You are on page 1of 25

SUTOPO and BRUCE WALDRIP

IMPACT OF A REPRESENTATIONAL APPROACH ON STUDENTS’


REASONING AND CONCEPTUAL UNDERSTANDING
IN LEARNING MECHANICS
Received: 27 November 2012; Accepted: 01 July 2013

ABSTRACT. The aim of this study was to explore whether a representational approach
could impact on the scores that measure students’ understanding of mechanics and their
ability to reason. The sample consisted of 24 students who were undergraduate, preservice
physics teachers in the State University of Malang, Indonesia. The students were asked to
represent a claim, provide evidence for it, and then, after further representational
manipulations, refinement, discussion, and critical thought, to reflect on and confirm or
modify their original case. Data analysis was based on the pretest–posttest scores and
students’ responses to relevant phenomena during the course. The results showed that
students’ reasoning ability significantly improved with a d-effect size of 2.58 for the
technical aspects and 2.51 for the conceptual validity aspects, with the average normalized
gain being 0.62 (upper–medium) for the two aspects. Students’ conceptual understanding
of mechanics significantly improved with a d-effect size of about 2.50 and an average
normalized gain of 0.63. Students’ competence in mechanics shifted significantly from an
under competent level to mastery level. This paper addresses statistically previously
untested issues in learning mechanics through a representational approach and does this in
a culture that is quite different from what has been researched so far using student-
generated representational learning as a reasoning tool for understanding and reasoning.

KEY WORDS: reasoning ability, representational approach, understanding of mechanics

INTRODUCTION

There has been widespread interest and concern amongst physics teachers
about student understanding of Newtonian mechanics. For example, some
research has investigated student difficulties in interpreting velocity and
acceleration (McDermott, Rosenquist & van Zee, 1987; Trowbridge &
McDermott, 1980, 1981), while other studies have utilized multiple-
choice tests such as the Force Concept Inventory (Hestenes, Wells &
Schwackhammer, 1992) and the Force and Motion Conceptual Evaluation
(Sokoloff & Thornton, 1998) that incorporate easily understood and non-
technical questions to investigate student knowledge. Some researchers have
shown that some students have difficulty with the concept of force
(Champagne, Klopfer & Anderson, 1980; Rosenblatt & Heckler, 2011). The
previous research has focused on upper high school or university students.

International Journal of Science and Mathematics Education (2014) 12: 741Y765


# National Science Council, Taiwan 2013
742 SUTOPO AND BRUCE WALDRIP

In addition, previous research tended to focus on identifying perceived


student misconceptions in mechanics (Clement, 1982; White & Gunstone,
1992). This research has stimulated strong interest in what is and how to
measure conceptual change (McDermott, 1984). This research facilitated
exploration of a change in teaching strategies. Some other researchers explored
the means of creating good classroom practice (Dykstra, Boyle & Monarch,
1992, Glynn & Duit, 1995; Hake, 1987; Visniadou, Ioannides,
Dimitrakopoulou & Papademetriou, 2001). Finally, more recent research has
shown strong interest in researching students’ reasoning and use of argumen-
tation in the science classroom (Alozie, Moje & Krajcik, 2010: 395; Osborne,
2010), with much of this research focused on strategies to replace traditional
transmission approaches with “more dialogic classroom interactions.”

Characterizing Reasoning
It is reasonable to claim that a person’s ability to explain can indicate their
level of understanding. The Science Framework (NRC, 2012) states that
there is a need to justify one’s understandings with peers to ensure that
they are adequate and implies that this process is interactive.

In science, reasoning and argument are essential for identifying the strengths and
weaknesses of a line of reasoning and for finding the best explanation for a natural
phenomenon. Scientists must defend their explanations, formulate evidence based on a
solid foundation of data, examine their own understanding in light of the evidence and
comments offered by others, and collaborate with peers in searching for the best
explanation for the phenomenon being investigated. (NRC, 2012: 67).

Reasoning ability is a prerequisite for developing improved scientific


explanations. Toulmin’s (1958) formal model of syllogistic reasoning
processes has been highly influential in recent research on student reasoning
in science, focusing on the identification of claims, assertions, connecting
evidence or data, warrants for claims, and backings for warrants (Brown,
Furtak, Timms, Nagashima & Wilson, 2010; Osborne, 2010; Jimenez-
Aleixandre & Erduran, 2007). There is a need for students to be able to give a
clear account of a representation as a claim, but this claim and its supporting
“evidence” will be embedded in complex ways in their representation, rather
than entail a predominantly linguistic account, as in the Toulmin model.
Dolan and Grady (2010: 40) have noted the complexity of reasoning
processes, claiming that the highest level of reasoning in inquiry entailed
students representing data in “multiple ways including tables, drawings,
graphs, or statistical representations, thoughtfully considering the
meaning of representations.” They noted that the most complex form of
reasoning by students arose when they communicated results and were
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 743

“using logical arguments to defend their findings” (p. 41). As noted by


Cox (1999), representations can serve many purposes. For example, they
can be used as tools for initial, speculative thinking, as in constructing a
diagram or model to imagine how a process might work.
From a technical aspect point of view, reasoning includes constructing a
claim (a statement or conclusion that addresses the question or problem),
evidence (data, theory, principles, etc., that supports the claim), and
argumentation (reasons that justify the connections between evidence and
the claim) (McNeill & Krajcik, 2008). The conceptual validity employed in
reasoning is another aspect of reasoning. The interplay between the two
aspects of reasoning plays a key role in constructing scientific explanation.
These reasoning processes also align with the TIMSS framework for
characterizing reasoning in school science (Mullis, Martin, Ruddock,
O’Sullivan & Preuschoff 2011). These researchers noted that students at
this level are expected to solve multi-step science problems in unfamiliar,
complex contexts. Mullis & colleagues (2011: 90–92) noted that students
are expected to integrate and synthesize understandings, evaluate claims,
and justify explanations through constructing a case to support the
reasonableness of their proposed explanations of topics.
A considerable amount of research has evaluated students’ reasoning
processes against formal, logical models, but there has been far less focus
on the reasoning strategies and processes students actually use when they
generate and critique their own and others’ representations. That is,
students using representations as a reasoning tool is a largely under
researched aspect.

