Creative Education
2012. Vol.3, No.5, 619-631
Published Online September 2012 in SciRes (http://www.SciRP.org/journal/ce) http://dx.doi.org/10.4236/ce.2012.35091
Copyright © 2012 SciRes. 619
Addressing Misconceptions about the Particulate Nature of Matter
among Secondary-School and High-School Students in the
Republic of Macedonia
Marina I. Stojanovska1, Bojan T. Šoptrajanov2, Vladimir M. Petruševski1
1Institute of Chemistry, Faculty of Natural Sciences and Mathematics, Ss. Cyril & Methodius University,
Skopje, Macedonia
2Macedonian Academy of Sciences and Arts, Skopje, Macedonia
Email: marinam@pmf.ukim.mk
Received July 3rd, 2012; revised August 10th, 2012; accepted August 20th, 2012
A study was conducted to identify concepts about the particulate nature of matter among secondary- and
high-school students (N = 187) and to address some misconceptions regarding this topic, especially the
misunderstandings related to the vague ideas of the relationship between the macro and micro world. Data
were collected using both quantitative (six-item multiple-choice instrument in a pre-test-post-test design)
and qualitative (semi-structured focus group interviews) methods. Paired-samples t-test analysis showed
that students experienced significantly higher results in the post-test when compared to the pre-test, thus
confirming the efficiency of the intervention program in facilitating the understanding of some basic ele-
ments of the theory and practice concerning the particulate nature of matter (widely known as particle
theory concepts, a term which will be used in this paper as well) among students of different levels of
study. The findings revealed seven misconceptions prevalent by more than 20% of students and some ad-
ditional ones emerged from the in-depth focus group discussions. The analysis of the content of textbooks
indicated that some erroneous chemical concepts might have been formed as a result of the teaching of
chemistry and that of physics, as well. The use of animations and molecular models had a positive effect
on students and pointed to the need of introducing, in the chemistry teaching, the new material more visu-
ally.
Keywords: Focus Group Interviews; Intervention Program; Misconceptions; Particle Theory Concepts;
Particulate Nature of Matter; Secondary- and High-School Chemistry Education
Introduction
Chemistry is a subject based on concepts, many of which are
abstract and are therefore hard to grasp and learn especially
when the students are put in a position to believe without seeing.
On the other hand, students are basically familiar with a num-
ber of relevant concepts as a result of their previous learning
(Roschelle, 1995). The potentially present preconceptions about
the world itself can be reflected in the chemistry lessons and
can sometimes grow into misconceptions.
It has been known that the misconceptions are powerful, ex-
tremely persistent and hard to change, creating obstacles to
further learning (Canpolat, 2006; Pabuçcu & Geban, 2006).
Many misconceptions concerning various chemistry (and sci-
ence in general) topics have been documented (Horton, 2007;
Kind, 2004). Also, many studies in the field of science miscon-
ceptions and difficulties in learning and understanding chemical
concepts have been reported (Chiu, 2005; Cliff, 2009; Çalýk et
al., 2005; Kariper, 2011; Levy Nahum et al., 2004; Morgіl &
Yörük, 2006; Taber, 2011; Wenning, 2008). In some of them,
students’ drawings or pictorial representations within the test
item play an important role in inspecting their thinking (Mul-
ford & Robinson, 2002; Nyachwaya et al., 2011; Onwu &
Randal, 2006). The misconceptions related to the particulate
nature of matter are perhaps some of the most investigated ones
and are widely reported in literature (Badrian et al., 2011;
Banda et al., 2011; Bindernagel & Eilks, 2009; de Vos & Ver-
donk, 1996; Bridle & Yezierski, 2011; Gabel & Samuel, 1987;
Margel et al., 2004; Özmen & Kenan, 2007; Skamp, 2009;
Yezierski & Birk, 2006).
Many misconceptions are due to the fact that students do not
distinguish between macroscopic and microscopic explanations
(Bucat, 2004; Chandrasegaran et al., 2007; Meijer, 2011).
According to Johnstone (2000), the chemical knowledge is
acquired at three levels: a) the macro and tangible (what can be
seen, touched and smelt); b) the sub-micro (atoms, molecules,
ions and structures); and c) the representational (symbols, for-
mulae, equations, mathematical manipulation, graphs etc.) and
the links between these levels should be explicitly taught. None
of the above aspects is superior to another; on the contrary, they
complement each other. Aware of the conflict between the in-
tuitive, continuous model and the “scientific”, particulate one,
Ben-Zvi et al. (1986) pose the question: “Is an atom of copper
malleable?” In their study, involving students from all levels,
Treagust et al. (2011) display the widely-held confusion be-
tween the macroscopic and sub-microscopic properties of mat-
ter held by the students. Namely, the authors highlight the ideas
of students that particles (sub-microscopic representation)
themselves are compressed and pushed closer together instead
of the gas being compressed (macroscopic representation).
Introducing all three levels to students simultaneously leads
to an “overload of their working memory space” (Sirhan, 2007)
M. I. STOJANOVSKA ET AL.
and can be a potential reason for misconceptions to emerge.
Furthermore, the process of going directly from the macro-
scopic to the symbolic level, neglecting the sub-microscopic
one, may also be responsible for the appearance of misconcep-
tions. Therefore, chemistry concepts must be taught progres-
sively, stepwise, starting with macroscopic observation and
only then go to the microscopic interpretations.
Misconceptions that stem from the teaching process are re-
ferred to as school-made misconceptions (Barke et al., 2009).
Perhaps the most resistant to change are the misconceptions
that students build in the early stages of their development.
Some of the reasons for their occurrence could be traced to
problems of the specific terminology and used wording, espe-
cially when introducing the concepts of substances, the parti-
cles of which they consist and chemical symbols used for their
representation. Taber (2001) argues that definitions of most
fundamental concepts are problematic. Similarly, Nelson (2003)
stresses that many basic chemistry concepts are difficult to
teach because “the definitions of these concepts given in text-
books either lack precision, or invoke ideas that beginners are
not familiar with, and have to accept on trust”.
In this respect, the situation in many Macedonian textbooks
(for both secondary schools and high schools) is not very dif-
ferent. Thus, statements can be found in which a substance
reacts with one or more particles (atoms, molecules, ions ···),
such as: “An acid is any substance (molecule, anion or cation)
capable, during chemical reaction, to give a proton to another
substance.” (Cvetković, 2002b: p. 191), “By accepting a proton,
the former base becomes an acid.” (Šoptrajanov, 2002b: p. 74).
Similar (basically incorrect) notions are present in statements
such as: “When forming ionic compounds, iron gives three
electrons.” (Aleksovska & Antonovska, 2010: p. 142), “The
elementary substance phosphorus consists of four atoms of
element phosphorus.” (Aleksovska & Antonovska, 2010: p. 39),
“··· glycerol has three hydroxyl groups ···” (Aleksovska &
Stojanovski, 2003: p. 176), “Deoxyribonucleic acids are mac-
romolecules.” (Aleksovska & Stojanovski, 2005: p. 112) etc.
All of the above statements may be at the origin of formation
of new misconceptions or strengthening already existing ones.
Students holding deep-rooted misconceptions often construct
new knowledge on a faulty basis to which further mental con-
nections are made. Many students (and some of their teachers,
too) believe that their adopted concepts are correct because they
make sense and fit within their previously established mental
framework. Faced with new information that differs from their
established conception, a cognitive conflict arises (Demircioğlu,
2009), the students being put in a position to either change their
rather naive view into a scientifically accepted one (for this to
happen, a conceptual change must be developed first) or, oth-
erwise, to reject or ignore this new information simply because
it seems wrong. Active experimenting, use of animations and
multi-frame illustrations (and even drawings) could help
achieve this cognitive transformation (Kabapınar, 2009). It is of
utmost importance to identify the misconceptions of students,
using diagnostic tests, confront the erroneous notions applying
appropriate intervention program(s) expecting that the miscon-
ceptions will eventually be eliminated, corrected and replaced
with ones that are correct and stable.
Regarding the chemistry in Macedonia, only few investiga-
tions have been conducted so far that address the misconcep-
tions present among students (Monković et al., 2007; Stoja-
novska et al., 2012a; Stojanovska et al., 2012b, Petruševski et
al., 2006; Petruševski et al., 2007). A need for a more thorough
study of misconceptions in chemistry necessarily emerged. As a
consequence, the present study was planned to investigate the
potential misconceptions associated with the particulate nature
of matter among secondary- and high-school students in Ma-
cedonia. It focuses on in-depth understanding of concepts held
by students and on the implementation of an instructional pro-
gram aimed to help students in overcoming erroneous concepts
they have. The results from the multiple-choice item testing and
focus group discussions are summarized in the present paper.
