You are on page 1of 23

Res Sci Educ (2013) 43:413435

DOI 10.1007/s11165-011-9275-9

Inquiry-Based Laboratory Activities in Electrochemistry:


High School Students Achievements and Attitudes
Burcin Acar Sesen & Leman Tarhan

Published online: 4 December 2011


# Springer Science+Business Media B.V. 2011

Abstract This study aimed to investigate the effects of inquiry-based laboratory activities
on high school students understanding of electrochemistry and attitudes towards chemistry
and laboratory work. The participants were 62 high school students (average age 17 years)
in an urban public high school in Turkey. Students were assigned to experimental (N=30)
and control groups (N=32). The experimental group was taught using inquiry-based
laboratory activities developed by the researchers and the control group was instructed
using traditional laboratory activities. The results of the study indicated that instruction
based on inquiry-based laboratory activities caused a significantly better acquisition of
scientific concepts related to electrochemistry, and produced significantly higher positive
attitudes towards chemistry and laboratory. In the light of the findings, it is suggested that
inquiry-based laboratory activities should be developed and applied to promote students
understanding in chemistry subjects and to improve their positive attitudes.
Keywords Attitude towards chemistry laboratory . Attitude towards chemistry lesson .
Chemistry education . Electrochemistry . Inquiry-based laboratory
Introduction
The laboratory setting has been recognized as a unique instructional environment in which
students can work cooperatively in small groups (Schwab 1962; Hurd 1969; Hofstein and
Lunetta 1982; DeBoer 1991; Lin 2007). Research has shown that students learning will be
more meaningful if they engage in laboratory activities (Domin 2007; Garnett et al. 1995;
Hodson 1990; Hofstein and Lunetta 1982, 2004; Lazarowitz and Tamir 1994; Lunetta 1998;
Tobin 1990). Unfortunately, as it is traditionally structured, science laboratory instruction
B. Acar Sesen
Hasan Ali Yucel Education Faculty, Department of Science Education, Istanbul University, 34452
Eminonu, Istanbul, Turkey
e-mail: bsesen@istanbul.edu.tr
L. Tarhan (*)
Science Faculty, Chemistry Department, Dokuz Eylul University, 35160 Buca, Izmir, Turkey
e-mail: leman.tarhan@deu.edu.tr

414

Res Sci Educ (2013) 43:413435

has the enduring reputation of failing to live up to this expectation (National Research
Council 2006). Consequently, alternative instructional approaches should be utilized to
improve student learning. For this purpose, the inquiry-based laboratory approach has
begun to gain more interest.
Learning Chemistry in Laboratory: Traditional Versus Inquiry-based
In traditional cookbook laboratory settings, students only follow step-by-step instructions to
complete an experiment. Because they concentrate on the completion of individual steps,
they often do not have a deep understanding of the experimental design, and so to many
students, laboratory activities mean manipulating equipment but not manipulating ideas
(Hofstein and Lunetta 2004). Wu and Hsieh (2006) underlined that whereas cookbook labs
can teach some laboratory techniques or serve as visual aids for concepts already studied,
they are largely ineffective as a tool for teaching science concepts. Therefore, cookbook
laboratory activities may work well as illustrations of concepts already studied and
understood but it is unlikely they will lead to new conceptual learning. However, in many
school science programs, laboratories have been used in a cookbook fashion to verify
scientific facts and not promote laboratory or science process skills to investigate the
natural phenomena (Kln 2004). Gunstone (1991) indicated that if students in the science
laboratory are usually involved primarily in technical activities, with few opportunities for
metacognitive activities, they may not construct the knowledge.
Meaningful learning in the laboratory will only occur if students are given ample time
and the opportunities for interaction and reflection to initiate discussion (Gunstone and
Champagne 1990; Tobin 1990). Hofstein and Lunetta (2004) advocated more intensive
research on the effect of science laboratory instruction on the development of students
conceptual understanding, and they indicated that when laboratory experiences are
integrated with other metacognitive learning experiences such as predictexplainobserve
demonstrations, etc., and when they incorporate the manipulation of ideas instead of simply
materials and procedures, they promote the learning of science. They also stated that
to acquire a more valid understanding . . . science educators need to conduct more
intensive, focused research to examine the effects of specific school laboratory
experiences and associated contexts on students learning. The research should
examine the teachers and students perceptions of purpose, teacher and student
behaviour, and the resulting perceptions and understandings (conceptual and
procedural) that the students construct (p. 33).
As a result of this research, interest in using inquiry-based teaching strategies has
increased in recent years as science teachers have become more critical about the efficacy of
cookbook-type laboratory activities and indeed the purposes, practices, and learning
outcomes of laboratory. In science instruction, laboratory activities have been a popular
vehicle for activity/performance-based science tasks for a long time. Thus, many science
educators have advocated the use of inquiry-based laboratory work (Abd-El-Khalick et al.
2004; Hodson 1990; Lunetta 1998; National Research Council 2000).
The National Science Education Standards use the term inquiry in two ways (Bybee
2000; Lunetta 1998): (a) inquiry as content understanding, in which students have
opportunities to construct concepts and patterns, and to create meaning about an idea to
explain what they experience; and (b) inquiry in terms of skills and abilities. Under the
category of abilities or skills, Bybee (2000) included identifying and posing scientifically
oriented questions, forming hypotheses, designing and conducting scientific investigations,

Res Sci Educ (2013) 43:413435

415

formulating and revising scientific explanations, and communicating and defending


scientific arguments. It is suggested that many of these abilities and skills are in alignment
with those that characterize inquiry-based laboratory work, an activity that puts the student
in the centre of the learning process. In the inquiry-based laboratory environment, students
may undertake the following activities: observing objects and events, posing questions,
designing investigations, proposing explanations, collecting and analyzing data, comparing
proposed explanations with new data.
Hofstein et al. (2004) reported that by conducting the experiments in the context of
high school chemistry curriculum, students were able to practice inquiry skills such as
asking questions, hypothesizing, and suggesting a question for further investigation. The
results of the study by Hofstein et al. (2005) also showed that laboratory experiments
increase students ability to ask more questions and better questions related to their
experimental observations and findings. Brown et al. (1989), Williams and Hmelo
(1998), Gunstone (1991) asserted that students construct their knowledge by solving
genuine and meaningful problems during laboratory activities. As mentioned by Taraban
et al. (2007), even when there is a shift in the direction of implementing inquiry learning
in the classroom, teachers can still take away from the true nature of science by giving
students cookbook experimentsactivities through which students carefully follow
step-by-step instructions and collect data without a clear understanding of the question
or concepts, and without opportunities to reason about their observations (National
Research Council 2005).
Thus, inquiry-based laboratory work is an effective mode of learning to improve
students understanding (Lord and Orkwiszewski 2006), scientific process skills (Deters
2005; Hofstein et al. 2004), attitudes toward school science (Gibson and Chase 2002; Jones
et al. 2000; Lord and Orkwiszewski 2006), motivation to learn science (Tuan et al. 2005),
understanding of the nature of science (Backus 2005), and communication skills (Deters
2005). On the other hand, it is well known that students in a typical middle or high school
have few opportunities to engage in inquiry-based activities, and learning environments are
needed in which students can engage in scientific discussion to explain and defend their
thinking (Leonard and Chandler 2003; Tsai 2001). As mentioned by Cheung (2006), there
has been a lack of effective inquiry materials, and he asserted that cookbook-style
laboratory activities are by far the most common among commercial chemistry curricula
and thus teachers have difficulties finding inquiry materials. For this reason in this study,
inquiry-based laboratory activities were developed and applied to high school students to
assess their efficacies.