A Representational Approach to Learning Physics


Recently researchers in science education argued that to learn science
effectively, students need to understand the different representations of
science concepts and processes, be able to translate a representation into
another one, and understand their coordinated use in representing scientific
knowledge (Hubber, Tytler & Haslam, 2010; Prain, Tytler & Peterson,
2009). The ability to use representations is considered as being a key for
learning physics (Kohl, Rosengrant & Finkelstein, 2007). Students with
higher representational ability have a higher chance of solving complex
problems successfully (Malone, 2008). Rosengrant, Heuvelen & Etkina
(2009) found evidence that students who frequently use representations are
successful in the Force Concept Inventory and the Mechanics Baseline Tests.
Ainsworth (2008, p. 197) argued that multiple representations play three
major functions in learning: (1) a complementary function, in that one
744 SUTOPO AND BRUCE WALDRIP

representation complements another representation as every representation


may differ in the information it expresses or in the processes it supports; (2)
constraining interpretations of a difficult representation (because of its
complexity or abstractness) using an easier representation (because of its
familiarity or concreteness); and (3) constructing a deeper understanding as
students could integrate information from more than one representation.
The emerging area of research into representation investigates the
value of student-generated representations to promote understanding in
science (Waldrip, Prain & Carolan, 2010). This approach is adopted in
this study even though it does include some representations provided by
lecturers. In this perspective, students learn to use material and symbolic
tools to think scientifically, incorporating both new and old technologies
(Cox, 1999; diSessa, 2004; Greeno & Hall, 1997; Waldrip & Prain,
2006). We agree with Kozma & Russell (2005) that representations can
function as conceptualizing and reasoning tools, rather than just as means
for knowledge display. It implies the need for learners to use their own
representations. diSessa (2004) argued that students can productively
construct new representations, if given enough time and support, even
approaching qualities of scientific representation such as precision (clear
or unambiguous), conciseness (give minimal but sufficient information),
and completeness (comprehensive for its purpose). It is sensible to
assume that a teaching approach that integrates reasoning and both
sanctioned and student-generated representations may be able to promote
students’ learning. In this respect, a representational approach to learning
means students being able to state claims, reflect on what is appropriate
evidence, as well as critique and modify the representation and then refine
both the initial reasoning and representation.
In this paper, we have utilized an approach that conceptualizes facilitation
of student learning through a teacher-guided focus on students’ representa-
tional reasoning (Waldrip et al., 2010), drawing on Roberts’ (1996, p. 423)
“trialogue” model. This model focuses on student learning as the
development of three-way reciprocal linkages between teachers’ and
students’ representations and domain knowledge. In past studies (Carolan,
Prain & Waldrip, 2008; Tytler, Haslam, Hubber & Prain, 2009; Waldrip &
Prain, 2006), we have found that if the teacher focuses on students’ thinking
and reasoning as they attempt to represent concepts and processes in a
sequence of representational challenges, this guided inquiry facilitates
quality conceptual learning. In this approach, the teacher needs to recognize
that students’ understandings often diverge from the teacher’s expert domain
knowledge and that there is a need to make explicit students’ reasoning
around representational adequacy to facilitate learning. We also recognize
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 745

that it is important to explore students’ initial understandings through in-


class discussion between students and also with their teacher and then build
toward reconciliation with more authorized views. We consider that students
benefit by generating their own understandings through making represen-
tations of their emerging conceptual knowledge, including drawing,
modeling, discussions, tables, graphs, multi-media products, role plays,
and photographs, in a process of guided inquiry (Ehrlén, 2009; Creagh,
2008). Where appropriate, they can also use and integrate tables, graphs,
multi-media products, role-plays, and photographs, to represent and justify
new understandings.

Aim of Study
The aim of this study is to explore whether a representational approach
can impact the learning of mechanics and students’ ability to reason. That
is, the paper addresses lecturers’ perceived concerns about the inadequate
preparation of physics content knowledge. It adopts a largely student-
generated representational approach where students need to display their
understanding of mechanics through negotiating, validating, and reason-
ing (Waldrip, Prain & Sellings, 2013). The focus of this paper is restricted
to the achievement measures in the pre- and post-test and relates them to
levels of reasoning skills and mastery of mechanics concepts, rather than
providing an in-depth description of the implementation of the represen-
tational approach, especially dealing with the student–student talks during
their group discussions. That is, to what extent did a representational
approach to teaching and learning mechanics impact on students’ ability
to reason and their achievement scores?

METHOD

Context of Study
This study was implemented in a preservice physics teacher program in
Indonesia. Indonesian physics teachers are prepared through an undergrad-
uate, 4-year program. The curriculum is concurrent in that students learn
both content knowledge and pedagogical knowledge concurrently through-
out the program. For preparing students with deep physics content
knowledge, the first year students learn introductory physics that covers all
topics typically covered in university physics textbooks (e.g. Halliday,
Resnick & Walker, 2011; Serway & Jewett, 2010; Young, Freedman &
Ford, 2008). In the following years, students study other advanced physics.
746 SUTOPO AND BRUCE WALDRIP

Students learn physics through practical and theoretical classes. The current
curriculum provides a course designed to help students to grasp the essential
topics closely related to the school physics curriculum. Some universities
name the course the Selected Topic of School Physics (STSP). Historically,
this course was developed to address students’ markedly deficient
understanding of essential concepts of school physics when they com-
menced their teacher preparation program. A deep content knowledge and
good reasoning ability are critical competencies for becoming an effective
physics teacher in Indonesia.
This research implemented a representational approach to learning
mechanics on the STSP course, in the Physics Education Department, at
the State University of Malang, Indonesia. The physics topics covered in
this course include mechanics, sound waves, thermodynamics, and
electromagnetism. However, this research focused on students’ learning
of mechanics. Students enrolling this course have previously learned
mechanics through introductory physics and mechanics courses. The
study was implemented from August 2011 to December 2011 and
involved 24 students. Nineteen students were female, indicating female
domination in Indonesian preservice physics teacher education programs.
The teaching was conducted in English and Bahasa Indonesia, the
national language of Indonesia, as the students involved in this study were
being prepared to teach physics in National-Plus schools where science
and mathematics should be taught in English. Both the lecturer and the
students were expected to use English as much as possible in order to
improve students’ English skill. All teaching materials, including the test,
were provided in English. However, the lecturer allowed the students to
use Indonesian during small group discussions as well as in responding to
the test, both in the pretest and posttest.
The “generic outline” of the teaching and learning process can be
described as follows: (1) The lecturer exposes an open-ended, multifaceted
problem (see Fig. 1 as an example) and asks the students to solve the problem
using coherent multiple representations including tables, graphs, diagrams,
equations, and words. (2) Through collaborative work in a group of three to
four students, the students should construct and critique their representations
by arguing, writing, collecting data, drawing, modeling, and graphing in order
to solve the problem using coherent multiple representations. They also need
to prepare their best presentation and be ready to defend their work in the next
whole class discussion. (3) During this group discussion, the lecturer moves
around the groups and gives necessary prompts or assistance according to the
need of each group depending on: (a) Whenever students have a high degree
of certainty about their representations, the lecturer prompts them to justify
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 747