Objectives of the Study
One of the objectives of this study was to check the capabil-
ity of students to transfer their knowledge through the three
levels of thinking as well as the ability to explain physical and
chemical properties mostly in terms of macro- and microscopic
approaches. At the same time, the study was intended to inves-
tigate the views of the students about the appearance and prop-
erties of substances, on one hand, and the entities they consist
of, on the other, and to address potentially present misconcep-
tions about the particulate nature of matter.
The study was conducted to evaluate the effectiveness of in-
tervention program on cognitive achievement towards particle
theory concepts of eighth-grade and high-school students. The
intervention program (that included deepened explanations,
molecular models, experiments, discussions and web anima-
tions) complemented with interviewing students was imple-
mented during instruction to facilitate understanding of particle
theory concepts among students and to attempt to eliminate
some of the misconceptions found.
The null hypothesis was tested at the 0.05 significance level
and was formulated as:
H0: There is no significant difference between the pre-test
and the post-test results.
The investigation was guided by the following research
questions:
1) Does the intervention program improve the achievement
of the students in the post-test?
2) Are there any trends in understanding the particle theory
concepts by the students across the various levels of study?
3) What are the misconceptions present in the students’
thinking regarding some issues on the particulate nature of
matter?
Methodology
Design
The study consisted of two parts: implementation of six-item
multiple-choice instrument in a pre-test-post-test design (quan-
titative approach) and realization of focus group interviews
(qualitative approach). It involved the evaluation of an instruc-
tional program in assessing the understanding of students about
concepts associated to particulate nature of matter. The con-
ceptual change intervention concerning the three levels of un-
derstanding consisted of active learning, including interactive
lectures using animations (BBC, 2012; Gregorius, n.d.), ex-
periments and discussions. We deemed the animations could be
helpful in visualizing atoms and molecules, their sizes and their
movement in a particular phase of a substance, as well as the
phase changes of the substance in question. Additional anima-
tions were found together with the students.
Copyright © 2012 SciRes.
620
M. I. STOJANOVSKA ET AL.
Copyright © 2012 SciRes. 621
The experiments carried out in front of (or together with) the
students followed by discussions were:
a) Freezing of water in a plastic cup and discussing phase
changes from the macro and micro points of view;
b) Burning of ethanol and discussing the fate of the C, H and
O atoms and;
c) Dissolving anhydrous copper(II) sulfate in water, evapo-
rating the solvent and comparing the chemical nature of the
substance(s) before and after the change.
The developmental stages of the investigation are briefly
stated as follows:
1) Administration of the pre-test;
2) Analysis of the pre-test data using the software package
PASW 18.0;
3) Recording misconceptions and identifying students hold-
ing those misconceptions;
4) Conducting focus group interviews;
5) Preparing transcripts and analyzing each interview;
6) Implementation of instruction program;
7) Administration of the post-test (two weeks after finishing
the stage 6);
8) Analysis of the post-test data using the software package
PASW 18.0.
Sample
The instrument was administered to a total of 187 sample
students both from a secondary school (N = 30) and from sev-
eral high schools (N = 157) in the 2011/12 school year. Further,
in order to get an insight into the particle theory concepts
among students of different age and the related misconceptions,
four sub-samples were formed according to the level of study in
which students were enrolled. Regarding the sampling, it is
mentioned (Hoque et al., 2011) that the minimum number of
subjects for experimental research is 30. Thus, the number of
students involved in this study was sufficient for further analy-
sis.
A total of 77 participants were purposively selected from the
sample students to take part in the group discussions. Accord-
ing to their earlier performance in chemistry, students were
categorized into two sub-groups (high achievers and low
achievers). Details concerning the participants involved in the
study are given in Table 1.
During the focus group interviews, mutual respect and con-
fidentiality was observed. The participants (students) were
carefully explained that the test results and interview discus-
sions would be used for research purposes only and that all
information would be kept confidential. For that reason, each
student was assigned a code (S1, S2 etc.) in the data analysis
process. Teachers approved the testing and interviewing stu-
dents. Students who agreed to participate in focus group dis-
cussions had the right to withdraw at any given moment and
were guaranteed that the research will not affect their grade.
Data Collection
A combination of quantitative and qualitative data collection
techniques was used and two kinds of instruments were imple-
mented: multiple-choice concept tests and focus group inter-
views. The usage of both kinds of methods led to an improve-
ment of the validity of results by means of triangulation (Hus-
sein, 2009; Jick, 1979). The advantages of the written test are
that it is easily administered to a wider population and that it
allows calculating the percentages of correct and incorrect an-
swers (Schmidt et al., 2003), thus enabling to draw conclusions
about the most prevalent choices that can indicate existence of a
misconception. Nevertheless, in many cases students who hold
misconceptions were able to give correct answers to open-
ended questions, especially to those that do not seek higher
mental activation. Therefore, it is advisable to use misconcep-
tions as distractors in a multiple-choice test (Herrmann Abell &
DeBoer, 2011) and, in that way, to provide students with plau-
sible answer choices to select from, so they are less likely to
simply guess the correct answer. Furthermore, interviews are an
appropriate technique when looking for in-depth explanations.
Interviews, both individual (Singh, 2008; Sözbilir et al., 2010)
and in a group (Eybe & Schmidt, 2004; Muzaffar et al., 2011;
Nair & Ngang, 2012; Osborne & Collins, 2001) have been
successfully used as data collection instruments in educational
studies, the focus groups providing a possibility for interaction
and exchanging ideas among participants.
Concept test: To identify notions that are present and to get
insight into the misconceptions regarding the particulate nature
of matter, a concept test was given both before and after group
discussions and intervention program (these are referred to as
pre-test and post-test respectively). The test was distributed to
secondary- and high-school students (Appendix 1). It was
conducted under normal class conditions and was completed
within 40 minutes. It consisted of 6 items written in a multi-
ple-choice format. No open-ended questions were provided for
testing because students (a part of them) were later interviewed
in a group, although occasionally an option existed to write
down their answers if the participants thought this would be
more appropriate. The test items were developed by the authors.
Some of the misconceptions known in the literature were added
as distractors (Horton, 2007; Kind, 2004; Anonimous, 2003;
Yezierski & Birk, 2006).
Focus group interviews: In this study, thirteen semi-struc-
Table 1.
Information on participants involved in the study.
Number of participants Number of focus groups
Level of study N (%) High-achievers Low-achievers
Secondary school VIII Grade 30 (16) 2 1
II Year 57 (31) 1 1
III Year 51 (27) 2 2
IV Year 49 (26) 2 2
High school
Total 187 (100) 13
M. I. STOJANOVSKA ET AL.
tured in-depth focus group interviews were conducted: three
with secondary-school and ten with high-school students. The
number of students in each group varied between 5 and 7 and
the time needed for completing an interview was in the range of
40 to 90 minutes. The interviews took place in an empty class-
room (or a laboratory) to enable uninterrupted conversation.
They were audio-taped and transcripts were made for each
focus group discussion.
The interviews were carried out according to the design pro-
posed by Kvale (1996). They consisted of three phases: the
briefing phase, the main phase and the debriefing phase. In the
main phase students were asked questions (in an open-ended
format) according to an interview guide prepared beforehand.
Details about the interview protocol are given in the Appendix
2. The misconceptions identified from the tests were used as a
starting point in the preparation of the interview guide and
conducting the interviews themselves. More questions emerged
during discussions depending on the answers of the partici-
pants.
Data analysis
In the concept test, each correct response was scored 1 point
and each incorrect one - 0 points, thus making the maximum
possible score on the test 6 points. Next, responses were ana-
lyzed using the Predictive Analytics Software PASW 18.0 and
were subjected to mean, standard deviations and significance
testing. In order to determine whether significance differences
were present between the pre- and post-test, a pared-sample
t-test was run.
Besides calculating the percentage of correct answers, the
number of wrong ones that were present in the answers by ap-
proximately 20% or above was considered. Namely, according
to Dhindsa & Treagust (2009), the incorrect responses (distrac-
tors) by more than 20% of the students indicate the presence of
misconception of the tested concepts. The above-mentioned
authors also proposed a criterion for satisfactory understanding
the tested concept in a four-distractors-item, according to which
it is necessary that at least 75% of the students give correct
answers.
In the present study, students preferred one or two distractors
for certain items. The misconceptions that were identified in the
answers of over 20% of the students are reported in the Results
section. The focus group interviews were not subject to a de-
tailed analysis and are used as representative for misconcep-
tions found in the written responses of the students. All inter-
views were conducted in Macedonian, so the excerpts quoted in
this paper were translated into English.
Results
The first research question was related to the effectiveness of
the intervention program on the improvement of the students’
achievement in the post-test. The comparison of the mean
scores of pre- and post-tests showed that an improvement in
cognitive achievement was evident (Figure 1), the positive
results being found for all sub-samples in the study. The results
of the paired-samples t-test analysis showed that that students
experienced significantly higher results in the post-test when
compared to the pre-test in each of the four sub-samples and in
the total sample (Table 2) which indicates that the intervention
program was efficient in improving the students’ knowledge
and understanding of particle theory concepts. Thus, the null
hypothesis was rejected.
