Sei sulla pagina 1di 13

P1: JLS

Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

23:24

Style file version June 20th, 2002

C 2004)
Journal of Science Education and Technology, Vol. 13, No. 2, June 2004 (

Some Student Misconceptions in Chemistry:


A Literature Review of Chemical Bonding
1

Haluk Ozmen

Students misconceptions before or after formal instruction have become a major concern
among researchers in science education because they influence how students learn new scientific knowledge, play an essential role in subsequent learning and become a hindrance in
acquiring the correct body of knowledge. In this paper some students misconceptions on
chemical bonding reported in the literature were investigated and presented. With this aim, a
detailed literature review of chemical bonding was carried out and the collected data was presented from past to day historically. On the basis of the results some suggestions for teaching
were made.
KEY WORDS: chemistry; misconception; chemical bonding.

INTRODUCTION

found after teaching has taken place. This constructivist view is the dominant paradigm of learning in science. According to constructivist theory of learning,
knowledge is uniquely constructed by each individual
learner and learners actively construct knowledge to
make sense of the world, interpreting new information in terms of existing cognitive structures (Taber
and Watts, 1997). The particular knowledge that is
constructed by an individual will be affected by the
learners prior knowledge and experience and the social context in which learning takes place (Grayson
et al., 2001; Von Glasersfeld, 1992).
Students preexisting beliefs influence how students learn new scientific knowledge and play an essential role in subsequent learning (Arnaudin and
Mintez, 1985; Boujaoude, 1991; Driver and Oldham,
1986; Shuell, 1987; Tsai, 1996). Hunt and Minstrell
(1997) stated that childrens difficulties in science occur because students conceptions before teaching are
not taken into account and therefore communication barriers between teachers and learners can not
be overcome. These ideas are logical, sensible, and
valuable from the students point of view, strongly
held by the students, but may be significantly different from accepted scientific viewpoints and may not
be in conformity with the true or the scientific explanation (Osborne, 1982; Schoon and Boone, 1998).

Learning science is a cumulative process and


each new piece of information is added to what students already know about the topic at hand. Research
has shown that children bring to lessons a lot of preexisting (alternative) conceptions about scientific phenomena that can interfere with students learning of
correct scientific principles or concepts (Driver and
Easley, 1978; Driver and Erickson, 1983; Fleer, 1999;
Palmer, 1999, 2001; Posner et al., 1982; Taber, 2000).
This understanding has caused science educators to
be increasingly concerned about revealing students
difficulties prior to, during, or after the instruction
in conceptualizing scientific knowledge and suggesting ways of remediation. Alternative conceptions may
arise as a result of the variety of contacts students
make with the physical and social world or as a result of personal experience, interaction with teachers,
other people, or through the media (Gilbert et al.,
1982; Gilbert and Zylberstajn, 1985; Griffiths and
Preston, 1992). These ideas may be present before
any teaching of a topic commences, and are often also
1 Department

of Science Education, Fatih Faculty of Education,


Karadeniz Technical University, 61335 Sogutlu-Trabzon, Turkey;
e-mail: hozmen61@hotmail.com and hozmen@ktu.edu.tr