Figure 1. A multiflash motion diagram for which students need to describe how
position, velocity, and acceleration change with time, as well as the net force acting on the
object. The images are taken successively every second

their reasoning through clarification. (b) Whenever students are uncertain


about their represented claim, or face deadlock in a discussion, the lecturer
provides them with the necessary scaffolding to prompt further reasoning. (c)
Whenever most groups do not have the necessary skill or knowledge to
construct the appropriate representation, or have no idea of how to critique the
appropriateness of their representation, the lecturer provides the necessary
scaffold(s) through class discussion. (4) After group discussion, the students
share their work with others through a whole class discussion. The lecturer
facilitates this discussion to consolidate students’ understanding.
Students’ activities in solving such problems as presented in Fig. 1 can be
outlined as in Fig. 2. Through collaborative work in their own group, students
should describe the motion using coherent multiple representations. They
need to work inductively based on the given multiflash diagram by generating
and tabulating data, graphing their resulting data, and formulating an equation
that best fits the resulting graph. Parallel to this activity, students need to
analyze the given multiflash diagram using another approach, i.e., vector
diagram. Based on the resulting tables, graphs, equations, and diagrams,
students describe the motion using concise sentences. Throughout these
processes, students are encouraged to check and critique the appropriateness
of their resulting representations as well as self-consistency among these
representations. For example, as shown by the two-headed dashed arrow in
Fig. 2, students are encouraged to check if their resulting equation of velocity,
for instance, matches with the first derivative of the resulting equation of
position with respect to time.
The mechanics topics discussed throughout the course included linear
motion, parabolic motion, circular motion, a simple harmonic oscillator, and
748 SUTOPO AND BRUCE WALDRIP

Given multiflash motion diagram

Generating & Graphing Formulating Drawing Constructing


tabulating data position equation of position verbal description
of position versus position as of how position
vectors
versus time time function of time changes with time

Generating & Graphing Formulating Drawing Constructing


tabulating data velocity equation of velocity verbal description
of velocity versus velocity as of how velocity
vectors
versus time time function of time changes with time

Generating & Graphing Formulating Drawing Constructing


tabulating data equation of verbal description
acceleration acceleration
of acceleration acceleration as of how acceleration
versus time vectors
versus time function of time changes with time

Inferring & describing net force


using multiple representations

Comprehensive description about the motion using coherent multiple representations

Figure 2. Students’ representational work in describing a given multiflash motion


diagram. A solid arrow indicates a typical working sequence; a two-headed dashed arrow
indicates an activity to check the consistency among representations

the physics of a roller coaster. It took 7 weeks, with one 3-h session a week, to
finish these topics. The representational task dealing with the first four topics
was similar, i.e., students constructed a comprehensive description of motion
using coherent multiple representations based on the given multiflash motion
diagram as described in advance. Students’ learning activities dealing with
the problem of “the physics of a roller coaster” were quite different. This
problem was designed to facilitate students in employing their skills and
knowledge that they have already acquired throughout solving the previous
problems. To solve this problem, students needed to apply their knowledge
about linear motion, parabolic motion, and circular motion in terms of both
their kinematics and dynamics aspects, as well as the basic concepts in
mechanics such as mechanical energy conservation and Newton’s laws of
motion. Students needed to have mathematical skills such as geometry,
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 749

trigonometry, vector manipulations, and numerical method as well. Because


of the complexity of the problem, the lecturer’s intervention occurs much
more often than before. However, the main characteristics of students’
learning, i.e., to construct multiple representations and use their representa-
tions to grasp a deep understanding of physics ideas through activities that
include arguing, writing, drawing, modeling, and graphing, are still
maintained.
We attribute the teaching approach as being a representational approach. It
adopts a largely student-generated representational approach where students
need to display their understanding of mechanics through a negotiating,
validating, and reasoning (Waldrip et al. 2013). We argue that the approach
could improve both students’ understanding of mechanics concepts and
ability to reason.

Research Design and Data Collection


This study implemented a mixed-methods, embedded experimental design
(Creswell & Clark, 2007). The same instrument was used as a pretest and
posttest. The posttest was conducted at the end of semester instead of upon
the finishing of the topics for mechanics.
The main instrument was an integrated test adapted from the
Mechanics Baseline Test (MBT) developed by Hestenes & Wells
(1992). The original MBT instrument consisted of 26 multiple-choice
items designed to assess student’s conceptual understanding of mechanics
“that cannot be grasped without formal knowledge about mechanics” and
“require algebraic manipulation or more than one-step reasoning”
(Hestenes & Wells: 159–161). According to Hake (1998: 65), the “MB
test requires conceptual understanding in addition to some mathematical
skill and critical thinking.” The instrument was used to simultaneously
assess students’ conceptual understanding of mechanics and reasoning
ability. For this purpose, students not only chose one alternative that best
represents their response, but also wrote explanations to justify their
answer. We measured student conceptual understanding of mechanics
based on their multiple choice responses and reasoning ability based on
their open explanations. We addressed students’ reasoning ability from
two aspects, i.e., the technical and conceptual validity. The technical
aspect refers to the fulfillment of Toulmin’s reasoning components. The
corresponding rubric (Table 1) was adapted from the work of Furtak,
Hardy, Beinbrech, Shavelson & Shemwell (2010). Conceptual validity
refers to the correctness and appropriateness of the physics concepts,
theories, principles, or laws employed in reasoning. The corresponding
750 SUTOPO AND BRUCE WALDRIP

TABLE 1
Rubric to code the technical aspect of student’s reasoning quality, adapted from Furtak et
al. (2010)

Category Score=level Description/indicator More detailed description

Inductive/ 4 Claim is backed up by The rationale consists of a


deductive generalized relationship, comprehensive data analyses
rule-based principle, theory, or law. supported by a principle,
reasoning theory, law, or definition that
are relevant to the data and
problem being solved. The
scientific correctness of the
theory, law, etc., used in this
reason is not important.
Evidence- 3 Claim is backed up by The rationale has considered
based evidence, including an amount of data (including
reasoning analogy and analysis implicit data) and applied a
of data. relevant data analyses, but not
enough to solve the problem
correctly.
Data-based 2 Claim is backed up by The rationale relies on limited
reasoning data. data or the surface feature of
the problem.
No reasoning 1 Claim is not backed up, The rationale, if any, is merely
or the backup is pseudo, a restatement of the claim
or tautology. (response) or not clearly
related to the problem nor
clear in meaning.
Unidentified 0 No response Student’s answer sheet is blank

rubric (Table 2) was constructed by combining the “validity outcome


space” and “conceptual sophistication outcome space” of reasoning
proficiencies proposed by Brown et al. (2010: 155–156).
It is noted that the lessons did not cover all the concepts being tested by
MBT (see Table 3). The concepts of MBT that were not mentioned
throughout the lessons included integrated displacement, work-energy
theorem, impulse–momentum, momentum conservation, friction, and
superposition principle. In addition, the lecturer tried to avoid “teaching
the test” throughout the lessons.
Based on the Indonesian pilot study in establishing the instrument with 52
students in a previous semester, the final instrument consisted of 22 multiple
choices items with the following characteristics. The coefficients of correlation
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 751