In addition, the effect size (Cohen’s d) was calculated as a
measure of strength of the difference between the pre- and
post-test mean scores. The effect size values in Table 2 are
high (d > 0.8) implying relatively large mean differences for
each sub-sample and for the total sample, as well.
With respect to the second research question for the trends in
understanding of particle theory concepts by students across the
various levels of study, one can notice that the t-values increase
from eighth-grade to the fourth year high-school students’ sam-
ples; the second value disagrees with this trend by being higher
(cf. Table 2).
As a rough estimate of the concept-grasping by the students,
the percentage of correct answers (both in the pre-test and the
post-test) for all test items are given in the Table 3. The data
summarized in this table show that in all cases but one (for
eighth-grade students), an improvement in understanding can
be observed. Students had problems solving items 1, 4 and 6
and understanding the concepts involved, but were more suc-
cessful at answering the other items in the pre-test.
Figure 1.
Comparison of the pre-test—post-test mean scores.
Table 2.
Paired-samples t-test analysis results comparing pre-test and post-test score.
Pre-test Post-test
Level of study N Mean SD Mean SD t Cohen’s d
VIII 30 1.38 0.84 2.95 0.63 9.88* 2.11
II 57 1.29 0.97 4.32 1.44 14.60* 2.47
III 51 2.27 1.47 4.06 1.19 10.43* 1.34
IV 49 2.1 1.25 3.76 1.05 11.00* 1.44
Total 187 1.78 1.25 3.88 1.25 20.58* 1.68
*p < 0.01.
Copyright © 2012 SciRes.
622
M. I. STOJANOVSKA ET AL.
As for the third research question, some distractors were fa-
voured for certain items in the test and thus seven misconcep-
tions were identified in the answers given by students. Several
more misconceptions emerged on analyzing the interview tran-
scripts. The misconceptions found in the written responses are
listed here and are discussed in more detail in the next section.
Percentages of students holding the particular misconceptions
were also calculated (according the level of study, as well as the
total sample percentage) and are summarized in the Table 4.
No strict regularity in their appearance in different sub-sample
students can be detected from the percentages of the prevalent
misconception represented in the table. Misconceptions (M)
identified through the written responses of the students were:
M1: Butadiene reacts with two bromines.
M2: Butadiene reacts with one molecule of bromine.
M3: Copper(II) sulfate from aqueous solution crystallizes as
a solid substance composed of copper(II) sulfate molecules.
M4: In a process of ice melting, the volume of the system
(mixture of ice and liquid water) increases.
M5: Particles change from solid to liquid ones.
M6: When ethanol is ignited, carbon, hydrogen and oxygen
atoms (of which ethanol is composed), will ignite too and will
burn just like the ethanol.
M7: The particles of a substance enlarge during heating that
substance.
Three more misconceptions were detected in the test answers
of certain sub-samples. Thus, among eighth-grade students, a
notion that particles receive heat and enlarge was prevalent. In
the pre-test, 33.3% of students have chosen this option. The
percentage was not much reduced in the post-test (26.7%),
underlining the fact that misconceptions are very resistant and
hard to change. Two other misconceptions were observed in the
second-year high-school sub-sample analysis. A statement
commonly present among such students was that carbon, hy-
drogen and oxygen atoms (of which ethanol is composed), will
disappear when ethanol is ignited (31.6% in the pre-test and
7.0% in the post-test). Another favored discractor that repre-
sented a misconception was the one about particles of a sub-
stance shrinking during heating that substance. The percentage
of students that held this erroneous notion prior to instruction
was 28.1% and it was decreased to a great extent after the in-
struction (3.5% of students have chosen this option in the
post-test).
Discussion
The overall findings resulting from this research were posi-
tive and suggested that the intervention program effectively
contributed to the improvement in the knowledge of students
and their understanding of particle theory concepts. The con-
siderable increase in the post-test scores evidently supports this
observation. However, several misconceptions were still pre-
sent in the post-test. Some students, perhaps due to their poor
understanding, might have been misled by the distractors. Mis-
leading role of distractors was, presumably, the most pro-
nounced in the eighth-grade sample, whose participants (by
being the youngest) have limited knowledge. Also, beginners in
learning chemistry might not have well developed chemistry
concepts nor experience in chemistry and this could be assigned
to the constant level of misconception for certain test items for
eighth graders. Such was the case for the misconceptions re-
ferred as M1 and M2. These were distractors for the same test
item. It is likely students who were doubtful in answering this
test item have preferred the M1 in the post-test instead of their
Table 3.
Percentage of correct responses of students on pre- and post-test.
VIII II III IV Total
Item No. Pre Post Pre Post Pre Post Pre Post Pre Post
1 10 13.3 8.8 47.4 0 51 0 49 4.3 43.3
2 63.3 96.7 29.8 82.5 52.9 92.2 44.9 89.8 45.5 89.3
3 53.3 16.7 40.4 80.7 52.9 66.7 40.8 42.9 46 56.7
4 10 26.7 3.5 56.1 5.9 31.4 8.2 42.9 6.4 41.2
5 23.3 80 29.8 71.9 76.5 88.2 73.5 95.9 52.9 84
6 10 83.3 15.8 87.7 47.1 72.5 40.8 55.1 29.9 74.3
Table 4.
Percentage of misconceptions present in each of the sub-samples and the total sample.
VIII II III IV Total
Misconception No. Pre Post Pre Post Pre Post Pre Post Pre Post
M1 36.7 30 54.4 15.8 15.7 19.6 2 22.4 27.3 20.9
M2 40 56.7 29.8 28.1 72.5 27.5 98 28.6 61 32.6
M3 26.7 3.3 49.1 5.3 19.6 3.9 40.8 6.1 35.3 4.8
M4 40 83.3 42.1 12.3 31.4 15.7 44.9 26.5 39.6 28.3
M5 33.3 30 49.1 19.3 43.1 13.7 69.4 22.4 50.3 20.3
M6 36.7 10 31.6 17.5 21.6 9.8 18.4 0 26.2 9.6
M7 76.7 6.7 49.1 5.3 35.3 15.7 51 32.7 50.3 15.5
Copyright © 2012 SciRes. 623
M. I. STOJANOVSKA ET AL.
previously chosen option in the pre-test (M2) or the correct
answer (option “c”). Furthermore, data in Table 4 indicate that
about 28% of the third- and the fourth-year students showed
consistency in their answer regarding the first item (unfortu-
nately, the wrong one) by choosing the M2 in the post-test. On
the other hand, approximately 50% of each of these sub-sample
students (cf. Table 3) experienced positive gain in respect to
the tested concept on the first item (note that the percentage of
correct responses for the third- and the fourth-year students is
0.0% in the pre-test), thus reducing the level of both M1 and
M2 in the post-test. This was not the case with the eighth-grade
students.
Another observation is interesting to note at this point. It was
perceived that percent distribution for chosen test options was
more or less equally represented for eighth-grade students, thus
lowering the percent of correct answer for account of other
distractors (which were potential misconceptions). To justify
this fact two examples are pointed out. It was mentioned in the
Result Section that 33.3% of eighth-grade students in the pre-
test and 26.7% in the post-test have chosen the option that par-
ticles receive heat and enlarge (option “f” in the fourth test
item). The percent of chosen option “f” was 3.9% and 0.0% for
the third-year students and 2.0% and 0.0% for the fourth-year
ones in the pre- and post-test respectively. This could explain
the more pronounced level of M5 in the pre-test for students
enrolled in higher grades. The second example refers to the first
test item. It was noticed that 10.0% of eighth-grade students
thought that the statement “one gram butadiene reacts with one
gram bromine” correctly describes the equation. None of the
third- or the fourth-year students has chosen this option.
It is fair to mention that, when dealing with multiple-choice
items, students have an opportunity to guess the correct answer
or, on the contrary, to be misled by the distractors. For this
reason, the interviews were an important part of the research.
During the focus group discussions, students tried to explain
the statements they have chosen (or written) and, thus, helped
the researchers involved in this investigation to understand the
reasoning behind their opinions. Excerpts from the interview
transcripts that are presented in the paper could enable readers
to examine the trustworthiness of the research procedure.
The warm-up questions in the focus group discussions were
aimed at the (mental) images that students had about atoms and
molecules. On the basis of the experience in the microscopic
world, an opinion that atoms and molecules are spherical pre-
vailed as most of students recalled the illustrations they have
seen in textbooks or models that they have been using earlier.