Purpose and Research Questions


In Turkey, chemistry teaching at high school level includes laboratory activities like a cookbook design, and few chemistry teachers use inquiry-based laboratory work as a teaching
aid as well as the other countries in the world (Deters 2005; Hackling et al. 2001).
Therefore, the purpose of this study was to investigate the effectiveness of inquiry-based
laboratory activities on high school students understanding of electrochemistry and
attitudes towards chemistry lessons and the chemistry laboratory. In order to enhance this
aim the following research questions were investigated;
1. What is the high school students prerequisite knowledge about their proficiency for
learning Electrochemistry?

416

Res Sci Educ (2013) 43:413435

2. Does inquiry-based laboratory instruction contribute to better conceptual understanding


of Electrochemistry in high school students than traditional cook-book design
laboratory instruction?
3. Does inquiry-based laboratory instruction contribute to the high school students
performance in the laboratory?
4. Does inquiry-based laboratory instruction contribute to the high school students
attitudes towards chemistry laboratory activities?
5. Does inquiry-based laboratory instruction contribute to the high school students
attitudes towards chemistry lessons?

Method
Participants
The sample of this study was 62 high school students (average age 17 years) from two
science classes in a high school in Izmir, in Turkey. Chemistry backgrounds of all the
students were the same, and they had been instructed by the same competent chemistry
teacher who has 19 years experience. Students in the classes were randomly assigned to the
experimental (N=30) and control groups (N=32). Students in both groups were instructed
according to the traditional approach by the same teacher during the same instructional
period. Additionally, while inquiry-based laboratory activities were accomplished in the
experimental group, traditional cook-book laboratory instruction was used in the control
group.
Instruments
The Pre-Test
Constructivism claims that existing concepts play an important role for learning new
concepts (Bodner 1986). To learn the subject of electrochemistry, students have to know the
prerequisite subjects of (a) Periodic table, (b) Electronegativity, (c) Ionisation energy, (d)
Electron affinity, (e) Metals and non-metals, (f) Chemical reactions, (g) Chemical
equilibrium, (h) Acids and bases, (i) Oxidation-reduction, (j) Redox reactions, (k) Element
activity. For this reason, a pre-test consisting of thirteen multiple-choice items was
developed to identify student prerequisite knowledge about their proficiency for learning
Electrochemistry. The content of the test was validated by seven chemistry educators and
11 high school chemistry teachers. The test was piloted with the sample of 146 high school
students for the reliability. After the item analysis, the reliability coefficient (KR-20) of the
test was found to be 0.81.
Students answers were classified as correct (1 points), incorrect and no answers
(0 points). The maximum score for the test, in which a student can achieve, is 13.
The Electrochemistry Achievement Test (EAT)
In this study, the Electrochemistry Achievement Test (EAT) developed by Acar and Tarhan
(2007) was used to determine students understanding of Electrochemistry. The test
involved 8 open-ended and 12 multiple-choice items, related to (a) Reactions in

Res Sci Educ (2013) 43:413435

417

electrochemical cells, (b) Construction of electrical current, (c) Identification of anode and
cathode and their charges, (d) Functions of salt bridge, (e) Functions of metal rods, (f)
Function of voltmeter, (g) Cell potential, (h) Half-cell and standard hydrogen electrode, (i)
Electrolysis. Prior to the development of the test items, the content boundaries and
instructional objectives had been defined. Test items had been constructed according to the
objectives and by considering misconceptions identified in the literature as seen in Table 1
(Garnett et al. 1990a, b; Garnett and Treagust 1992a, b; Sanger and Greenbowe 1997,
1999). The content of the tests had been validated by six experts in chemistry education and
four high school chemistry teachers. In addition, the test had been piloted with 150 high
school students for reliability. The reliability coefficient (KR 20) of the test was 0.86.
For the statistical analysis of EAT, multiple choice items were scored as correct (1 point)
and incorrect (0 point) and blank (0 point). In addition to this, open-ended items were
categorized as correct (2), partially correct (1); incorrect (0) and no-response (0). The
correct answers category involved completely correct explanations and the response reflects
the learning objectives in a detailed and clear manner. The incorrect answers category
included incorrect ideas and alternative conceptions on the related topics. On the other
hand, both correct answers with inadequate explanations were placed into the partially
correct answers category. Due to the fact that each correct multiple-choice item and openended item were graded with one point and two points respectively, the maximum score
which can be obtained from the test was 32.
Attitudes Toward Chemistry Lesson (ATCS) and Laboratory Scales (ATCLS)
To determine students attitudes toward the chemistry lesson before and after the
instruction, a 5-point Likert type Attitude toward Chemistry Lesson Scale (ATCS) with
25 items was used (Acar 2008). Before development of the items, literature related to the
attitudes towards science and chemistry had been reviewed (Berberolu and alkolu
1992; Freedman 1997; Hofstein and Lunetta 1982; Koballa 1988; Koballa et al. 1990; Salta
and Tzougraki 2004). The items were constructed by considering the attitude scale
developed by Salta and Tzougraki (2004). For the validity, the scale was reviewed by seven
educators in the different universities. After the corrections, the scale was applied to 168
high school students for the reliability. Cronbachs alpha reliability coefficient was found to
be 0.81. The ATCS has four dimensions:
(1)
(2)
(3)
(4)

Interest in chemistry lesson (items: 1, 3, 8, 18, 20, 22)


Understanding and learning chemistry (items: 2, 5, 7, 12, 13, 14, 15, 17, 21, 23)
The importance of chemistry in real life (items:4, 6, 11, 19, 25)
Chemistry and occupational choice (items: 9, 10, 16, 24)

Students attitudes toward the chemistry laboratory before and after the instruction were
determined by using a 5-point Likert type Attitude toward Chemistry Laboratory Scale
(ATCLS) developed by Tarhan (2008). Before development of the items, literature reviews
were done (Carlo and Bodner 2004; Freedman 1997). For the validity, the scale was reviewed
by seven educators in the different universities. It was applied on 191 high school students,
and Cronbachs alpha reliability coefficient was to be 0.87. The ATCLS has four dimensions:
(1)
(2)
(3)
(4)

Laboratory environment and using equipment (items: 1, 6, 9, 13)


Experimental process in the laboratory (items: 2, 3, 4, 5, 7, 8, 10, 12, 23, 25)
Assessment in the laboratory (items: 11, 15, 16, 17, 18, 19, 20, 21, 26)
Cooperative learning in the laboratory (items: 14, 22, 24, 27).