TABLE 2
Rubric to code the conceptual validity aspect of student’s reasoning

Category Score/level Definition

Fully valid 3 Claim is correct and follows from the relevant and correct
backup
Partially 2 • Claim is correct but the backup is not fully appropriate
valid (incomplete or partially irrelevant), or
• Claim is incorrect since it follows from an inappropriate
backup
Invalid 1 • Claim is incorrect since it follows a fully incorrect
backup or does not logically follow from backup, or
• Claim is correct but fully follows incorrect backup
Unidentified 0 • No rationale, or the rational is tautological

between item score with total score varied from 0.32 (pG0.05) to 0.62 (pG
0.01); the item’s discrimination index varied from 0.29 (moderate) to 0.86
(very high), and the item difficulty index varied from 0.12 (difficult) to 0.63
(moderate). This pilot study corroborated the claim of Hestenes & Wells
(1992) that the test was not easy. The Cronbach’s Alpha coefficient of this
instrument was 0.81, indicating a very good reliability (Everitt & Skrondal,
2010). For checking the coding reliability, 25 % of the reasoning units were
coded by the primary rater (researcher) and one secondary independent rater.
There were 528 reasoning units (i.e. 24 students×22 items) for each data set.
The percentage agreement of the two raters was 78 % for the technical aspect
and 87 % for the conceptual validity aspect; yielding Cohen’s kappa of 0.71
and 0.80, respectively.
Quantitative analysis included Cohen’s d-effect size (Ellis, 2010) and
normalized gain (N-gain) (Hake, 1998). The average N-gain is defined as
〈g〉 ¼ 〈posttest〉−〈pretest〉
100−pretest , where the maximum score has been scaled to 100. In
this study, the categorizations used were: low if 〈g〉 G 0.25, lower–medium if
0.25 ≤ 〈g〉 G 0.45, upper–medium if 0.45 ≤ 〈g〉 G 0.65, and high if 〈g〉 ≥ 0.65.

RESULTS AND DISCUSSION

Improvement of Students’ Understanding of Mechanics Concepts


Students’ scores on the pretest and posttest, as well as their normalized
gains, are summarized in Fig. 3. These tests measured whether they could
752 SUTOPO AND BRUCE WALDRIP

provide an adequate response but did not explore why this response was
correct.
The mean of the posttest was greater than that of the pretest (Fig. 3, right).
A paired-samples ttest yielded t = 14.12, p = 0.00 (two-tailed), suggesting
that the representational approach implemented in this research improved
students’ understanding of mechanics concepts. The improvement was
strong as its d-effect size was 2.5, in the category of “much larger than
typical” (Morgan, Leech, Gloeckner & Barrett, 2004, p.91), and Hake’s
average N-gain was 0.63 (upper–medium category). Students with lower

TABLE 3
Mechanics concepts covered by the MBT and the lesson

Topics/concepts involved on MBT Topics/concepts discussed in the lessons

A. Kinematics A. Linear motion


Linear motion: constant acceleration, Kinematics: position, velocity and
average acceleration, average velocity, acceleration (instantaneous and average)
and integrated displacement Newton’s second law
Curvilinear motion: tangential B. Parabolic motion
acceleration, normal acceleration, and Kinematics: position, velocity and
centripetal acceleration (a = v2/r) acceleration (instantaneous and average),
tangential acceleration, normal acceleration
Newton’s second law
B. General principles C. Circular motion with constant speed
Newton’s laws of motion, superposition Kinematics: position, velocity and
principle, work-energy, energy acceleration (instantaneous and average),
conservation, impulse-momentum, and angular velocity and acceleration,
momentum conservation centripetal acceleration (a = v2/r)
Newton’s second law
C. Specific forces D. Harmonic oscillator
Gravitational free-fall and friction Kinematics: position, velocity and
acceleration (instantaneous and average),
displacement, amplitude, period, and
frequency
Hooke’s law, Newton’s third law, and
Newton’s second law
E. The physics of a roller coaster
Kinematics of linear, parabolic, and circular
motion
Newton’s third law, normal force, weight
sensation, weightless sensation,
gravitational free-fall, energy conservation
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 753

Figure 3. Scatter plots of N-gain versus the pretest (left) and posttest versus pretest
(right). The mean of the pretest (dashed vertical line), posttest (horizontal line in figure
right), and N-gain (horizontal line in figure left) was 39.2 (SD = 15.3), 77.5 (SD = 13.0),
and 0.64 (SD = 0.17) respectively. The bullets represent single data, while the squares
represent double data

(less than the average) pretest score tended to get a lower gain, and vice versa
(Fig. 3, left). This suggests that most students took proportional advantage
from the teaching-learning process to improve their understanding of
mechanics concepts.
It is useful to examine the improvement of students’ understanding in
terms of their competence in mechanics. Hestenes & Wells (1992) argued that
a score of 60 % on the MBT is the threshold for problem-solving competence.
Below this threshold, students’ grasp of Newtonian concepts is too limited for
effective problem solving. A score of 80 % is the threshold for mastery of
basic Newtonian concepts. They believed that when this is approached, other
goals of physics instruction would be much easier to attain. We define
students’ mechanics competence levels as follows: “under-competent” if the
MBT score is G60.0 %, “competent in problem solving” if there is a 60.0 %≤
MBT score G80.0 %, and “mastery in basic Newtonian mechanics” if the
MBT score is ≥80.0 %. Figure 3 (right) shows that the approach successfully
assisted most students to jump from an “under-competent” level to the
“mastery” level. Moreover, there were no students in the mastery level before
instruction, whereas 50 % of students were in this level after instruction.
The low result of the students’ mechanics competence before instruction
was not expected as they had previously taken first semester courses in
Mechanics and Introductory Physics. This phenomenon was similar to our
previous study (Sutopo, Waldrip & Rusdiana, 2011) involving 35 students
who had taken Introductory Physics, plus 24 students who had taken both
754 SUTOPO AND BRUCE WALDRIP