Further, considering the phase changes, some students re-
ferred to molecular distances and molecular forces as factors
responsible for the change. Unfortunately, notions that mole-
cules of liquid (or gaseous) water weigh less that the molecules
of ice (or liquid water), that molecules can expand and stretch
when a substance is heated and that the particles (atoms and
molecules) can change their shape seem to be widespread even
among students that do not fall in the category of low achievers.
Here, an emphasis is given to those misconceptions found in
the written responses of students that were represented by more
than 20% (the list of such misconceptions was given in the
Results Section but in this part they are reviewed in more de-
tail). In addition, excerpts from the conversations between the
researcher (R) and student (S) are given to support the findings
from the test analysis. After every excerpt a brief description is
given that includes the type of school (SS or HS for secondary
and high school respectively), the level of education within the
school in question (given by Roman numerals) and the sub-
group to which the student belongs.
M1 and M2: Butadiene reacts with two bromines or with
one molecule of bromine.
These two misconceptions were among the most persistent
ones. The corresponding distractors were present with more
than 20% in the pre-test (see Table 4) and even in the post-test,
although the percentage of students who gave such answers
decreased from 27.3 to 20.9% for the first and from 61.0 to
32.6% for the second option. Therefore, they can be considered
to reflect potential misconceptions (M1 and M2). Although
some improvement was obvious, the percentage of students
holding erroneous views was still high, pointing once again to
the general property of misconceptions: lots of efforts are
needed to eliminate them and to engrave correct knowledge in
the minds of the students.
During the focus group discussion, students, for example,
explained: “I have calculated correctly: there are two atoms of
bromine and that is why butadiene reacts with two atoms of
bromine.” (SS-VIII-high achievers) or “I considered that Br2 is
one molecule of bromine, so my answer is that butadiene reacts
with one molecule of bromine.” (HS-II-low achievers)
These excerpts clearly point out to the vagueness of the ideas
of students and their inability to properly use the macro and
micro concepts and the corresponding terminology. Surely, a
substance can not react with only one or a few particles.
M3: Copper(II) sulfate from aqueous solution crystallizes
as a solid substance composed of copper(II) sulfate mole-
cules.
The second item of the test covers the concept of crystalliza-
tion of copper(II) sulfate from aqueous solution and the offered
options were: with five molecules of water; as a solid substance
composed of copper(II) sulfate molecules; as а pentahydrate; as
an anhydrous salt. We assumed that the first distractоr would be
an option of choice for the majority of students, but, instead, it
turned out that such was the case with the second distractor
(M3). The first option was chosen by less than 20% of the stu-
dents in the pre-test and the second one with 35.3%. This indi-
cates that most of the students have heard about crystalline
hydrates and can recognize the formula of blue vitriol, thus
connecting it with the term pentahydrate.
Sometimes erroneous notions are not perceptible right away
and can be discovered only when looking at the in-depth re-
sponses. The next dialog is representative of that.
R:S12, what is your opinion? You have answered that
copper(II) sulphate crystallizes as pentahydrate from water
solutions. Why do you think this is the correct answer?”
S12:Last year we learned that ··· er ··· copper sulphate
has five water molecules.”
R:Where? Five water molecules in ···?”
S12:In its structure.” (SS-VIII-high achievers)
Discussing within certain groups, some students claimed that
CuSO4·5H2O can be formed if five drops of water are added to
CuSO4. It seems likely that for these students five drops are
equivalent to five molecules (or five moles, perhaps?). In rela-
tion to this, Tasker (1998) warns about misleading molecular
animations available as CD supplements to some textbooks,
thus generating deeply embedded misconceptions such as the
one that assert presence of only a few ionic species in a drop of
water.
The presence of another very important misconception be-
Copyright © 2012 SciRes.
624
M. I. STOJANOVSKA ET AL.
comes apparent from the second item analysis. Namely, the
above mentioned high value of 35.3% strongly indicates that
students fail to distinguish ionic from covalent substances, so
they refer to all building particles as molecules. It seems that
this type of misconception is widely spread.
R:Do you think that molecules of copper(II) sulfate exist?”
S10:Yes.
R:Could you explain in more detail what do you mean?”
S10:I dont know ··· Im not sure.”
S9:I think that molecules of copper(II) sulfate do not exist
because copper (II) sulfate is a compound.”
R:What about water? Is water a compound?”
S9:Well ··· yes.”
R:Does it consist of molecules?
S9:Yes, it does.
R:When is it right to say molecules”?”
S12:When we talk about ions, it must be an ionic bonded
compound.” (HS-II-high achievers)
One possible reason for such a wrong concept can be found
in some textbooks (Cvetković, 2002b: p. 33) where the author
represents structural formulae of ionic compounds in the same
way as those of the covalent ones, with a line between symbols
of atoms implying the existence of mutual electron pairs. Many
teachers apparently accept this scheme and make calotte models
for every entity of every substance, including the ionic ones
(Aleksovska, 2011). In this way, they create a distorted picture
in the minds of their students and such a picture lasts for a very
long time (in many cases, we dare say, it is a lifelong miscon-
ception).
There are some definitions of crystalline hydrates (Alek-
sovska & Antonovska, 2010: p. 130) that use both macro and
micro viewpoints. The above textbook defines crystalline hy-
drates as: “A large number of salts obtained from aqueous solu-
tions, in their structure, contain one or more molecules of wa-
ter. These salts are called crystalline hydrates.” Obviously, if
such definitions are encountered in the very early stages of the
chemistry (or science) education, they might have harmful
consequences and make difficult the discrimination between
substances and their building blocks (particles).
M4: In a process of ice melting, the volume of the system
(mixture of ice and liquid water) increases and M5: Parti-
cles change from solid to liquid ones.
The most common misconceptions were those that reflected
ideas about the dimensions and shape of molecules when the
substance undergoes a phase change or any other change pro-
voked by the changes in temperature or pressure. We tried to
gauge the opinions that students have concerning very familiar
phenomena: ice melting and freezing of liquid water.
The total percentage of the incorrect answer (M4) decreases
in the post-test (cf. Table 4), as one should expect. However,
the scores of the eighth-grade students are not in agreement
with the overall outcome (the reasons for this finding are not
quite clear, but one could speculate that, being the youngest
group, their overall chemical knowledge is restricted).
To make items 3 and 4 more understandable, the researcher
and students discussed (through interviews and instruction)
about many everyday examples, such as the lakes freezing on
the surface, the density of ice cubes, the bursting of plastic
water-filled bottles when put in deep-freeze, the rock breaking
due to freezing of water inside the gaps etc. It was found
(Demircioğlu et al., 2005) that when chemistry concepts were
related to everyday life during teaching, their retention in the
learner’s mind was greater. On the other hand, Oloruntegbe et
al. (2010) reported that students could not relate science con-
cept learnt in school to home activities.
During focus group discussions, most students, explaining
the ice melting, recalled their everyday experiences but the
offered explanations were not always the scientifically accepted
ones: “My answer is that particles change from solid to liquid
ones. For example, if we put ice cubes in a drink, after same
time they will melt and turn to liquid state. According to this, I
thought that this answer is the most appropriate.” (SS-
VIII-high achievers)
The responses of students about changing the “state” of the
particles are reported in the literature (Griffiths & Preston,
1992). In his study, Taber (2001) claims that it is common for
learners to make statements such as: “The substance melt be-
cause its molecules melt.”
As indicated before, 33.3% of secondary (eighth-grade)
school students thought that particles receive heat and change
their size. This view was clearly shown in the following ex-
cerpts from one of the secondary school groups’ discussion:
An ice cube ··· when exposed to sunlight and heat, starts to
melt and its particles broaden.” (SS-VIII-low achievers) or “I
think that particles in ice are smaller than those in
water ··· and when ice begins to melt, the volume decreases.
(SS-VIII-low achievers)
M6: When ethanol is ignited, carbon, hydrogen and oxy-
gen atoms, of which ethanol is composed, will ignite too and
will burn just like the ethanol.
The fifth item on the test deals with the nature of atoms in a
chemical reaction. As an example of chemical reaction we con-
sidered the combustion of ethanol. The concept of burning or
combustion of certain substances can be found in the
eighth-grade curriculum, and in the subsequent education it is
studied in more details. Explanations about reactions such as
burning of magnesium, ethyne or ethanol as well as the chemi-
cal equations for the corresponding reactions can be found in
many textbooks. One should, therefore, expect that students
will be able to write down a chemical equation for this reaction
and give a proper explanation. It was found, however, that stu-
dents hold both tested misconceptions: (a) 26.2% of 187 stu-
dents thought that atoms may be ignited and burned (M6) and
(b) 14.4% thought that atoms disappear during chemical reac-
tions. The latter was found to be the most prevalent option in
the second-year high-school sub-sample.
However, during interviewing we could notice some failures
in knowledge and vagueness about the students’ perceptions
about atoms. These are illustrated bellow.