418

Res Sci Educ (2013) 43:413435

Table 1 Students misconceptions determined in the literature


Student misconceptions
Electrical current
Electrons enter the electrolyte at the cathode, move through the electrolyte and emerge at the anode to
complete the circuit
Protons and electrons flow in the opposite directions to constitute an electric current
Electrons can flow through aqueous solutions without assistance from the ions
Only negatively charged ions constitute a flow of current in the electrolyte and the salt bridge
Identify cathode and anode and their function
The anode is negatively charged and because of this it attracts cations. The cathode is positively changed
and because of this it attracts anions
Anodes, like anions, are always negatively charged; cathodes, like cations, are always positively charged
The anode is positively charged because it has lost electrons
The cathode is negatively charged because it has gained electrons
The identity of the anode and cathode depends on the physical placement of the half-cells
In electrochemical cells oxidation occurs at the anode and reduction occurs at the cathode while in
electrolytic cells oxidation occurs at the cathode and reduction occurs at the anode
Function of metal rods
Metal rods only act as an electron carrier during redox reactions and so there will be no change on the
electrodes physical structure
No reaction will occur if inert electrodes are used
Inert electrodes can be oxidized or reduced
Function of salt bridge
The salt bridge supplies electrons to complete the circuit
The salt bridge assists the flow of current (electrons) because positive ions in the bridge attract electrons
from one half-cell to the other cell
Potential differences and cell potential
In electrochemical cells, as the attractive forces between anions and cations affects the ions velocity towards
electrodes, different potentials are read when different solutions are used in the cells
Protons and electrons flowing in opposite directions cause a potential difference between the two ends of the
wire
Half-cell
A standard half-cell is not necessary
Electrolysis
In electrolytic cells the polarity of the terminals of the applied voltage has no effect on the site of the anode
and cathode
Water does not react during the electrolysis of aqueous solution
The same products are produced in both aqueous and molten substances of salt electrolysis
There is no association between the calculated e.m.f. of an electrolytic cell and the magnitude of the applied voltage
In electrolytic cells with identical electrodes connected to the battery, the same reactions will occur at each electrode
It is not important which sides of the battery are connected to the electrodes, as the same reactions occur at
the electrodes

For statistical analysis, positive items in both the scales were assigned a numeric value
ranging across choices of (5) strongly agree, (4) agree, (3) undecided, (2) disagree, (1)
strongly disagree, and negative items were assigned their reverse. The maximum scores
which a student can obtain from the ATCS and ATCLS were 125 and 135 respectively.

Res Sci Educ (2013) 43:413435

419

Laboratory Assessment Form (LAF)


To evaluate students performance for each experiment, Laboratory Assessment Form
(LAF) was developed by considering the literature review (Reynolds et al. 1995; Tamir et
al. 1982; Lunetta et al. 1981; Doran et al. 2002). LAF includes three sub-scales as Readiness for laboratory experiment, -Behaviour in the laboratory, and -Evaluating the
findings and report writing (Table 2). To determine the validity of the form, it was
examined by ten chemistry educators and 11 competent chemistry teachers. After the
correction, ten chemistry teachers were required to observe and assess three students in their
class while they perform the experiment by using the LAF. The intraclass correlation, which
is found as 0.89, indicated that the internal consistency and reliability of the LAF is high.
Laboratory Assessment Form was used for all the students in the experimental group after
each experiment and then evaluated by the teacher and two observer teachers. For the
assessment of the students performance, the items were scored as (1) fail, (2) satisfactory,
(3) good and (4) excellent. The maximum score which can be obtained from LAF was 56.
Procedure
The Turkish high school chemistry curriculum contains the subject of electrochemistry under
the unit of Oxidation and Reduction Reactions. Before the subject of Electrochemistry,
students are taught the subjects of Oxidation and Reduction; Oxidant and Reductant;
Table 2 The sub-scales in the Laboratory Assessment Form (LAF)
Assessment criteria
Readiness for laboratory
Understanding of theoretical knowledge and concepts related
to experimental subjects
Assimilate the aim of the experiment
Investigate the pre-experiment questions in the worksheet
Behaviour in the laboratory
Knowledge about how to carry out the experiment
Use laboratory equipment and materials in a good manner
Record experimental observations and results in a meaningful
and accurate manner
Share, discuss, aid and stick together with friends during the
laboratory process
Make an effort to solve the problems encountered during
the laboratory process
Adhere to the laboratory rules
Evaluating the findings and writing laboratory report
Present results and conclusion clearly
Analyze experimental data accurately using the most
appropriate representation
Associate the findings with their theoretical knowledge
Associate the answers of the post-experiment questions and
experimental results with theoretical knowledge
Analyze experimental error sources

Very good

Good

Average

Poor

420

Res Sci Educ (2013) 43:413435

Oxidizing and Reducing Agent; Redox Reactions; Half Reactions; Electron Taking and
Giving; Elements Activities; Oxidation States, Determining Oxidation States, and Balancing
Redox Equations. In the context of Electrochemistry, they learn Electrode potentials,
Electrochemical cells, Cell potentials, Standard hydrogen electrode, and Electrolysis.
In this study, the aim was to investigate whether inquiry-based laboratory activities are
more effective in understanding the subject of Electrochemistry than a traditional
laboratory instruction. For this reason, pre- and post-tests with a control group design was
used. Before the instruction, the pre-test was applied to both control and experimental
groups to identify student prerequisite knowledge about their proficiency for learning
Electrochemistry. In addition, ATCS and ATCLS were applied to identify their preattitudes. The independent t-test was conducted to compare the scores of groups with no
significant differences with respect to students mean scores of pre-test, ATCS and ATCLS.
Students in both group were taught the subject of Electrochemistry by the same teacher
during the 3 week period including 4 h per week. Throughout the lesson, the teacher
presented the subject by using the blackboard, asked some questions related to the subject
and students solved the problems. The students were instructed with the regular chemistry
textbook. They listened to the teacher carefully, took notes and solved algorithmic
problems. In addition to this treatment, the experimental group were engaged in inquirybased laboratory activities and the control group students studied regular chemistry
experiments. Before the instruction, a chemistry teacher of 19 years experience was trained
in how to implement the instruction based on an inquiry-based laboratory. Inquiry-based
learning requires students to work together rather than receiving direct instructions on what
to do from the teacher. The teachers job in this setting is therefore not to provide
knowledge, but instead to help students along the process of discovering knowledge for
themselves. For this reason, the teacher was required to act as a facilitator, visit and monitor
the groups, explain using laboratory equipment, warn about the possible laboratory hazards
and ask leading questions to encourage students to think, discuss and research. Before the
instruction, the teacher gave information about the laboratory process, using laboratory
equipment and materials, rules of working in the laboratory, how to conduct data evaluation
and report writing techniques, and laboratory safety. The teacher explained what should be
done if laboratory accidents occurred explaining emergency equipment and procedures.
Students were also required to wear a laboratory coat, glasses and gloves, and follow the
experimental stages indicated in the hand-outs. Students were informed that they would be
assessed according to their performances and the group report for each laboratory work. In
the group report, students were required to write the aim of the experiment, experimental
procedure, results, evaluation and discussion and conclusion.
Conducting the Inquiry-based Laboratory Activities
Five inquiry-based laboratory activities related to Electrochemistry were developed based
on constructivism by considering students misconceptions and learning difficulties
determined in the literature (see Table 1). Inquiry-based laboratory activities were carefully
constructed to integrate the learning sequences with conceptual networks. The headings of
(a) Aims of the experiments, (b) Equipment and chemicals, (c) The warnings, (d)
Experimental procedures were clearly indicated in the laboratory hand-outs for all the
laboratory activities. The leading questions were especially constructed to require
construction of knowledge by encouraging students to research, discuss, and share their
knowledge in their small groups. The laboratory activities were examined by seven
chemistry educators and 11 high school chemistry teachers. The activities were piloted with