Introductory Physics and Mechanics. The average score was 33 % (SD =


17 %) with a maximum score of 68 %. These findings indicate that student
deficiency in understanding Newtonian mechanics had persisted beyond the
school physics and could not be adequately addressed throughout the two
university courses.
To understand the effectiveness of this approach, it is useful to elaborate on
the qualitative findings drawn from students’ open explanations on the pre-
and post-test, as well as from phenomena encountered throughout the lessons.
Based on students’ responses on the pretest, there was evidence that most
students held some misconceptions about acceleration, including: (1)
acceleration is always in the same direction as the motion or velocity; (2)
the magnitude of acceleration is proportional to the magnitude of velocity; (3)
acceleration is in the direction to which the object tends to move; (4) in any
frictionless track, the object’s acceleration is zero; and (5) if there is no friction
and other external applied force, the acceleration of an object moving under
the influence of gravity is equal to the gravitational acceleration g, even though
there exists such a “passive force” as normal force or tension in a rope.
Students also misunderstood the relationship between acceleration and force
as stated in Newton’s second law of motion, F = ma. They considered this
equation as being causality rather than a covariant relationship.
Based on the phenomena encountered throughout the first few lessons,
there was evidence that students held some prior deficient understandings.
They: (1) tended to define velocity as “distance divided by time,” instead of
“displacement divided by the time elapsed” or “the change of position divided
by time”; (2) could not apply the operational definition of velocity
v≡limΔt→0 Δx Δv
Δt ¼ dt and acceleration a≡limΔt→0 Δt ¼ dt to analyze a motion
dx dv

diagram; (3) could not define the change of position (Δx) even though they
had been able to make vector operations such as (A+B) and (A−B) where A
and B are vectors unconnected to any physical quantity; and (4) believed that
acceleration was always in the same direction as the velocity, the minus sign
on acceleration meant deceleration, and acceleration should be inferred from
the resultant force; but not in reverse. It can be believed that these deficiencies
in understanding have caused the low student scores on the pretest.
The students worked in small groups where they developed solutions
as a group and developed a common response as to why their solution
was adequate. During this process, students made and justified a claim as
to the solution, through student–student and a group–teacher dialogue.
The following transcript shows an example of a dialogue in a group when
the group attempted to describe how velocity changes with time. This
dialogue happened when the group had described how position changes with
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 755

time. Students discussed the problem (lines 1–9) and faced a deadlock (line
9). The lecturer’s prompt (line 10) inspired students to propose a new idea
(line 11). Afterward, the lecturer required students to construct another
representation (line 14).
1. S1: “Now we should describe how the position changes with time”.
2. S2: “From diagram we can know the distance traveled for some
interval of time. For example at t = 2 s, the distance is 36 m then v =
18 m/s. At t = 4 s …
3. S1: “Is 63 m then v = 15.75 m/s?”
4. S2: “At t = 6 s, the distance is 80 m then v = 13.33 m/s ....”
5. S3: “Yah, It means that the velocity is decreasing …”
6. S1: “Wait! Why do we always start from t zero? … how about if we
calculate the velocity at every interval of 2 seconds, from t = 0 to 2 s,
then from t = 2 s to 4 s, etc.?”
7. S2: “Could be, let’s try. … For the first interval, v=36/2=18 m/s.
For second interval, v = (63−36)/2 = 13.5 m/s, … for the third … v =
8.5 m/s … Yah, it’s decreasing too”.
8. S1: “Yah, but the result is different from the previous one. Look, v3 is
very different! Our previous result was 13.3 but now is 8.5. …”
9. S3: “So how?” [Discussion faced deadlock]
10. L: “What are you talking about? You are really talking about speed
rather than velocity. Moreover, you are talking about average speed,
instead of instantaneous speed. Your task is to describe how velocity
changes with time. It means you need to talk about instantaneous
velocity. For example, what’s the velocity at t = 2 s, at t = 3 s, and so
on?”
11. S1: “I have an idea. We’ve had an equation of position x (t) =
−1.2 t2 + 20.4 t. To get velocity, we need to simply take its derivative
v=dx/dt … the result is v(t) = −2.4 t + 20.4.”
12. S2: “Yah, this equation describes how the velocity changes with
time.”
13. S3: “We can also use this equation to define the velocity at any
instant of time. For example, at t = 0, v = 20.4 m/s; at t = 2, v = −4.8 +
20.4 = 16.4 m/s, etc.”
14. L: “Good job! So, you’ve gotten a mathematical representation of the
velocity. Now, please describe it using a diagram, that is, draw the
velocity vector at any instant of time”

Throughout the lessons, there was evidence that most students had
improved their conceptions and developed a better understanding in the
areas outlined above. It can be argued that the representational approach
756 SUTOPO AND BRUCE WALDRIP

implemented in this study could remediate students’ misconceptions, or


common sense, about Newtonian mechanics concepts, even though the
lessons were not designed initially to assess and remediate these
misconceptions.
We argued that student success in grasping a better understanding of
the operational definition of velocity and acceleration was the key for
their success in the posttest, especially in dealing with kinematics
problems. Therefore, it is useful to describe how students reached this
understanding. Students’ deficiency in understanding appeared directly in
the first meeting when they solved the problem of linear motion. They
had no idea of how to define instantaneous velocity and acceleration from
the given multiflash motion diagram. They tried to find out by opening
some physics textbooks such as Serway & Jewett (2010), Halliday et al.
(2011), and Young et al. (2008). Unfortunately, they just found the
Δx
definition of instantaneous velocity of v ¼ dx dt ¼ limΔt→0 Δt and its
difference from average velocity. They did not get any clues to using
this equation to solve the problem. Because all the groups faced the same
difficulty, the lecturer facilitated the students to be able to develop an
alternative method. He used a basic concept of statistics, i.e., the average
value of a data set is usually equal to the middle score, as the entry point.
1. L: “Look at the series of numbers: 1, 3, 5, 7, 9. What is the average
value of this data set?”
2. S: “5”

5. L: “Now, assume that the position of a moving particle is x(t) = c0 +
c1t + c2t2 where c0,c1,and c2 are constants. What is the particle’s
average velocity between t = 0 and t = t0? … What is the particle’s
velocity at t ¼ t0 =2 , that is v (t 0 =2 )?”
6. S (After discussion): “Average velocity is v½0;t0 Š ¼ xðt0t0Þ−x ð0Þ
−0 …  ¼ 
c1 þ
c2 t 0 …Instantaneous velocity is vðtÞ ¼ c1 þ 2c2 t::::→v t 0 =2 ¼
c1 þ c2 t 0 ”
7. L: “Good … What can you say about these values?”
8. S: “Exactly the same …. The average velocity during the interval
time of t0 is equal to the instantaneous velocity at t ¼ t 0 =2 .”
9. L: “Okay, now replace position with velocity, and velocity with
acceleration. Do this with the same mathematical model. … What do
you get?”
10. S (After discussion): “Average acceleration during the interval time
of t0 is equal to the instantaneous acceleration at t ¼ t 0 =2 .”
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 757