R:Why do you think that atoms will burn?”
S15:Well, because ethanol has oxygen in its structure and
it is able to burn.” (SS-VIII-high achievers)
Some other type of misconceptions regarding the phase
changes and the size of the particles were also documented.
Some students believed that the particle size must be changed,
while others failed in distinguishing between physical and
chemical change. Here are comments of the students regarding
these issues.
R:What will happen with the particles when a substance is
ignited?”
S21:They change their phase.”
S22:I stick to my opinion that they will either increase their
size or decrease it, because when they undergo a phase transi-
tion their shape will be changed. They must be changed; they
Copyright © 2012 SciRes. 625
M. I. STOJANOVSKA ET AL.
can not stay the same. In such a case, there will be no chemical
change, nor will a phase transformation occur.” (HS-III-low
achievers)
S3:They will get into some other substances because etha-
nol passes from liquid to gas during burning.”
S17:Since molecular space is changing as a result of
burning, their size must change, too.” (HS-III-high achievers)
The findings show that some of the students retained their
misconception, although they were faced with opposing opin-
ions during focus group discussions.
M7: The particles of a substance enlarge during heating
that substance.
This distractor was once again used to test the consistence in
opinion of students concerning the size of the particles when
the substance is heated. Several interesting (but not scientifi-
cally correct) opinions are presented to support the presence of
this misconception.
S11:I think that ··· when water evaporates, molecules
spread out because of the temperature ··· they are heated. I
think they enlarge.” (SS-VIII-high achievers)
R:Will the molecule of water at 80 C be smaller, bigger or
equally large to the one of water at 20 C?”
S11:I think that when temperature changes, particle size
decreases.” (HS-II-high achievers)
S70:When heating an object, particles become wider.
(HS-IV-high achievers)
S8:When heat gets inside the particles, they are heated and,
due to that, increase their size.” (SS-VIII-low achievers)
On the other hand, some students claimed that molecules of
ice are bigger than those of liquid water, in consideration of the
fact that the volume of ice is bigger when compared to that of
the same amount of liquid water.
R:Y ou said that freezing will make the water molecule lar-
ger. Could you explain what did you mean?”
S2:Well ··· the volume of ice ··· water in the solid state is
larger.
R:Do you think that if the volume is larger, the molecule
must be larger, too?”
S2:Thats right.” (HS-II-low achievers)
Several more categories of erroneous notions were registered
during the focus group discussions. Fortunately, those were not
widely spread among students. Some of them are the following
ones:
S29:Molecules that evaporate are lighter and go upward
and the ones that stay in the vessel are heavier.” (HS-III-low
achievers)
S32: [When water evaporates] “a vapor is formed and water
molecules are decomposed into hydrogen and oxygen.” (HS-III-
low achievers)
There was a good attempt by some students to give correct
explanation in line with the fact that in many chemistry text-
books used in Macedonia (Cvetković, 2002b: p. 130; Šoptraja-
nov, 2002a: p. 113; Šoptrajanov, 2002b: p. 62) a number of
illustrations indicate that when a substance changes its phase,
the space/distance between particles is changing, not the parti-
cles themselves. On the other hand, some less precise state-
ments can also be found. One of them is the following: “Be-
cause of that, molecules in a liquid can not substantially move
apart from one another and turn into vapor” (Cvetković, 2002b:
p. 132) indicating that molecules can turn into a gaseous state.
Similar findings are reported in the literature (Tatar, 2011).
According to Taber (2001), the students have represented the
longest O-H bond in water molecule in the gaseous state of the
substance, explaining the “fact” that the rise of the volume (due
to the phase change) implies increasing the size of the molecule
itself. Other studies (Mayer, 2011; Regan et al., 2011) notify
that, when asked about water evaporation, most of the students
have chosen the drawing according to which the water mole-
cules separate into hydrogen and oxygen atoms.
At the end of the focus group discussion, as a conclusion, the
following question was asked: “Will the particles (atoms,
molecules or ions) possess similar properties as the parent ma-
terial?” There were different ideas coming from students, but
some improvements in thinking regarding their earlier stand-
points were noticeable (see below).
S14:It is my opinion ··· that this statement is true. It is said
in chemistry textbooks ··· that the atom is the smallest particle
that possesses the same physical and chemical properties as the
substance it originates from.” (HS-III-high achievers)
S2: Well ··· because we are dealing with one substance, I
think that the atoms will have the same physical and chemical
properties ··· because substances are made up of particles.”
(SS-VIII-high achievers)
S5:Atoms and all the other particles of which the sub-
stances are composed do not have the same properties as com-
pounds have, because ··· properties of that particular element
or compound are added to the atoms.” (SS-VIII-high achievers)
S50:In my opinion, molecules do not change, only sub-
stances change their state.” (HS-III-high achievers)
S64:Physical properties are not the same. For example, we
can see and touch sulfur, but not its atoms. So, atoms and
molecules of sulfur as a whole constitute one object that has
different properties from those characteristic of atoms and
molecules themselves.” (HS-IV-high achievers)
It is likely that some erroneous concepts are formed in the
teaching of physics, as well since imprecise definitions present
in some physics textbooks can also lead to appearance of mis-
conceptions that are evident in testing of chemistry concepts.
Namely, according to Gešoski & Nonkulovski (2009) “All
substances are composed of small invisible particles that pos-
sess the same properties as the parent material. These particles
are called molecules.” With respect to the atom, the authors
have written: “The atom is the smallest particle of a chemical
element that retains its physico-chemical properties.” In fact, an
analogous statement is found in the chemistry textbook for 7th
grade (Cvetković, 2002a: p. 35) where it is written “The small-
est particles that posses all properties of iron, copper and mer-
cury and that can not be divided further are called atoms.”
Similar definitions were reported by Papageorgiou & Sakka
(2000). It is more than obvious that some adjustments need to
be done in clarifying similar concepts present in the teaching of
chemistry and physics.
Conclusion and Implications for Teaching
Several points need to be addressed concerning this study.
The higher scores in the post-test showed that the intervention
program was efficient, enabling students to gain more scientific
explanations. The overall instruction was successful in facili-
tating understanding of particle theory concepts among students
of different levels of study. In order for students to gain better
knowledge and understanding, the researchers used lecture,
discussions, written notes and clarifications and demonstrated
several experiments. The improvement of the mean scores was
Copyright © 2012 SciRes.
626
M. I. STOJANOVSKA ET AL.
shown to be statistically significant for all sub-samples. Fur-
thermore, the better results in the post-tests (for all samples) are
in favor of the conclusion that indeed there is a noticeable im-
provement in the achievements of students for all test items
(except for one in the sample of eighth-grade students).
With respect to the trends in understanding the particle the-
ory concepts by students across the various levels of study, an
increase of the t-values was noticed, although the second value
in Table 2 is opposite to this reasoning. Nevertheless, it was
practically impossible to detect a strict regularity about the
learning progress and the achievement of students since there
was not a big difference in understanding of students of differ-
ent age. Similar results are reported in the literature (Özmen,
2011).
One of the main points of this research was to detect misun-
derstandings and difficulties regarding the particle theory con-
cepts, especially those related to the three levels of thinking
(macroscopic, sub-microscopic and symbolic). Seven miscon-
ceptions represented by more than 20% were registered among
students of all levels of study and three additional misconcep-
tions were found in certain sub-samples. Many students
claimed that the particles either decrease or increase their size
when a substance is heated, their notions originating (errone-
ously) from the observations of the macroscopic properties of
substances. Approximately 35% of the students fail to recog-
nize the ionic nature of copper(II) sulfate and even 88% of
them believe that the substance (butadiene) reacts with one or
more entities (atoms or molecules of bromine). At this point, it
must be emphatically stated that neither the substances are
molecules nor molecules are substances and such statements
should be explicitly given in the textbooks and repeated, over
and over again, by the teachers in the classrooms. In this re-
spect, it is of little comfort that even the Nobel Foundation
reported that the Nobel Prize in Physiology or Medicine 1998
was awarded jointly to Robert F. Furchgott, Louis J. Ignarro
and Ferid Murad “for their discoveries concerning nitric oxide
as a signalling molecule in the cardiovascular system” (The
Nobel Prize in Physiology or Medicine 1998), thus showing
how widespread is the imprudent mixing of macroscopic (nitric
oxide, a substance) and sub-microscopic (molecule, a particle)
points of view.
We tried to get insight into ideas of the students about the
particles and their connection to substances using a concept test
and group discussions. The test items (which were developed
by the authors) are provided in Appendix 1 and can be used to
repeat the research to a different sample and to compare the
findings. Although generalization beyond the data of this sam-
ple can not be guaranteed, some considerations could be valu-
able. In this study, we found out that a large portion of students
held the belief that the molecules have the properties of the
bulk matter. The reasons for emergence of these misconcep-
tions are mainly due to the failure to distinguish between mac-
roscopic and microscopic properties (Ben-Zvi & Gai, 1994;
Bucat, 2004; Chandrasegaran et al., 2007; Treagust et al., 2011).