Res Sci Educ (2013) 43:413435

421

the sample of 23 high school students attending a different high school and revision was
made according to their feedback.
The main application of inquiry-based laboratory activities was accomplished with
the participation of 30 high school students who were randomly assigned into their
cooperative groups based on their scores obtained from the pre-test, ATCS and ATCLS.
There were six groups with five students. All the laboratory activities were performed
in these small cooperative groups under the guidance of the teacher. Each laboratory
activity began with a problem related to the laboratory activities. Then the students
were asked to define or describe the problem. During this process, the teacher asked
some leading questions to arouse their interest. After the brainstorming, students began
to conduct activities. Students asked relevant questions of each other and discussed
their observations in their groups, and finally, analyzed the findings. During the
inquiry-based laboratory process, it was required students construct the knowledge step
by step by using existing ideas.
Laboratory Activity-1 The first laboratory activity began with a problem related to
Galvanis observations on a trembling frogs leg connected to
Zn and Cu metals, and group discussion began around this
problem. Students were then required to conduct the laboratory
activity named Potential differences of different metal rods in
different fruits to solve the problem. They designed their
experiments and observed the potential differences between the
copper rod and some metal rods such as Zn, Sn, Mg, or Ni
immersed into a fruit like lemon, apple, and orange using a
simple voltmeter, and then noted the values into a Table. Students
were required to inquire about the reasons for changes in the
potential differences according to the type of the metal pairs and
fruits. During this period, they activated their prior knowledge
such as Redox Reaction, Oxidation and Reduction, and Element
Activities.
Laboratory Activity-2 After the first laboratory work, students learned that potential
difference is dependent on the concentration of a solution
where metal rods are immersed. Students were asked how to
construct a standard cell system for international validity.
After the brain storming, a simulation was presented related
to the cell system where Zn and Cu rods were immersed in
1 M HCl solution, and it required students to discuss the
reason for the lighting of the lamp by considering the
chemical reactions occurring in the cell, and the oxidation
tendency of the metals. After constructing a system including
one cell, laboratory activity-2 was conducted by the students,
titled If Zn and Cu rods were immersed into two different cells,
does the lamp light up? . Before this activity, students activated
their previous knowledge as Anion, Cation, Electrolyte, Oxidation, Reduction, Redox reactions and Element activities. In the
first step of the experiment, students made a system by
immersing the Cu rod into a beaker include 1 M CuSO4 and
the Zn rod into the other beaker including 1 M ZnSO4 solutions,
and the rods and the lamp were connected via a conductive wire

422

Res Sci Educ (2013) 43:413435

to complete the circuit. Students were required to inquire why


the lamp did not light up and interpret their observations with
their previous knowledge. In the second step, the teacher gave
out to all the groups a U shaped glass tube filled with a saturated
electrolyte (potassium chloride) and closed with a cotton wool.
Students were encouraged to connect the beaker to each other
using the salt bridge and then to comment on their observations.
Students defined the reactions that occurred in the beakers,
electron and ion flow, salt bridges function and energy
transformation in the system. As a result of this experiment,
the aim was for students to learn the terms Electrode, Anode,
Cathode, Half-cell and Salt-bridge and understand the working
principle of the electrochemical cells.
Laboratory Activity-3 After completion of the second laboratory activity, a sample
from daily life that showed water flowing spontaneously over
a waterfall from high potential energy to low potential energy
was given to students as a simulation for electron flow from
anode to cathode demonstrating electromotor forces (cell
potential). They began to express their opinions about the
affective parameters of the cell potential. After the brainstorming, laboratory activity-3 named Effective Factors on Cell
Potential was conducted by students to understand the change of
cell potential depending on concentration and temperature. In this
activity, students made four different electrochemical cells by
using Cu and Zn electrodes and CuSO4 and ZnSO4 solutions in
the different concentration as 0.05 M and 2 M with the same
volume and salt bridge filled with saturated Na2SO4. The cell
potentials for four electrochemical cells were measured via a
voltmeter. Students inquired about the effect of concentration on
cell potential. Then, students designed another electrochemical
cell composed of Cu and Zn electrodes; 2 M CuSO4 and 2 M
ZnSO4 solutions in two conditions of ice water and boiling water,
and they measured the potential differences. Students were
encouraged to inquire about the effect of temperature on cell
potential.
Laboratory Activity-4 To arouse students interest, they were asked whether electrical
energy could be transferred to chemical energy. After brainstorming, they performed a laboratory activity-4 named Water
Electrolysis. Students secured two test tubes filled with water in
the beaker in a way that test tubes were upside down over the
beaker, mounted the cupper and carbon electrodes and then
connected the 12 V battery. The events occurring in the system
were observed and noted by students. Then, 23 mL of 1 M
Na2SO4 was added to the water and observations were recorded.
After completion of the experiment, pH of water was measured.
While students interpreted the results of the experiment, they were
encouraged to inquire which gases were released in the anode and
cathode by writing the half reactions, the change of pH of the
water and the reason for adding Na2SO4.

Res Sci Educ (2013) 43:413435

423

Laboratory Activity-5 The last activity was related to electroplating. Firstly, students were
asked, how to make jewellery or watch plating with silver or gold.
After the brainstorming, students formulated their hypothesis, and
began to design their experiments. During this activity, students
made an electrolysis system and plated a spoon with copper. They
were required to inquire about the examples of electrochemistry
from daily life.
Conducting the Traditional Laboratory Activities
The same laboratory activities were conducted in the control group based on traditional
cookbook settings. Students were given laboratory hand-outs with the description of the
laboratory for performing the experiment. Students read the hand-out and then followed
step-by-step directions to conduct the experiments. They were not required to ask questions
and discuss the reason for the findings.

Results
Results of the Pre-Test
In order to identify students prior knowledge of Electrochemistry, the pre-test was
administered to both control and experimental groups. An independent sample t-test was
conducted to compare the mean scores of experimental and control groups. As seen in
Table 3, the analysis results expressed that there was no statistically significant difference
among the control and experimental groups in terms of pre-test mean scores (t=0.16, p>.05).
Results of the Electrochemistry Achievement Test
In order to identify students understanding of Electrochemistry, the Electrochemistry
Achievement Test (EAT) was applied after the implementation. The mean scores of both
control and experimental groups were compared by conducting an independent sample ttest, and the results showed there was a statistically significant difference between groups (t
=12.07, p<.05, Table 4).
Based on the EAT results, it was found that in the experimental group students had fewer
misconceptions and understood the concepts more meaningfully than students in the control
group. It was found that the control group students commonly failed to explain -Flow of
electrons and ions, -Function of salt bridge, -Identification of anode and cathode, and Confused electrolytic cell with electrochemical cells.
Table 3 Independent sample t-test results of the pre-test, ATCS, ATCLS before the instruction
Groups

Experimental group (N=30)

Control group (N=32)

SD

SD

Pre-test

8.23

1.87

8.16

1.88

0.16

0.87

ATCS

67.10

12.73

67.34

13.79

0.07

0.94

ATCLS

77.73

7.75

79.81

13.79

0.72

0.47

424

Res Sci Educ (2013) 43:413435

Table 4 Independent sample t-test results of EAT, ATCS and ATCLS after the instruction
Groups

Experimental group (N=30)

Control group (N=32)