11. L: “Excellent. But, be careful to avoid overgeneralization. Now, use


x(t)=c0 +c1t+c2t2 +c3t3 as your model. What will you get?”
This lecturer–students dialogue led students to formulate the procedure
that instantaneous velocity (acceleration) at the middle interval of time is
equal to the average velocity (acceleration) during the interval. Students
successfully applied the procedure to analyze the given multiflash
diagram of parabolic motion, circular motion, and the harmonic oscillator.
Figure 4 is an example of the multiple representations constructed and
presented (using power point) by a group.
Based on the students’ responses to the questionnaire, there was
evidence that students have obtained considerable benefits from their
learning. All groups asserted that their understanding of the concepts
being discussed improved. The gaining of conceptual understanding
included the strengthening or deepening of the existing understanding,
remediating misconceptions, constructing a new concept, and identifying
key concepts involved in the physics problems. It also: (1) improved their
problem solving skills, including the confidence to use their own ideas in
solving problems, instead of relying on the textbook; (2) improved their
skills and confidence in communication, discussion, and collaboration;
(3) improved their ability to verify the validity of an opinion by checking
its consistency with other representations; (4) increased their ability to
describe physics phenomena scientifically and with evidence; and (5)
improved skills in using computer software to make graphs, calculations,
and data analysis.

By trying to describe our idea using a variety of representations, we can more deeply understand
the concepts being discussed … are able to describe physics phenomena scientifically (using the
relevant physics ideas appropriately) and convincingly (using a variety of coherent
representations). .... Now we are able to dare to not rely on the textbooks in solving physics
problem; instead we need to deepen our understanding of the essential concepts and use them
consistently. … Learning physics becomes enjoyable as we can find out the key concepts
involved in a problem and successfully solve the problem; we engage in the finding of the key
concepts throughout the lessons. .... We improve our skills in using computer software such as
Microsoft Excel to make graphs, calculations, and data analysis. We improve our skills of
discussion and collaboration in a group.

The effectiveness of this approach confirms some previous research


about representations. Some researchers argue that to learn science
effectively, students need to understand the different representations of
science concepts and processes, be able to translate representations into
others, and understand their coordinated use in representing scientific
knowledge (Hubber et al., 2010; Prain et al., 2009). Similarly, Carolan et
al. (2008) found that the teaching of science in which the teacher focuses
on students’ thinking and reasoning enabled students to grasp deep
758 SUTOPO AND BRUCE WALDRIP

Figure 4. A slide of a student power point presentation: Multiple representations


constructed by a group to describe the motion diagram shown in Fig. 1

understandings of science concepts. This present research also


confirms the “construct deeper understanding” function of multiple
representations in learning science (Ainsworth, 2008). However, in
this present research, students did not just respond to the represen-
tations provided by the lecturer or available in textbooks. Instead,
students constructed their own representations and used them to form a
meaningful understanding of the concept being represented. According to
Ainsworth, Prain & Tytler (2011), engaging students in constructing their
own representations will deepen students’ conceptual understanding. The
finding of this research also confirms the claim of Rosengrant et al.
(2009) that students who frequently use representations will be successful on
the MBT.

Improvement of Students’ Reasoning Ability


Students’ reasoning scores on the pretest and posttest as well as their N-
gains are summarized in Table 4. Students were asked to justify their
responses to the conceptual test. The validity of the reasons rather than
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 759

TABLE 4
Descriptive statistics of students’ reasoning scores

Technical aspect (scale 0−4) Conceptual validity aspect (scale 0−3)

Statistics Pre Post N-gain Pre Post N-gain

Minimum 1.2 2.3 0.26 0.7 1.5 0.24


Maximum 3.0 4.0 0.98 2.1 3.0 1.00
Mean 1.99 3.23 0.63 1.32 2.37 0.64
(SD) (0.44) (0.51) (0.21) (0.41) (0.43) (0.20)
Skewness 0.11 −0.29 −0.15 0.10 −0.43 0.01

the correctness of their conceptual responses were the measure of these


scores. Theoretically, it was possible to achieve 100 % in the conceptual
test and fail the reasoning component. Mercier & Sperber (2011) stated
that reasoning allows one to evaluate the soundness of a claim and to
accept a view that would not otherwise be accepted.
A paired samples ttest yielded t=14.96, p=0.00 (two-tailed) for the
technical aspect and t=14.49, p=0.00 (two-tailed) for the conceptual
validity aspect, suggesting that the means difference between pairs of data
sets are statistically significant. The d-effect size on the technical and
conceptual validity aspect was 2.58 and 2.51, respectively. These effect
sizes are in the category of much larger than typical. The Hake’s average
N-gain was 0.62 (upper–medium category) for both the technical and
conceptual validity aspect. It can be argued that the representational
approach implemented in this research has high effectiveness in
improving students’ reasoning ability.
We also analyzed the change in the reasoning quality performed by
students. Students’ reasons were coded according to the level of reasoning
quality using the rubrics presented in Tables 1 and 2. The resulting data are
summarized in Tables 5 and 6. The Chi-square test resulted in χ2 =74.98 (p=
0.00) and χ2 =77.12 (p=0.00) for the technical and conceptual validity
aspects, respectively. This test suggested that the posttest distribution was
significantly different from the pretest distribution for both the technical and
conceptual validity aspects. For the technical aspect (Table 5), the proportion
of the highest level (Level 4) changed from about 19 % in the pretest to about
61 % in the posttest, whereas the lowest two levels decreased from about
44 % to about 12 %. For the conceptual validity aspect (Table 6), the
proportion of the highest level (Level 3) changed from about 17 % in the
760 SUTOPO AND BRUCE WALDRIP

TABLE 5
Posttest–pretest cross-tabulation of reasoning levels for the technical aspect

Posttest Total pretest

Lev-0 Lev-1 Lev-2 Lev-3 Lev-4 Count %a

Pretest Lev-0 2 2 4 3 9 20 3.8


Lev-1 7 37 19 24 126 213 40.3
Lev-2 1 11 31 25 77 145 27.5
Lev-3 0 2 11 10 27 50 9.5
Lev-4 0 0 3 13 84 100 18.9
Total posttest Count 10 52 68 75 323 528 100
%a 1.9 9.8 12.9 14.2 61.2 100
a
Relative to the total of reasoning units (528)

pretest to about 63 % in the posttest, whereas the lowest two levels decreased
from about 65 % to about 21 %.
The improvement in students’ abilities to reason was as expected according
to the anecdotal comments from the lecturer and students’ reflection on their
learning. In this respect, the lecturer argued that:

The approach encouraged students to think critically and reason logically throughout the
lessons. Through constructing representations to solve problems of discussion, students
learned to check the consistence of their representations with the available data and the
relevant theories, concepts, and principles underlying the problems, as well as to check
the coherence of their tentative representation with the previous representation(s) they
had already constructed. Through group and whole class discussions, students attempted
to make the best argumentation for defending their idea.