Furthermore, the data analysis from the tests and focus group
interviews clearly showed that students had certain difficulties
about: 1) recognizing symbolic representations (e.g. confusing
symbols for atoms and formulae for molecules); 2) making
distinction between ionic and covalent substances and their
particles and 3) the ability to distinguish substances from
molecules (i.e. differentiate concepts at macroscopic and mi-
croscopic level).
The low understanding about the particle theory concepts
may originate from the unstable previous knowledge of stu-
dents and emphasising the rote learning (Dhindsa & Treagust,
2009; Salame et al., 2011) without making connections between
tangible macroscopic phenomena and changes at the molecular
level. The abstract nature of many chemical terms (such as
atoms, molecules, ions, electrons etc.) may also be a source of
misconceptions. The fact that we are not able to see particles
and their interconnection and interaction, compels us to apply a
model which would help in acquiring understanding. As Coll &
Treagust (2003) noted “To understand and explain the macro-
scopic, we need to have a picture or image-a model-of what is
going on at the microscopic level. Therefore, the necessity of
using visualization techniques (models, animations or computer
software) is obvious and it was mentioned as a required tool in
teaching abstract concepts (José & Williamson, 2005; Milne,
1999). However, introducing models or images to students
must be done with caution; otherwise it could be unproductive
or even counterproductive. Eilks et al. (2012), with reference to
their earlier studies, state that curriculum developers and text-
book authors not always carefully incorporate important re-
search evidence when preparing learning materials, as a result
of which potentially misleading visualisations can be found in
some German textbooks.
Particle theory concepts are an integral part of the eighth-
grade secondary-school curriculum in Republic of Macedonia.
At the very beginning of learning chemistry as a subject, stu-
dents encounter the bulk properties (physical and chemical) of
substances, their changes, classifications etc. Next, they learn
about the structure of substances. Chemical symbols, formulae
and equations come later. Nonetheless, the results showed that
the sample of students have not developed an accurate under-
standing of these concepts. Furthermore, a remarkable consis-
tency in reasoning was observed among students of different
levels of study, thus confirming the fact that many students
retain their misconceptions over the years.
Therefore, it is of utmost importance to make a clear differ-
entiation between sub-microscopic entities and macroscopic
phenomena in chemistry (Silberberg, 2006) and, consequently,
their appropriate implementation in chemistry teaching. By all
means, a special caution should be taken when teaching the
youngest students in order to enable them to create proper
mental images concerning the basic chemical concepts.
There are, at least, two main starting points useful in teaching
chemistry: a) the “cognitive conflict” strategy that includes the
use of diagnostic questions whose purpose is to stimulate stu-
dents to talk about their ideas and utilizing the answers of the
participating students to establish and organize teaching and b)
carefully introducing the new material (ideas or concepts) visu-
ally, using pictures, animations and models of atoms and
molecules and presenting new ideas consistently in other topics,
thus forming an all-inclusive, integrated whole, instead of
teaching different topics isolated.
Unfortunately, within the educational system in Macedonia
many obstacles exist (including financial, experimental, per-
sonal prerequisites and, especially, the time allowed to the
teachers) which would hamper the implementation of our deep
conviction outlined above.
Our future studies will be aimed at other aspects of this (or a
similar) research. An important issue will be testing the influ-
ence of gender on achievements of students. Preparing teaching
materials to reduce erroneous notions (and help teacher to di-
Copyright © 2012 SciRes. 627
M. I. STOJANOVSKA ET AL.
agnose and address misconceptions before the instruction proc-
ess) is deemed equally important.
Acknowledgements
The authors wish to express their sincere thanks to Prof. Bo-
rislav Toshev, D.Sc. (University of Sofia, Bulgaria) for critical
reading of the manuscript and the valuable suggestions he
made.
REFERENCES
Aleksovska, S., & Antonovska, L. (2010). Chemistry for the 7th grade
of eighth-year secondary education. Skopje: Ministry of Education
and Science of Republic of Macedonia (in Macedonian).
Aleksovska, S., & Stojanovski, K. (2003). Chemistry for the 3rd year of
reformed gymnasium education. Skopje: Prosvetno Delo (in Mace-
donian).
Aleksovska, S., & Stojanovski, K. (2005). Chemistry for the 4th year of
reformed gymnasium education. Skopje: Prosvetno Delo (in Mace-
donian).
Aleksovska, S. (2011). Personal communication.
Anonimous (2003). Common misconceptions about matter. URL (last
checked 2 July 2012).
http://www.arps.org/users/ms/pricen/one%20pagers/common_misco
nceptions%20chemistry.pdf
Badrian, A., Abdinejad, T., & Naseriazar, A. (2011). A cross-age study
of Iranian students’ various conceptions about the particulate nature
of matter. Journal of T u rkish Science Education, 8 , 49-63.
Banda, A., Mumba, F., Chabalengula, V. M., & Mbewe, S. (2011).
Teachers’ understanding of the particulate nature of matter: The case
of Zambian pre-service science teachers. Asia-Pacific Forum on
Science Learning and Teaching, 1 2 , Article ID: 4.
Barke, H. D., Hazari, A., & Yitbarek, S. (2009). Chapter 2: Students’
misconceptions and how to overcome them. In: Misconceptions in
chemistry addressing perceptions in chemical education (pp. 21-36).
Berlin: Springer-Verlag.
BBC (2012). KS3 Bitesize: Science—Behaviour of matter. URL (last
checked 2 July 2012).
http://www.bbc.co.uk/schools/ks3bitesize/science/chemical_material_b
ehaviour/behaviour_of_matter/activity.shtml
Ben-Zvi, N., & Gai, R. (1994). Macro- and micro-chemical compre-
hension of real-world phenomena. Journal of Chemical Education,
71, 730-732. doi:10.1021/ed071p730
Ben-Zvi, R., Eylon, B., & Silberstein, J. (1986). Is an atom of copper
malleable? Journal of Chemical Education, 63, 64-66.
doi:10.1021/ed063p64
Bindernagel, J. A., & Eilks, I. (2009). Evaluating roadmaps to portray
and develop chemistry teachers’ PCK about curricular structures
concerning sub-microscopic models. Chemistry Education Research
and Practice, 10, 77-85. doi:10.1039/b908245j
Bridle, C. A., & Yezierski, E. J. (2011). Evidence for the effectiveness
of inquiry-based, particulate-level instruction on conceptions of the
particulate nature of matter. Journal of Chemical Education, 89, 192-
198. doi:10.1021/ed100735u
Bucat, B. (2004). Pedagogical content knowledge as a way forward:
Applied research in chemistry education. Chemistry Education Re-
search and Practice, 5, 215-228. doi:10.1039/b4rp90025a
Çalýk, M., Ayas, A., & Ebenezer, J. V. (2005). A review of solution
chemistry studies: Insights into students’ conceptions. Journal of
Science Education and Techno l ogy, 14, 29-50.
Canpolat, N. (2006). Turkish undergraduates’ misconceptions of eva-
poration, evaporation rate, and vapour pressure. International Jour-
nal of Science Education, 28 , 1757-1770.
doi:10.1080/09500690600779957
Chandrasegaran, A. L., Treagust, D. F., & Mocerino, M. (2007). The
development of a two-tier multiple-choice diagnostic instrument for
evaluating secondary school students’ ability to describe and explain
chemical reactions using multiple levels of representation. Chemistry
Education Research and Practice, 8, 293-307.
doi:10.1039/b7rp90006f
Chiu, M.-H. (2005). A national survey of students’ conceptions in
chemistry in Taiwan. Chemical Education Internationa l, 6, 8 p.
Cliff, W. H. (2009). Chemistry misconceptions associated with under-
standing calcium and phosphate homeostasis. Advances in Physiol-
ogy Education, 33, 323-328. doi:10.1152/advan.00073.2009
Coll, R. K., & Treagust, D. F. (2003). Investigation of secondary school,
undergraduate, and graduate learners’ mental models of ionic bond-
ing. Journal of Research in Science Teaching, 40, 464-486.
doi:10.1002/tea.10085
Cvetković, S. (2002a). Chemistry for the 7th grade. Skopje: Prosvetno
Delo (in Macedonian).
Cvetković, S. (2002b). Chemistry for the 1st year of reformed gymna-
sium education. Skopje: Prosvetno Delo (in Macedonian).
De Vos, W., & Verdonk, A. H. (1996). The particulate nature of matter
in science education and in science. Journal of Research in Science
Teaching, 33, 657-664.
doi:10.1002/(SICI)1098-2736(199608)33:6<657::AID-TEA4>3.0.C
O;2-N
Demircioğlu, G. (2009). Comparison of the effect of conceptual change
texts implemented after and before instruction on secondary school
students’ understanding of acid-base concepts. Asia-Pacific Forum
on Science Learning and Teaching, 10, Article ID: 5.