SD

SD

EAT

23.93

2.08

14.31

3.96

12.07

.00

ATCS

87.73

10.49

67.97

12.94

6.58

.00

ATCLS

95.20

18.88

79.84

13.76

3.64

.00

Results of Attitudes towards Chemistry Lesson Scale


In order to measure students pre- and post-attitudes towards chemistry, The Attitudes
towards Chemistry Lesson Scale (ATCS) was used. The independent sample t-test was
used to compare the mean scores of the experimental and control groups. Statistical
results showed that while there were no significant differences among experimental and
control groups with respect to pre-attitude towards chemistry lessons (t=0.07, p>.05;
Table 3), significant differences were found between groups after the instruction (t=6.58,
p<.05; Table 4).
The mean scores of both control and experimental groups obtained from ATCS before
and after the instruction were compared by conducting paired sample t-test analysis. As
seen in Table 5, in contrast to the control group (t=1.00, p>0.05), the mean score of the
experimental group increased significantly after the instructions (t=6.37, p<0.05).
Twenty-five items in the ATCS were investigated in four sub-dimensions as; (a) Interest
in chemistry lessons; (b) Understanding and learning of chemistry; (c) The importance of
chemistry in real-life, and (d) Chemistry and occupational choice. As seen in Table 6, while
the experimental group students attitudes for all dimensions increased, there were no
significant changes for control group students. If students attitudes towards the first
dimension (interest in chemistry lesson) is analysed, it is seen that experimental group
students require chemistry lessons more often. While 50.00% of the experimental group
students would like to have fewer chemistry subjects in their chemistry lessons before the
instruction, the percentage of those students decreased to 16.67 after the instruction. In
harmony with these results, students positive attitudes towards finding chemistry lessons
more interesting and necessary increased significantly (Table 6).
Students answers to the second dimension (understanding and learning chemistry)
indicated that students in the experimental group began to appreciate the importance of
learning the basic concepts for understanding chemistry. They also began to find the use of
chemical symbols and concepts to be easier after the instruction. The percentage of students
in the experimental group who had positive thoughts about how some knowledge in

Table 5 Paired sample t-test results of ATCS before and after the instruction
Group

Pre-attitude

Post-attitude

Experimental

30

67.10

12.73

30

87.73

10.49

6.37

.00

Control

32

67.34

13.79

32

67.96

12.94

1.00

.33

26.67
36.67

8. I think chemistry lessons are unnecessary

The importance of chemistry in real-life

Understanding and learning chemistry

33.33

1. I like chemistry lessons

3. I would like the teaching period of the chemistry lesson more often

13.33
63.33

17. I believe that some knowledge in chemistry helps us understand the other
science lessons more easily

21. I can understand chemistry concepts easily

23. I have difficulties while using my knowledge in solving chemistry problems

25. I think chemistry has a great role in solving environmental problems

21.88

28.13

59.38

15.63

28.13

15.63

37.50
34.38

26.67

15. I find using chemical symbols to be easy

65.63

56.25

33.33
30.00

13.33

14. I make many efforts to understand chemistry

37.50

63.33

13. Chemistry is a sophisticated and impalpable lesson

56.25

23.33

56.67

12. Most of the concepts in chemistry are not concrete

34.38

18.75

40.00

60.00

7. I think, learning the basic concepts are important for understanding chemistry

11. I think the level of chemistry technology in a country is an important


indicator for development of the country
19. I think chemistry has a great role in modern life

33.33

5. I can solve chemistry problems easily

34.38
34.38

6. I think developments in chemistry improve the quality of our lives

13.33

22. I find chemistry lessons interesting


2. Chemical symbols are unintelligible as a foreign language that I do not know

21.88
50.00

30.00

36.67
36.67

20. I would like to have fewer chemistry topics in the lessons

4. I believe that chemical knowledge helps us to interpret seriously


events in our daily life

20.00
50.00

18. I hate chemistry lessons

37.50

25.00

28.13

Cont.

Exp.

Interest in chemistry lesson

Post-attitude

Pre-attitude

60.00
76.67

60.00

56.67

53.33

30.00

53.33

46.67

60.00

40.00

33.33

30.00

76.67

60.00

66.67
16.67

16.67

10.00

16.67

60.00

63.33

Exp.

Percentages of students answers

Items in the attitudes towards chemistry lesson scale

Dimension

Table 6 The percentages of students answer to the ATCS before and after the instruction

40.63
37.50

34.38

25.00

31.25

59.38

15.63

31.25

15.63

62.50

53.13

59.38

31.25

25.00

31.25
37.50

46.88

21.88

34.38

25.00

31.25

Cont.

Res Sci Educ (2013) 43:413435


425

56.67
53.33
46.67
16.67

9. I do not believe that chemistry knowledge will be useless after my graduation

10. I believe that I do not need chemistry knowledge for my target career

16. I do not find the jobs related to chemistry as attractive

24. My target career is chemist/chemistry teacher/chemical engineer

15.63

46.88

43.75

56.25

Cont.

Exp.

Chemistry and occupational choice

Post-attitude

Pre-attitude

20.00

20.00

26.67

23.33

Exp.

Percentages of students answers

Items in the attitudes towards chemistry lesson scale

Dimension

Table 6 (continued)

18.75

43.75

43.75

59.38

Cont.

426
Res Sci Educ (2013) 43:413435

Res Sci Educ (2013) 43:413435

427

chemistry helped them to understand the other science lessons more easily, increased
significantly from 26.67 to 46.67. It was also found that students began to find chemical
symbols intelligible and concepts as concrete (Table 6).
According to the students answers to the third dimension related to attitudes towards the
importance of chemistry in real-life, it was found that the experimental group of students
began to believe that chemical knowledge helped to interpret events in daily life in the
percentage of 53.33%, and chemistry has a great role in the modern life in the percentage of
60.00%. The percentage of students who think that the level of chemistry technology is an
important indicator for development of the country and chemistry has a great role in solving
environmental problems, also increased significantly.
The lowest significant increase was found for the fourth dimension which related to
chemistry and occupational choice. The highest increase in students attitudes was found in
the items such as believing chemistry knowledge will be useless after graduation and
needing chemistry knowledge for their target career. While the percentage of experimental
group students who thought the jobs related to chemistry as unattractive decreased from
46.67 to 20.00, their thoughts about choices of jobs related to chemistry were not changed
significantly.
Results of Attitudes Towards Chemistry Laboratory Scale
In order to measure students pre- and post-attitudes towards chemistry laboratory, the
Attitudes towards Chemistry Laboratory Scale (ATCLS) was used. The independent sample
t-test was used to compare the mean scores of the experimental and control groups.
Statistical results showed that while there were no significant differences among
experimental and control groups with respect to pre-attitudes towards chemistry lessons
(t=0.72, p>.05; Table 3), significant differences were found between groups after the
instruction (t=3.64, p<.05; Table 4).
The mean scores of both control and experimental groups obtained from ATCLS before
and after the instruction were compared by conducting paired sample t-test analysis. As
seen in Table 7, while the mean score of experimental groups significantly increased from
77.73 to 95.20 (t=4.73, p<0.05), the increases in the mean score of the control group from
79.81 to 79.84 was not significant (t=0.05, p>0.05).
Twenty seven items in the ATCLS were investigated in four sub-dimensions as; (a)
Laboratory environment and using equipments, (b) Experimental process in the laboratory,
(c) Assessment in the laboratory, and (d) Cooperative learning in the laboratory. The results
showed that experimental group students mean scores for all dimensions increased
significantly after the instruction compared to the control group.
Table 8 shows the percentages of experimental and control group students pre- and
post-attitudes towards chemistry laboratory in terms of the dimensions. The answers for the

Table 7 Paired sample t-test results of ATCLS before and after the instruction
Group

Pre-attitude
N

Experimental

30

77.73

Control

32

79.81

Post-attitude
S

7.75

30

95.20

18.88

4.73

.00

13.79

32

79.84

13.14

0.05

.96

46.67 43.75 96.67 43.75


60.00 53.13 30.00 56.25
63.33 40.63 36.67 40.63

9. Fear of damaging apparatus in the laboratory decreases my desire to perform experiments

13. Because I am afraid to crash the glass apparatus, I do not like to perform experiments

Assessment in the laboratory

Experimental process in the


laboratory

36.67 40.63 66.67 43.75

6. Laboratory environment should be safe for the experiments which will be performed

36.67 31.25 26.67 31.25


46.67 50.00 20.00 53.13
46.67 46.88 16.67 50.00
43.33 37.50 56.67 40.63
30.00 28.13 50.00 28.13