And students claimed that:

By trying to describe our idea using a variety of representations, we were able to describe
physics phenomena scientifically (using the relevant physics ideas appropriately) and
convincingly (using a variety of coherent representations).

It can be argued that the approach has provided students with ample
opportunities to develop their ability to construct the claims that were
backed up by appropriate theory and comprehensive data analysis. As a
result, students could develop both reasoning and representational skills
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 761

TABLE 6
Posttest–pretest cross-tabulation of reasoning levels for the conceptual validity aspect

Posttest Total pretest

Lev-0 Lev-1 Lev-2 Lev-3 Count %a

Pretest Lev-0 14 23 12 56 105 19.9


Lev-1 12 47 39 142 240 45.5
Lev-2 1 8 33 52 94 17.8
Lev-3 0 4 5 80 89 16.9
Total posttest Count 27 82 89 330 528 100
%a 5.1 15.5 16.9 62.5 100
a
Relative to the total of reasoning units (528)

concurrently as they tried to explain and justify their represented understand-


ing or idea.

CONCLUSION

The approach indeed improved students’ reasoning and understanding of


mechanics concepts. Students’ conceptual understanding was improved
with a very large effect size and upper-medium N-gain. Their competence
in mechanics significantly improved from an under competent level to
mastery level. Students’ ability to reason was improved significantly with
a very large effect size and upper-medium N-gain for both the technical
and conceptual validity aspect. The reasoning quality levels shifted from
the first two lowest levels to the highest level.
While this paper has focused on measures of reasoning skills rather
than explaining how this happens in detail, the reasoning shifts as
students use representations to develop their understanding in mechanics,
and the results of this paper support the claim that students can develop
both their reasoning and their representational skills concurrently as they
seek to explain and justify their represented understandings. Lecturers in
this process function as expert guides and respondents to students’
individual and group emerging accounts of topics.
762 SUTOPO AND BRUCE WALDRIP

In this paper, the students were asked to represent a claim, provide


evidence for it, and then after further representational manipulation,
refinement, discussion, and critical thought, to reflect on and confirm or
modify their original case. It is in line with the TIMSS framework for
characterizing reasoning processes in science (Mullis et al., 2011). Students
at this level are expected to solve multi-step science problems in unfamiliar,
complex contexts and are expected to integrate and synthesize understand-
ings, evaluate claims, and justify explanations through constructing a case to
support the reasonableness of their proposed accounts of topics. By using
these representations as contestable artifacts needing justification and
elaboration, students are practicing habits of mind and reasoning skills
central to scientific literacy. In this context, analogous to Mullis et al.’s
(2011) views of reasoning processes, students were invited to be assessors of
their own learning (mediators) and also to function as an audience and
sounding board for other students, thereby co-operatively fostering scientific
reasoning and literacy development aligned to scientific practice on a micro
learning-community scale. Importantly, the lecturer facilitates this guided
inquiry through critical feedback on the adequacy of student-generated
claims evident in their representations. As noted by Ford & Forman (2006),
unless school students learn to construct and interpret accounts of their
observations and reasoning and become active in the learning process, their
learning would be limited and superficial. This paper addresses statistically
previously untested issues in learning mechanics through a representational
approach and does this in a culture that is quite different to what has been so
far researched using student-generated representational learning as a
reasoning tool to gain understanding and allow reasoning.

REFERENCES

Ainsworth, S. (2008). The educational value of multiple representations when learning


complex scientific concepts. In J. K. Gilbert, M. Reiner & M. Nakhleh (Eds.), Visualization:
Theory and Practice in Science Education (pp. 191–208). New York: Springer.
Ainsworth, S., Prain, V. & Tytler, R. (2011). Drawing to learn in science. Science, 333,
1096–1097.
Alozie, N., Moje, E. & Krajcik, J. (2010). An analysis of the supports and constraints for
scientific discussion in high school project-based science. Science Education, 94(3),
395–427.
Brown, N., Furtak, E., Timms, M., Nagashima, S. & Wilson, M. (2010). The evidence-
based reasoning framework: Assessing scientific reasoning. Educational Assessment,
15(3), 123–141.
Carolan, J., Prain, V. & Waldrip, B. (2008). Using representation for teaching and
learning in science. Teaching Science, 54(1), 18–23.
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 763

Champagne, A. B., Klopfer, L. E. & Anderson, J. H. (1980). Factors influencing the


learning of classical mechanics. American Journal of Physics, 48(2), 1074–1079.
Clement, J. (1982). Students’ preconceptions in introductory mechanics. American Journal of
Physics, 50(1), 66–71.
Cox, R. (1999). Representation construction, externalized cognition and individual
differences. Learning and Instruction, 9, 343–363.
Creagh, C. (2008). Diagrams: Useful tools for investigating a student’s understanding of
buoyancy. Teaching Science, 54(4), 48–50. DEECD, (2011).
Creswell, J. W. & Clark, V. L. P. (2007). Designing and conducting mixed methods research.
Thousand Oaks, California: Sage Publications.
diSessa, A. A. (2004). Metarepresentation: Native competence and targets for instruction.
Cognition and Instruction, 22(3), 293–331.
Dolan, E. & Grady, J. (2010). Recognizing students’ scientific reasoning: A tool for
categorizing complexity of reasoning during teaching by inquiry. Journal of Science
Teacher Education, 21, 31–55.
Dykstra, D. I., Boyle, C. F. & Monarch, I. A. (1992). Studying conceptual change in
learning physics. Science Education, 76, 615–652.
Ehrlén, K. (2009). Drawings as representations of children’s conceptions. International
Journal of Science Education, 31(1), 41–57.
Ellis, P. D. (2010). The Essential Guide to Effect Sizes: Statistical Power, Meta-analysis,
and the Interpretation of Research Results. New York: Cambridge University
Press.
Everitt, B. S. & Skrondal, A. (2010). The Cambridge Dictionary of Statistics (4th ed.).
New York: Cambridge University Press.
Ford, M. & Forman, E. (2006). Redefining literacy learning in classroom contexts. Review
of Research in Education, 30, 1–32.
Furtak, E. M., Hardy, I., Beinbrech, C., Shavelson, R. J. & Shemwell, J. T. (2010). A
framework for analyzing evidence-based reasoning in science classroom discourse.
Educational Assessment, 15(3), 175–196.
Glynn, S. & Duit, R. (1995). Learning Science in the Schools: Research Reforming
Practice. Mahwah, NJ: Lawrence Erlbaum Associates.
Greeno, J. G. & Hall, R. P. (1997). Practicing representation: Learning with and about
representational forms. Phi Delta Kappa, 78(5), 361–368.
Hake, R. (1987). Promoting student crossover to the Newtonian world. American Journal
of Physics, 55(10), 878–884.
Hake, R. (1998). Interactive-engagement versus traditional methods: A six-thousand-student
survey of mechanics test data for introductory physics courses. American Journal of
Physics, 66(1), 64–74.
Halliday, D., Resnick, R. & Walker, J. (2011). Fundamentals of Physics (9th ed.). New
Jersey: John Wiley and Sons, Inc.
Hestenes, D. & Wells, M. (1992). A mechanics baseline test. The Physics Teacher, 30(3),
159–166.
Hestenes, D., Wells, M. & Schwackhammer, G. (1992). Force concept inventory. The
Physics Teacher, 30(3), 141–158.
Hubber, P., Tytler, R. & Haslam, F. (2010). Teaching and learning about force with a
representational focus: Pedagogy and teacher change. Research in Science Education, 40,
5–28.
764 SUTOPO AND BRUCE WALDRIP