Demircioğlu, G., Ayas, A., & Demircioğlu, H. (2005). Conceptual
change achieved through a new teaching program on acids and bases.
Chemistry Education Research and Practi ce, 6, 36-51.
doi:10.1039/b4rp90003k
Dhindsa, H. S., & Treagust, D. F. (2009). Conceptual understanding of
Bruneian tertiary students: Chemical bonding and structure. Brunei
International Journal of Science & Mathematics Education, 1, 33-51.
Eilks, I., Witteck, T., & Pietzner, V. (2012). The role and potential
dangers of visualisation when learning about sub-microscopic ex-
planations in chemistry education. Centre for Educational Policy
Studies Journal, 2, 125-145.
Eybe, H., & Schmidt, H. (2004). Group discussions as a tool for inves-
tigating students’ concepts, Chemistry Education Research and
Practice, 5, 265-280. doi:10.1039/b4rp90028f
Gabel, D. L., & Samuel, K. V. (1987). Understanding the particulate
nature of matter. Journal of Chemical Educatio n, 64, 695-697.
doi:10.1021/ed064p695
Gešoski, S., & Nonkulovski, F. (2009). Physics for the 7th grade of
eighth-year-secondary education. Skopje: Ministry of Education and
Science of Republic of Macedonia (in Macedonian).
Gregorius, R. M. (n.d.). An eBook in flash to support inductive learning.
URL (last checked 2 July 2012).
http://www3.canisius.edu/~gregorir/ic2home/eBookPaper.pdf
Griffiths, A. K., & Preston, A. K. (1992). Grade-12 students’ miscon-
ceptions relating to fundamental characteristics of atoms and mole-
cules. Journal of Research in Science Teaching, 29, 611-628.
doi:10.1002/tea.3660290609
Herrmann Abell, C. F., & DeBoer, G. E. (2011). Using distractor-given
standard-based multiple-choice assessment and Rasch modelling to
investigate hierarchies of chemistry misconceptions and detect
structural problems with individual items. Chemistry Education Re-
search and Practise, 12, 184-192. doi:10.1039/c1rp90023d
Hoque, K. E., Alam, G. M., Ariff, M. R. B. M., Mishra, P. K., & Rabby,
T. G. (2011). Site-based management: Impact of leader’s roles on in-
stitutional improvement. African Journal of Business Management, 5,
3623-3629.
Horton, C. (With Other Members of the Modeling Instruction in High
School Chemistry Action Research Teams at Arizona State Univer-
sity) (2007). Student Alternative Conceptions in Chemistry. URL
(last checked 2 July 2012).
http://www.scribd.com/doc/52664732/student-alternative-conception
s-in-chemistry
Hussein, A. (2009). The use of triangulation in social sciences research:
Can qualitative and quantitative methods be combined? Journal of
Comparative Soci al Work, 1, 1-12.
Jick, T. D. (1979). Mixing qualitative and quantitative methods: Trian-
gulation in action. Administrative Science Quaterly, 24, 602-611.
Copyright © 2012 SciRes.
628
M. I. STOJANOVSKA ET AL.
doi:10.2307/2392366
Johnstone, A. H. (2000). Teaching of chemistry—Logical or psycho-
logical? Chemistry Education Research and Practice in Europe, 1,
9-15. doi:10.1039/a9rp90001b
José, T. J., & Williamson, V. M. (2005). Molecular visualization in
science education: An evaluation of an NSF-sponsored workshop.
Journal of Chemical Ed uc a t io n , 82, 937-943.
doi:10.1021/ed082p937
Kabapınar, F. M. (2009). Multi-frame illustrations: A molecular visual
strategy in learning and teaching chemistry concepts. Australian
Journal of Education i n C h em i s t r y , 69, 11-16.
Kariper, İ. A. (2011). An investigation into the misconceptions, erro-
neous ideas and limited conception of the pH concept in pre-service
science teacher education. The Chemical Education Journ a l, 14.
Kind, V. (2004). Beyond appearances: Students’ misconceptions about
basic chemical ideas (2nd ed.). Durham: Durham University. URL
(last checked 2 July 2012).
http://www.rsc.org/images/Misconceptions_update_tcm18-188603.p
df
Kvale, S. (1996). Chapter 7: The interview situation. In Interviews: An
Introduction to Qualitative Research Interviewing (pp. 124-135).
London: Sage.
Levy Nahum, T., Hofstein, A., Mamlok-Naaman, R., & Bar-Dov, Z.
(2004). Can final examinations amplify students’ misconceptions in
chemistry? Chemistry Education Research and Pract ice, 5, 301-325.
Margel, H., Eylon, B.-S., & Scherz, Z. (2004). “We actually saw atoms
with our own eyes” conceptions and convictions in using the scan-
ning tunneling microscope in junior high school. Journal of Chemi-
cal Education, 81, 558-566. doi:10.1021/ed081p558
Mayer, K. (2011). Addressing students’ misconceptions about gases,
mass, and composition. Journal of Chemical Ed u c a ti o n , 88, 111-115.
doi:10.1021/ed1005085
Meijer, M. R. (2011). Macro-meso-micro thinking with structure-
property relations for chemistry education. An explorative design-
based study. Ph.D. Thesis, Utrecht: Utrecht University.
Milne, R. W. (1999). Animating reactions. A low-cost activity for par-
ticle conceptualization at the secondary level. Journal of Chemical
Education, 76, 50-51. doi:10.1021/ed076p50
Monković, M., Ivanovski, V., & Petruševski V. M. (2007). Demonstra-
tions as a tool for ironing-out misconceptions: 3. A note on the reac-
tion Pb2+(aq) + 2I(aq) = PbI2(s): Exceptions from a priori expecta-
tions. The Australian Journal of Education in Chemistry, 68, 17-19.
Morgіl, İ., & Yörük, N. (2006). Cross-age study of understanding of
some concepts in chemistry subjects in science curriculum. Journal
of Turkish Science Educat i o n, 3, 15-27.
Mulford, D. R., & Robinson, W. R. (2002). An inventory for alternate
conceptions among first-semester general chemistry students. Jour-
nal of Chemical Education, 79, 739-744. doi:10.1021/ed079p739
Muzaffar, H., Castelli, D. M., Goss, D., Scherer, J. A., & Chapman-
Novakofski, K. (2011). Middle school students want more than
games for health education on the internet. Creative Education, 2,
393-397. doi:10.4236/ce.2011.24056
Nair, S., & Ngang, T. K. (2012). Exploring parent’s and teacher’s
views of primary pupils’ thinking skills and problem solving skills.
Creative Education, 3, 30-36. doi:10.4236/ce.2012.31005
Nelson, P. G. (2003). Basic chemical concepts. Chemistry Education
Research and Practic e, 4, 19-24. doi:10.1039/b2rp90033e
Nyachwaya, J. M., Mohamed, A., Roehrig, G. H., Wood, N. B., Kern,
A. L., & Schneider, J. L. (2011). The development of an open-ended
drawing tool: An alternative diagnostic tool for assessing students’
understanding of the particulate nature of matter. Chemistry Educa-
tion Research and Practice, 12, 121-132. doi:10.1039/c1rp90017j
Oloruntegbe, K. O., Ikpe, A., & Kukuru, J. D. (2010). Factors in stu-
dents’ ability to connect school science with community and real-
world life. Ed uc a t io n a l Research and Reviews, 5, 372-379.
Onwu, G. O. M., & Randal, E. (2006). Some aspects of students’ un-
derstanding of representational model of the particulate nature of
matter in chemistry in three different countries. Chemistry Education
Research and Practic e, 7, 226-239. doi:10.1039/b6rp90012g
Osborne, J., & Collins, S. (2001). Pupils’ views of the role and value of
the science curriculum: A focus group study. International Journal
of Science Education, 23 , 441-467. doi:10.1080/09500690010006518
Özmen, H., & Kenan, O. (2007). Determination of the Turkish primary
students’ views about the particulate nature of matter. Asia-Pacific
Forum on Science Learning and Teaching, 8, Article ID: 1.
Özmen, H. (2011). Turkish primary school students’ conceptions about
the particulate nature of matter. Internationa Journal of Environ-
mental & Science Education, 6, 99-121.
Pabuçcu, A., & Geban, Ö. (2006). Remediating misconceptions con-
cerning chemical bonding through conceptual change text. Hacettepe
University Journal of Education, 30, 184-192.