7. I do not believe that faults occurring during the experimental procedure reduce my motivation

8. If purpose of the experiment is not clearly indicated, my interest in the experiment is reduced

10. I prefer theoretical lessons than performing experiments in the laboratory


12. I think carrying out an experiment is a stressful and unnecessary process

23. I believe that advancing an opinion during the experimental process increases my motivation

25. Laboratory experiments contribute to developing my hands-on skills

43.33 50.00 26.67 50.00


23.33 21.88 56.67 25.00
43.33 50.00 20.00 50.00
36.67 25.00 63.33 28.13

18. Comprehending the experimental results increases my analytical thinking capacity


19. I worry about misinterpreting the experimental findings

20. Investigation of the reasons for experimental errors arouses my interest

16. I am apprehensive of my teachers knowledge about my experimental findings

17. Researching the reason for experimental errors arouses my interest

36.67 34.38 70.00 31.25


50.00 53.13 33.33 56.25

15. Proving a scientific law in the laboratory experiment increases my confidence

43.33 40.63 66.67 43.75

20.00 12.50 43.33 15.63

5. I believe that if I am informed about the issues that I must be careful, my self confidence is increased

11. Obtaining correct experimental findings increases my interest in experiments

13.33 15.63 10.00 18.75


33.33 37.50 50.00 40.63

4. I do not believe in the importance of laboratory rules

20.00 25.00 50.00 28.13


20.00 18.75 46.67 21.88

3. I believe that conducting experiments in the laboratory increases my achievement by strengthening the
theoretical knowledge I learnt in the lesson

Cont.

2. While I perform experiments, I feel like a scientist

Cont. Exp.

Post-attitude

1. I would like to be informed about the use of laboratory equipment

Exp.

Pre-attitude

Laboratory environment and


using equipment

Percentages of Students
Answers

Items in the attitudes towards chemistry laboratory scale

Dimension

Table 8 The Percentages of students answer to the ATCLS before and after the instruction

428
Res Sci Educ (2013) 43:413435

Cooperative learning in the


laboratory

Dimension

Table 8 (continued)

60.00 40.63 30.00 37.50


43.33 43.75 70.00 43.75
26.67 31.25 63.33 31.25
33.33 40.63 70.00 40.63

14. I think that teaching something related to the experiment to my group mates is a waste of time
22. I find group work in the laboratory enjoyable

24. Group solidarity during the experimental process in the laboratory strengthens our friendly relations

27. The more I contribute to the experimental process in my group the more I feel better myself

36.67 37.50 70.00 37.50

Cont.

26. The consistency between theoretical knowledge and experimental results increases my motivation

Cont. Exp.

Post-attitude

33.33 31.25 53.33 34.38

Exp.

Pre-attitude

Percentages of Students
Answers

21. I believe that teachers emphasizing my errors in laboratory reports contributes to my improvement
in learning

Items in the attitudes towards chemistry laboratory scale

Res Sci Educ (2013) 43:413435


429

430

Res Sci Educ (2013) 43:413435

first dimension (laboratory environment and using equipments) indicated that while the
percentage of experimental group students who want to be informed about using laboratory
equipments increased from 36.67% to 66.67%, the increases in the control group was from
40.63 to 43.75. Experimental group students also began to believe in the importance of
laboratory safety in a higher percentage.
The percentage of students answers to the second dimension (experimental process in
the laboratory) of the ATCLS showed that the experimental group students attitudes related
to feeling like a scientist while experimenting increased in the ratio of 30%. 46.67% of the
experimental group students also believed that conducting experiments in the laboratory
increased their achievement by strengthening the theoretical knowledge they learned in the
lesson.
The percentage of students answers to the third dimension (assessment in the
laboratory) showed that 70.00% of the experimental group of students began to believe
that proving a scientific law in the laboratory experiment increased their confidence. While
their attitudes about comprehending the experimental results increased, their analytical
thinking capacity increased from 23.33% to 56.67%, and their attitudes about worrying
about misinterpreting the experimental findings decreased from 43.33% to 20.00%.
The highest increase in the experimental group students attitudes was determined for the
fourth dimension (cooperative learning in the laboratory). After the instruction, over 60% of
the students began to find group work in the laboratory enjoyable and that group solidarity
during the experimental process strengthens their friendship. Students negative attitudes
about teaching something related to the experiment to their group mates is a waste of time,
decreased by 30%.
Results of Laboratory Assessment Form
Experimental group students performance in the laboratory for each experiment was
assessed by using the Laboratory Assessment Form. Students mean scores were compared
by conducting ANOVA and Bonferroni test. The mean scores were found as 22.47, 26.19,
37.57, 43.87 and 47.53 for each laboratory activity respectively. As seen in Table 9, the
ANOVA results showed there was a statistically significant difference between mean scores
of students performances (F=103.79, p<0.05).
As seen in the Figure 1, experimental group students mean scores for sub-scales of LAF
increased significantly. While their mean score related to readiness for laboratory
experiments was 5.73 after the first experiment, the mean score significantly increased to
11.96 after the last laboratory activity. Their mean scores in the second sub-scale titled
behaviour in the laboratory and in the third sub-scale titled laboratory report also increased
from 7.75 and 8.99 to 19.44 and 16.12 respectively.

Table 9 ANOVA results of LAF


Number of laboratory activities

Mean

SD

30

22.47

5.76

30

26.19

5.04

30

37.57

8.45

30

43.87

6.70

30

47.53

5.78

Significant difference

103.79

0.00

23, 24, 25

13, 14, 15
34, 35

Res Sci Educ (2013) 43:413435

25,00

431

Readiness for lab work


Behaviour in the lab
Lab report

20,00

19,44

Mean Scores

17,27
16,12

15,25

15,00

13,44
10,56

10,00

8,99
7,75

12,90
11,22

11,37

11,96

8,72
6,91

5,73

5,00

0,00

2
3
4
Number of Experiments

Fig. 1 Experimental group students mean Scores for subscales of LAF

Discussion and Implications


In this study, the effect of inquiry-based laboratory activities on high school students
understanding of Electrochemistry, and attitudes toward chemistry and laboratory work was
investigated.
As mentioned in the literature, laboratory activities have long had a distinctive and
central role in science and especially chemistry and educators have suggested that many
benefits accrue from engaging students in science laboratory activities (Garnett and
Hackling 1995; Hofstein and Lunetta 1982; Hofstein et al. 2004; Lunetta 1998; Tobin
1990). Inquiry-based laboratories have the potential to develop students abilities and skills
such as: posing scientifically oriented questions, collecting and analysing data, forming
hypotheses, designing and conducting scientific investigations, formulating and revising
scientific explanations, and communicating and defending scientific arguments (Hofstein et
al. 2005). The results of this study cohere with these findings. According to the results, it
can be concluded that inquiry-based laboratory activities caused significantly better
acquisition of the scientific conceptions, and increased students laboratory performances,
attitudes towards chemistry lessons and laboratory work.
The results of the EAT showed that the mean score of experimental group students who
carried out inquiry-based laboratory activities were significantly higher than those in the
control group (t=12.07, p<0.05). On the other hand, students responses to the EAT showed
that the number and percentage of misconceptions of the experimental group was
significantly fewer than the control group of students. This result showed the positive effects
of inquiry-based laboratory instruction on students understanding of Electrochemistry and
on preventing misconceptions. It was found that control group students commonly failed to
explain -Flow of electrons and ions, -Function of salt bridge, -Identification of anode and
cathode, and -Confused electrolytic cell with electrochemical cells. These results reflect that
inquiry-based laboratory activities help students to reach higher learning achievements, and
promote the development of cognitive abilities such as problem solving, analysing and
decision making. These results were also supported by the findings of LAF. The results
showed that experimental group students mean scores obtained from LAF increased