Jimenez-Aleixandre, M. & Erduran, S. (2007). Argumentation in science education: An


overview. In S. Erduran, and M.P. Jiménez-Aleixandre (Eds.), Argumentation in
Science Education (pp. 3–28). Springer.
Kohl, P. B., Rosengrant, D. & Finkelstein, N. D. (2007). Strongly and weakly directed
approaches to teaching multiple representation use in physics. Physics Review Special
Topics - Physics Education Research, 3, 010108.
Kozma, R. & Russell, J. (2005). Students becoming chemists: Developing representational
competence. In J. Gilbert (Ed.), Visualization in Science Education (pp. 121–146).
London: Kluwer.
Malone, K. L. (2008). Correlations among knowledge structures, force concept inventory,
and problem-solving behaviors. Physics Review Special Topics-Physics Education
Research, 4, 020107.
McDermott, L. (1984). Research on conceptual understanding in mechanics. Physics
Today, July, 24–32.
McDermott, L. C., Rosenquist, M. L. & van Zee, E. H. (1987). Student difficulties in
connecting graphs and physics: Examples from kinematics. American Journal of
Physics, 55, 503–513.
McNeill, K. L. & Krajcik, J. (2008). Inquiry and scientific explanations: Helping students
use evidence and reasoning. In J. R. Luft, L. Bell & J. Gess-Newsome (Eds.), Science
as Inquiry in the Secondary Setting (pp. 121–133). Arlington, VA: NSTA Press.
Mercier, H. & Sperber, D. (2011). Why do humans reason? Arguments for an
argumentative theory. The Behavioral and Brain Sciences, 34, 57–111.
Morgan, G. A., Leech, N. L., Gloeckner, G. W. & Barrett, K. C. (2004). SPSS for
Introductory Statistics: Use and Interpretation (2nd ed.). New Jersey: Lawrence
Erlbaum Associates Inc.
Mullis, I., Martin, M., Ruddock, G., O’Sullivan, C., & Preuschoff, C. (2011). TIMSS 2011
Assessment Frameworks. TIMSS & PIRLS International Study Center, Boston College
National Research Council (2012). A Framework for K-12 Science Education: Practices,
Crosscutting Concepts, and Core Ideas. Washington D.C: National Academy Press.
Osborne, J. (2010). Arguing to learn in science: The role of collaborative, critical
discourse. Science, 328, 463–466.
Prain, V., Tytler, R. & Peterson, S. (2009). Multiple representation in learning about
evaporation. International Journal of Science Education, 31(6), 787–808.
Roberts, D. (1996). Epistemic authority for teacher knowledge: The potential role of
teacher communities: A response to Robert Orton. Curriculum Inquiry, 26, 417–431.
Rosenblatt, R. & Heckler, A. F. (2011). Systematic study of student understanding of the
relationships between the directions of force, velocity, and acceleration in one
dimension. Physical Review Special Topic - Physics Education Research, 7, 020112.
Rosengrant, D., Heuvelen, A. V. & Etkina, E. (2009). Do student use and understand free-
body diagrams? Physics Review Special Topics - Physics Education Research, 5, 010108.
Serway, R. A. & Jewett, J. W., Jr. (2010). Physics for Scientists and Engineers with
Modern Physics (8th ed.). Belmont, CA: Brooks/Cole.
Sokoloff, D. R. & Thornton, R. K. (1998). Assessing student learning of Newton’s laws:
The force and motion conceptual evaluation and the evaluation of active learning
laboratory and lecture curricula. American Journal of Physics, 66(4), 338–352.
Sutopo, Liliasari, Waldrip, B., & Rusdiana, D. (2011, November 12). The need of
representational approach to provide prospective physics teacher with better reasoning
IMPACT OF A REPRESENTATIONAL APPROACH ON SUDENTS 765

ability and conceptual understanding. Paper presented at 5th International Seminar of


Science Education, Indonesia University of Education, Bandung, Indonesia
Toulmin, S. (1958). The Uses of Argument. Cambridge, MA: Cambridge University Press.
Trowbridge, D. E. & McDermott, L. C. (1980). Investigation of student understanding of the
concept of velocity in one dimension. American Journal of Physics, 48(12), 1020–1028.
Trowbridge, D. E. & McDermott, L. C. (1981). Investigation of student understanding of the
concept of acceleration in one dimension. American Journal of Physics, 49(3), 242–253.
Tytler, R., Haslam, F., Hubber, P. & Prain, V. (2009). An explicit representational focus for
teaching and learning about animals in the environment. Teaching Science, 55(4), 21–27.
Visniadou, S., Ioannides, C., Dimitrakopoulou, A. & Papademetriou, E. (2001). Designing
learning environments to promote conceptual change in science. Learning and Instruction,
11, 381–419.
Waldrip, B. & Prain, V. (2006). Changing representations to learn primary science
concepts. Teaching Science, 54(4), 17–21.
Waldrip, B., Prain, V. & Sellings, P. (2013). Explaining Newton’s laws of motion: Using
student reasoning through representations to develop conceptual understanding.
Instructional Science. doi:10.1007/s11251-012-9223-8.
Waldrip, B., Prain, V. & Carolan, J. (2010). Using multi-modal representations to improve
learning in junior secondary science. Research in Science Education, 40, 65–80.
White, R. & Gunstone, R. (1992). Probing Understanding. London: The Falmer Press.
Young, H. D., Freedman, R. A. & Ford, A. L. (2008). Sears and Zemansky’s University
Physics with Modern Physics (12th ed.). San Francisco: Pearson Education, Inc.

Sutopo
Physics Department
State University of Malang
Malang, Indonesia
E-mail: sutopo@fisika.um.ac.id

Bruce Waldrip
Faculty of Education
Monash University
Churchill, Australia
E-mail: bruce.waldrip@monash.edu

You might also like