Papageorgiou, G., & Sakka, D. (2000). Primary school teschers’ views
of fundamental chemical concepts. Chemistry Education Research
and Practice, 1, 237-247. doi:10.1039/a9rp90025j
Petruševski, V. M., Monković, M., & Šoptrajanov, B. (2007). Demon-
strations as a tool for ironing-out preconceptions: 1. On the reactions
of alkali metal sulfates with concentrated sulfuric acid. The Chemical
Educator, 12, 1-4.
Petruševski, V. M., Taseska, M., & Monković, M. (2006). Demonstra-
tions as a tool for ironing-out misconceptions: 2. Can a strong acid be
displaced by a weak one? The Chemical Education Journal, 9 .
Regan, A., Childs, P., & Hayes, S. (2011). The use of an intervention
programme to improve undergraduate students’ chemical knowledge
and address their misconceptions, Chemistry Education Research
and Practice, 12, 219-227. doi:10.1039/c1rp90027g
Roschelle, J. (1995). Learning in interactive environments: Prior knowl-
edge and new experience. In: J. H. Falk, & L. D. Dierking (Eds.),
Public institutions for personal learning: Establishing a research
agenda (pp. 37-51). Washington DC: American Association of Mu-
seums.
Salame, I. I., Sarowar, S., Begum, S., & Krauss, D. (2011). Students’
alternative conceptions about atomic properties and the periodic table.
The Chemical Educator, 16, 190-194.
Schmidt, H., Baumgärtner, T., & Eybe, H. (2003). Changing ideas
about the periodic table of elements and students’ alternative concep-
tions of isotopes and allotropes. Journal of Research in Science
Teaching, 40, 257-277. doi:10.1002/tea.10076
Silberberg, M. S. (2006). Chemistry. The molecular nature of matter
and change (4th ed.). Boston, MA: McGraw-Hill.
Singh, P. (2008). The unexpected reward of qualitative research in
assessment: A case example. The Qualitative Report, 13, 278-300.
Sirhan, G. (2007). Learning difficulties in chemistry: An overview.
Journal of Turkish Science Education, 4, 2-20.
Skamp, K. (2009). Atoms and molecules in primary science: What are
teachers to do? Australian Journal of Education in Chemistry, 69, 5-
10.
Šoptrajanov, B. (2002a). Chemistry for the 1st year of reformed gym-
nasium education. Skopje: Prosvetno Delo (in Macedonian).
Šoptrajanov, B. (2002b). Chemistry for the 2nd year of reformed gym-
nasium education. Skopje: Prosvetno Delo (in Macedonian).
Sözbilir, M., Pınarbaşı, T., & Canpolat, N. (2010). Prospective chemis-
try teachers’ conceptions of chemical thermodynamics and kinetics.
Eurasia Journal of Mathematics, Science & Technology Education, 6,
111-121.
Stojanovska, M. I., Petruševski, V. M., & Šoptrajanov, B. T. (2012a).
On the existence of hydrogen salts of monoprotic acids. Journal of
Chemical Education, 89, 1168-1170. doi:10.1021/ed2005483
Stojanovska, M., Petruševski, V. M., & Šoptrajanov, B. (2012b). The
concept of sublimation—Iodine as an example. Educatión Quimica,
23, 171.
Taber, K. (2001). Building the structural concepts of chemistry: Some
considerations from educational research. Chemistry Education Re-
search and Practice in Europe, 2, 123-158. doi:10.1039/b1rp90014e
Taber, K. S. (2011). Models, molecules and misconceptions: A com-
mentary on “Secondary school students’ misconceptions of covalent
bonding”. Journ a l of Turkish Science E d u c a t i o n , 8, 3-18.
Tasker, R. (1998). The VisChem project: Molecular level animations in
chemistry—Potential and caution. UniServe Science News, 9, 12-16.
Tatar, E. (2011). Prospective primary school teachers’ misconceptions
about states of matter. Educational Research and Reviews, 6, 197-
200.
The Nobel Prize in Physiology or Medicine (1998). URL (last checked
Copyright © 2012 SciRes. 629
M. I. STOJANOVSKA ET AL.
Copyright © 2012 SciRes.
630
2 July 2012).
http://www.nobelprize.org/nobel_prizes/medicine/laureates/1998/
Treagust, D. F., Chandrasesaran, A. L., Zain, A. N. M., Ong, E. T.,
Karpudewan, M., & Halim, L. (2011). Evaluation of an intervention
instructional program to facilitate understanding of basic particle
concepts among students enrolled in several levels of study. Chemis-
try Education Research and Practic e, 12, 251-261.
doi:10.1039/c1rp90030g
Wenning, C. J. (2008). Dealing more effectively with alternative con-
ceptions in science. Journal of Physics Teacher Education Online, 5,
11-19.
Yezierski, E. J., & Birk, J. P. (2006). Misconceptions about the par-
ticulate nature of matter. Using animations to close the gender gap.
Journal of Chemical Ed uc a t io n , 83, 954-960.
doi:10.1021/ed083p954
M. I. STOJANOVSKA ET AL.
Appendix 1: The Multiple-Choice Instrument
1) Which of the following statements regarding the bromi-
nation of butadiene represented by the equation
CH2=CH-CH=CH2 + Br2 CH2Br-CHBr-CH=CH2
correctly describes what happens?
a) butadiene reacts with two bromines.
b) butadiene reacts with one molecule of bromine.
c) one mole butadiene reacts with one mole bromine.
d) one gram butadiene reacts with one gram bromine.
e) ______________________________________________
2) Copper(II) sulfate from aqueous solution crystallizes:
a) with five molecules of water.
b) as a solid substance composed of copper(II) sulfate
molecules.
c) as а pentahydrate.
d) as an anhydrous salt.
e) ______________________________________________
3) In the process of ice melting, the volume of the system
(mixture of ice and liquid water):
a) decreases.
b) increases.
c) remains the same.
4) The reason for your answer to the third question is:
a) ice particles are smaller than the water particles.
b) ice particles are bigger than the water particles.
c) particles change from solid to liquid ones.
d) the density of the system increases.
e) the density of the system decreases.
f) particles receive heat and enlarge.
g) the change in temperature changes the size of the parti-
cles.
h) ______________________________________________
5) What happens with the carbon, hydrogen and oxygen at-
oms, of which ethanol is composed, when it is ignited?
a) they will ignite too and will burn just like the ethanol.
b) they will disappear.
c) they will become part of some other substance(s), differ-
ent from ethanol.
d) ______________________________________________
6) What happens with the particles of a substance during
heating that substance?
a) they enlarge.
b) they shrink.
c) their size does not change.
d) ______________________________________________
Appendix 2: Interview Protocol
Phase 1 (Briefing phase, not recorded)
Information about the purpose of the research and explana-
tions about the procedure of the interviewing process.
Ethical considerations (permission for audio-tape record-
ings, confidentiality of the obtained information etc.).
Introductory questions, such as: “How do you picture atoms
and molecules? What can you say about their appearance?”,
“Name several substances and write down their symbolic
representations”. “Are these substances composed of mole-
cules?”
Phase 2 (Main phase, recorded)
Discussion related to M1 and M2:
Write down one example of a chemical reaction. What do
these symbolic representations stand for? What are the
qualitative and quantitative information that can be drawn
out?
Discussion related to M3:
Write down an example of a crystalline hydrate. What does
this symbolic representation stand for? What are the quali-
tative and quantitative information that can be drawn out?
The discussion continues with students’ views of ionic and
covalent substances, their distinction and their building par-
ticles.
Discussion related to M4 and M5:
Have you ever put a plastic water-filled bottle in a deep-
freeze? What happened? What happens if you put ice cubes
in a glass of water? Why? Further discussions on the vol-
ume and density of ice compared to those of liquid water
(the lakes freezing on the surface, the rock breaking due to
freezing of water inside the gaps etc.).
Discussion related to M6:
What happens with atoms of a certain substance upon
burning? Is burning of alcohol different from the other
types of burning?
Discussion related to M5 and M7:
[Making models of water molecules]. Water evaporates at
any temperature, but this process is more significant upon
heating and boiling. Could you describe what happens with
the water molecules during evaporation?
Water as a substance exists in three states: solid, liquid and
gaseous. Explain what happens with the water molecules
during a phase change. Do you think they are different in
different phases? If you do, explain how. [An assignment
was given to those students who claimed that molecules are
different, to make new models of water molecules in dif-
ferent phases].
Do you think that temperature, pressure or some other exter-
nal factor have an influence on atoms or molecules? If you do,
explain how. Do you have further examples of this?
Gold has a characteristic golden color, copper is brick red
and iodine is dark purple. Having this in mind, is it possible to
draw a conclusion about the color of the atoms (or molecules)
of which these substances consist?
As a conclusion: Do you think that the properties of the
building particles are the same as those of the parent material?
Phase 3 (Debriefing phase, not recorded)
Asking the participants whether they have anything else to
add to the discussion or additional questions to pose.
Copyright © 2012 SciRes. 631