432

Res Sci Educ (2013) 43:413435

significantly after each laboratory activity (F=103.79, p<0.05). This shows that inquirybased laboratory activities not only increase students learning achievement, but also support
the development of cognitive abilities, hands-on and social skills. While the mean score on
the readiness for laboratory activity was 5.73 after the first experiment, this value increased
significantly to 11.96. This result showed that students began to understand the nature of
inquiry-based laboratory activities and to study the theoretical knowledge about the laboratory
activity before laboratory instruction. It was also observed that students mean score on the
behaviour in the laboratory increased significantly from 7.75 to 19.44. This result reflected
that students began to behave as a scientist in the laboratory. Their laboratory skills such as
using laboratory equipment and materials in a good manner, recording experimental
observations and results in a meaningful and accurate manner, sharing, discussing, aiding
and sticking together with friends during the laboratory process, making an effort to solve the
problems also increased significantly. Students score on preparing reports as required also
increased from 8.99 to 16.12. This result provides evidence of the development of students
abilities to think critically, analyze experimental findings, and comment on the results.
Research on inquiry-based laboratory work showed similar results (Bybee 2000; Domin
1999; Hofstein et al. 2004).
According to the results of this study, it was found that inquiry-based laboratory
instruction positively effects students attitudes as mentioned in previous research
(Freedman 1997; Ledbetter 1993; Gibson and Chase 2002; Jones et al. 2000; Lord and
Orkwiszewski 2006). In this study, the experimental group of students, who conducted
inquiry-based laboratory activities, scored significantly higher on ATCS (t=6.58, p<0.05)
and ATCLS (t=3.64, p<0.05) than the control group of students. The increases in the
percentage of experimental group students, who like chemistry and want chemistry lessons
more often, showed that inquiry-based laboratory activities positively affected students
interest in chemistry lessons (first dimension of ATCS). The studies by Hewson and
Hewson (1983), Stavy (1991), Sanger (2000) indicated that the difficulties of chemical
concepts cause negative attitudes about these concepts. The results of ATCS indicated that
if students learned these concepts in a meaningful way, their negative attitudes would
decrease. Students attitudes towards the second dimension titled understanding and
learning chemistry, also increased significantly. Experimental group students began to think
learning the basic concepts are important for understanding chemistry and to find the use of
chemical symbols and concepts easily. The significant increase of experimental students
attitudes related to understanding chemical concepts supports the results of EAT and reflects
that their higher order cognitive skills developed during the instructional process.
According to the results, the percentages of experimental group students who indicated
that they began to understand chemistry concepts easily increased to 53.33%, and those
who used chemical symbols and solved chemistry problems easily increased to 60.00%.
This significant increase also shows that inquiry-based laboratory applications helped
students to understand chemistry. Students answers to items related to the attitudes towards
the importance of chemistry in real-life indicated that while the control group of students
beliefs did not change significantly, experimental group students began to realise the
importance of chemistry. These results especially indicate that inquiry-based laboratory
activities derived from daily life attract students interests and help them to integrate
chemistry in their life.
The increase of students mean scores in the ATCLS is important evidence for the effects
of inquiry-based laboratory instruction on attitude towards the chemistry laboratory.
According to the results in Table 8, it is seen that while over 60% of the experimental group
of students did not want to perform experiments because of the fear of breaking or

Res Sci Educ (2013) 43:413435

433

damaging experimental apparatus before the instruction, this percentage decreased to 30%.
The results supported that if students are informed about using laboratory equipment,
possible hazards, laboratory safety and rules of working in the laboratory before the
laboratory activities, their self-reliance will increase and their fears will decrease. The
increase in the percentage of students who think their achievements increase by
strengthening the knowledge learnt in the class during the experimental process, showed
inquiry-based laboratory activities and inquiry-based laboratory instruction promoted
meaningful learning as indicated in the EAT and LAF results.
In the light of these results, while the students who conducted inquiry-based laboratory
activities had few misconceptions about electrochemistry, the students in the traditional
laboratory class had more misconceptions. This situation shows the power of inquiry-based
laboratory instruction for improving students understanding and preventing misconceptions. It was also found that inquiry-based laboratory instruction improves students
laboratory skills, and attitudes towards chemistry and laboratory work. Therefore, inquirybased laboratory activities should be constructed and used widely in chemistry lessons.
Acknowledgement This study was supported by The Scientific and Technological Research Council of
Turkey (TUB-105K058).

References
Abd-El-Khalick, F., BouJaoude, S., Duschl, R., Lederman, N. G., Mamlok-Naaman, R., Hofstein, A., et al.
(2004). Inquiry in science education: International perspectives. Science Education, 88, 397419.
Acar, B. (2008). An active learning application based on constructivism for the subject of acid and bases in
high school chemistry lesson. Dissertation, University of Dokuz Eylul.
Acar, B., & Tarhan, L. (2007). Effect of cooperative learning strategies on students understanding of
concepts in electrochemistry. International Journal of Science and Mathematics Education, 5, 349373.
Backus, L. (2005). A year without procedures. The Science Teacher, 72, 5458.
Berberolu, G., & alkolu, G. (1992). The construction of a Turkish computer attitude scale. Studies in
Educational Evaluation, 24, 841845.
Bodner, G. (1986). Constructivism: a theory of knowledge. Journal of Chemical Education, 63, 873878.
Brown, J. S., Collins, A., & Duguid, P. (1989). Situated cognition and the culture of learning. Educational
Researcher, 18, 3241.
Bybee, R. (2000). Teaching science as inquiry. In J. Minstrel & E. H. Van Zee (Eds.), Inquiring into inquiry
learning and teaching in science. Washington: AAAS.
Carlo, D. I., & Bodner, G. M. (2004). Students perceptions of academic dishonesty in the chemistry
classroom laboratory. Journal of Research in Science Teaching, 41, 4764.
Cheung, D. (2006). Learning about lithium through inquiry-based practical work. Hong Kong Science
Teachers Journal, 23, 18.
DeBoer, G. E. (1991). A history of ideas in science education. New York: Teachers College Press.
Deters, K. M. (2005). Student opinions regarding inquiry-based labs. Journal of Chemical Education, 82,
11781180.
Domin, D. S. (1999). A review of laboratory instruction styles. Journal of Chemical Education, 76, 543547.
Domin, D. S. (2007). Students perceptions of when conceptual development occurs during laboratory
instruction. Chemistry Education Research and Practice, 8, 140152.
Doran, R., Chan, F., Tamir, P., & Lenhardt, C. (2002). Science Educators Guide to Laboratory Assessment,
NSTA.
Freedman, M. P. (1997). Relationships among laboratory instruction, attitudes toward science, and
achievement in science knowledge. Journal of Research in Science Teaching, 34, 343357.
Garnett, P. J., & Hackling, M. W. (1995). Refocusing the chemistry lab: a case for laboratory based
investigations. Australian Science Teachers Journal, 41, 2632.
Garnett, P. J., & Treagust, D. F. (1992a). Conceptual difficulties experienced by senior high school students
of electrochemistry: electric circuits and oxidation-reduction equations. Journal of Research in Science
Teaching, 29, 121142.

434

Res Sci Educ (2013) 43:413435

Garnett, P. J., & Treagust, D. F. (1992b). Conceptual difficulties experienced by senior high school students
of electrochemistry: electrochemical (galvanic) and electrolytic cells. Journal of Research in Science
Teaching, 29, 10791099.
Garnett, P. J., Garnett, P. J., & Treagust, D. F. (1990a). Implications of research of students understanding of
electrochemistry for improving science curricula and classroom practice. International Journal of
Science Education, 12, 147156.
Garnett, P. J., Garnett, P. J., & Treagust, D. F. (1990b). Common misconceptions in electrochemistry: Can we
improve students understanding of this topic? Chemeda: Australian Journal of Chemical Education, 27,
311.
Garnett, P., Garnett, P., & Hackling, M. (1995). Students alternative conceptions in chemistry: a review of
research and implications for teaching and learning. Studies in Science Education, 25, 6995.
Gibson, H., & Chase, C. (2002). Longitudinal impact of an inquiry-based science program on middle school
students attitudes toward science. Science Education, 86, 693705.
Gunstone, R. F. (1991). Reconstructing theory from practical experience. In B. E. Woolnough (Ed.),
Practical science (pp. 6777). Milton Keynes: Open University Press.
Gunstone, R. F., & Champagne, A. B. (1990). Promoting conceptual change in the laboratory. In E. HegartyHazel (Ed.), The student laboratory and the science curriculum (pp. 159182). London: Routledge.
Hackling, M. W., Goodrum, D., & Rennie, L. (2001). The state of science in Australian secondary schools.
Australian Science Teachers Journal, 47(4), 617.
Hewson, M. G., & Hewson, P. W. (1983). Effect on instruction using students prior knowledge and
conceptual change strategies on science learning. Journal of Research in Science Teaching, 20, 731743.
Hodson, D. (1990). A critical look at practical working school science. School Science Review, 71, 3340.
Hofstein, A., & Lunetta, V. N. (1982). The role of the laboratory in science teaching: neglected aspects of
research. Review of Educational Research, 52, 201217.
Hofstein, A., & Lunetta, V. N. (2004). The laboratory in science education: foundations for the twenty-first
century. Science Education, 88, 2854.
Hofstein, A., Shore, R., & Kipnis, M. (2004). Providing high school chemistry students with opportunities to
develop learning skills in an inquiry-type laboratory: a case study. International Journal of Science
Education, 26, 4762.
Hofstein, A., Navon, O., Kipnis, M., & Mamlok-Naaman, R. (2005). Developing students ability to ask
more and better questions resulting from inquiry-type chemistry laboratories. Journal of Research in
Science Teaching, 42, 791806.
Hurd, P. D. (1969). New directions in teaching secondary school science. Chicago: Rand McNally.
Jones, M. E., Gott, R., & Jarman, R. (2000). Investigations as part of the key stage 4 science curriculum in
Northern Ireland. Evaluation and Research in Education, 14, 2337.
Kln, A. (2004). The opinions of Turkish high school pupils on inquiry-based laboratory activities. The
Turkish Online Journal of Educational Technology, 6, 4, 6, 5671.
Koballa, T. R., Jr. (1988). Attitude and related concepts in science education. Science Education, 72, 115
126.
Koballa, T. R., Jr., Crawley, F. E., & Shrigley, R. L. (1990). A summary of science education-1988. Science
Education, 74, 369381.
Lazarowitz, R., & Tamir, P. (1994). Research on using laboratory instruction in science. In D. L. Gabel (Ed.),
Handbook of research on science teaching (pp. 94127). New York: Macmillan.
Ledbetter, C. E. (1993). Qualitative comparison of students constructions of science. Science Education, 77,
611624.
Leonard, W. H., & Chandler, P. M. (2003). Where is the inquiry in biology textbooks? The American Biology
Teacher, 65, 485487.
Lin, J. (2007). Responses to anomalous data obtained from repeatable experiments in the laboratory. Journal
of Research in Science Teaching, 44(3), 506528.
Lord, T., & Orkwiszewski, T. (2006). Moving from didactic to inquiry-based instruction in a science
laboratory. The American Biology Teacher, 68, 342345.
Lunetta, V. N. (1998). The school science laboratory: historical perspectives and context for contemporary
teaching. In B. Fraser & K. Tobin (Eds.), International handbook of science education (pp. 249264).
Dordrecht: Kluwer.
Lunetta, V. N., Hofstein, A., & Giddings, G. G. (1981). Evaluating science laboratory skills. The Science
Teacher, 48, 2224.
National Research Council. (2000). Inquiry and national science education standards. Washington: National
Academy Press.
National Research Council. (2005). How students learn: Science in the classroom. Washington: The National
Academy Press.

Res Sci Educ (2013) 43:413435

435

National Research Council. (2006). Americas lab report: Investigations in high school science. Washington:
National Academies.
Reynolds, D. S., Doran, R. L., Allers, R. H., & Agruso, S. A. (1995). Alternative assessment in science: A
teachers guide. Buffalo: University of Buffalo.
Salta, K., & Tzougraki, C. (2004). Attitudes toward chemistry among 11th grade students in high schools in
Greece. Science Education, 88(4), 535547.
Sanger, M. J. (2000). Addressing student misconceptions concerning electron flow in aqueous solutions with
instruction including computer animations and conceptual change strategies. International Journal of
Science Education, 22, 521537.
Sanger, M. J., & Greenbowe, T. J. (1997). Common student misconceptions in electrochemistry: galvanic,
electrolytic, and concentration cells. Journal of Research in Science Teaching, 34(4), 377398.
Sanger, M. J., & Greenbowe, T. J. (1999). An analysis of college chemistry textbooks as sources of
misconceptions and errors in electrochemistry. Journal of Chemical Education, 76, 853860.
Schwab, J. J. (1962). The teaching of science as inquiry. In J. J. Schwab & P. F. Brandwein (Eds.), The
teaching of science (pp. 1103). Cambridge: Harvard University Press.
Stavy, R. (1991). Using analogy to overcome misconceptions about conservation of matter. Journal of
Research in Science Teaching, 28, 305313.
Tamir, P., Nussinovitz, R., & Friedler, Y. (1982). The design and use of practical tests assessment inventory.
Journal of Biological Education, 16, 4250.
Taraban, R., Box, C., Myers, R., Pollard, R., & Bowen, C. W. (2007). Effects of active-learning experiences
on achievement, attitudes, and behaviors in high school biology. Journal of Research in Science
Teaching, 44, 960979.
Tarhan (2008). Development of a material supported with active learning methods based on constructivism to
prevent formation and remediation of misconceptions in the subject of acids and bases in the level of
high school and university. Project supported by The Scientific and Technological Research Council of
Turkey (TBTAK) (Project number: TUB-105K058)
Tobin, K. G. (1990). Research on science laboratory activities. In pursuit of better questions and answers to
improve learning. School Science and Mathematics, 90, 403418.
Tsai, C. C. (2001). A review and discussion of epistemological commitments, metacognition, and critical
thinking with suggestions on their enhancement in internet-assisted chemistry classrooms. Journal of
Chemical Education, 78, 970974.
Tuan, H., Chin, C., & Shieh, S. (2005). The development of a questionnaire to measure students motivation
towards science learning. International Journal of Science Education, 27, 639654.
Williams, S. M., & Hmelo, C. E. (1998). Guest editors introduction. Journal of the Learning Sciences, 7,
265270.
Wu, H. K., & Hsieh, C. E. (2006). Developing sixth graders inquiry skills to construct explanations in
inquiry-based learning environments. International Journal of Science Education, 28, 11, 15, 1289
1313.

You might also like