147
C 2004 Plenum Publishing Corporation
1059-0145/04/0600-0147/0

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

148
And also, it is found that these beliefs are widely
held by learners in various grade levels, they are fairly
pervasive, stable, and resistant to change by conventional teaching strategies and are often held intact by
children and adults alike even after the completion
of years of formal science instruction (Champagne
et al., 1982; Clement, 1982; Guzzetti, 2000; Halloun
and Hestenes, 1985a; Hewson and Hewson, 1984;
Osborne and Cosgrove, 1983; Osborne and Wittrock,
1983; Stavy, 1991; Tsai, 1998; Wandersee et al., 1994).
According to Niaz (2001a), students preconceptions
that resist change can be considered as part of students hard-core beliefs. Students conceptions that
are different from those accepted by the scientific
community are variously labeled in the science education literature as misconceptions (Abimbola, 1988;
Brown, 1992; Chambers and Andre, 1997; Din, 1998;
Driver and Easley, 1978; Gonzalez, 1997; Griffiths,
1994; Griffiths et al., 1988; Griffiths and Preston,
1992; Helm, 1980; Hewson and Hewson, 1984; Lawson
and Thompson, 1988; Michael, 2002; Nakhleh, 1992;
Nussbaum, 1981; Schmidt, 1997; Treagust, 1988),
alternative conceptions (Astudillo and Niaz, 1996;
Driver and Easley, 1978; Gilbert and Swift, 1985;
Niaz, 2001a; Palmer, 2001; Taber, 2001; Wandersee
et al., 1994), preconceptions (Hashweh, 1988; Novak,
1977), alternative frameworks (Driver, 1981; Driver
and Easley, 1978; Gonzalez, 1997; Kuiper, 1994;
Taber, 1999, 2001), nave beliefs (Caramazza et al.,
1981), nave theories (Resnik, 1983), nave conceptions
(Champagne et al., 1983), childrens scientific intuitions (Sutton, 1980), conceptual frameworks
(Southerland et al., 2001), childrens science (Gilbert
et al., 1982; Osborne et al., 1983), common sense
understanding (Hills, 1983), common sense concepts
(Halloun and Hestenes, 1985b), alternative conceptual
framework (Taber, 1998), intuitive conceptions (Lee
and Law, 2001), intuitive science (Preece, 1984), common alternative science conceptions (Gonzalez, 1997),
students intuitive theories (Boujaoude, 1992), prescientific conceptions (Good, 1991), alternate perceptions
(Carter and Brickhouse, 1989), students descriptive
and explanatory systems (Champagne et al., 1982), and
spontaneous knowledge (Pines and West, 1986). In the
science education context, these terms refer to ideas
that students have about natural phenomena that are
inconsistent with scientific conceptions and reflect the
complex nature and multiple causes of childrens erroneous conceptions as viewed by science educators.
Although Wandersee et al. (1994) presented an analysis of the subtle distinctions in the usage of these
terms, no consensus has been reached on the term of

23:24

Style file version June 20th, 2002

Ozmen
choice. For simplicity, the term of misconception will
be used in this paper and it means any concept that
differs from the commonly accepted scientific understanding of the term.
Of course, chemistry is one of the most important branches of science and has been regarded as
a difficult subject for young students by chemistry
teachers, researchers, and educators. Although the
reasons for this vary from the abstract nature of
many chemical concepts to the difficulty of the language of chemistry (Ayas and Demirbas, 1997), there
are two major reasons for students having difficulties in these areas; firstly, the topics are very abstract
(Ben-Zvi et al., 1988), and secondly, words from everyday language are used but with different meanings (Bergquist and Heikkinen, 1990). Because students misconceptions in school sciences at all levels
constitute a major problem of concern to science educators, scientistresearchers, teachers, and students
(Johnstone and Kellett, 1980; Nussbaum, 1981), the
identification of the students understandings and misconceptions have been the goal of many of the studies
carried out over the last years. Some of the conceptual
areas in which most studies have been conducted are
element, compound, and mixture (Ayas and Demirbas,
1997; Papageorgiou and Sakka, 2000), chemical re
actions (Andersson, 1990; Ayas and Ozmen,
2002;
Ben-Zvi et al., 1987; Boo and Watson, 2001; Hesse

and Anderson, 1992; Ozmen


and Ayas, 2003), chemical bonding (Birk and Kurtz, 1999; Boo, 1998; Coll
and Taylor, 2001, 2002; Coll and Treagust, 2001,
2002, 2003; Harrison and Treagust, 2000; Niaz, 2001b;
Nicoll, 2001; Peterson et al., 1986, 1989; Robinson,
1998; Taber, 1994; Tan and Treagust, 1999), chemical equilibrium (Banerjee and Power, 1991; Bergquist
and Heikkinen, 1990; Chiu et al., 2002; Gorodetsky
and Gussarsky, 1986; Gussarsky and Gorodetsky,
1988, 1990; Hackling and Garnett, 1985; Hameed
et al., 1993; Huddle and Pillay, 1996; Maskill and
Cachapuz, 1989; Niaz, 1995, 1998, 2001a; Pedrosa and
Dias, 2000; Quilez-Pardo and Solaz-Portoles, 1995;
Tsaparlis et al., 1998; Tyson et al., 1999; Van Driel,
2002; Voska and Heikkinen, 2000; Wheeler and Kass,
1978), atoms and molecules (Ben-Zvi et al., 1986;
Griffiths and Preston, 1992; Harrison and Treagust,
2000; Lee et al., 1993; Nakhleh and Samarapungavan,
1999; Skamp, 1999; Tsaparlis, 1997), acids and bases
(Bradley and Mosimege, 1998; Hand and Treagust,
1991; Nakhleh and Krajcik, 1994; Sisovic and Bojovic,
2000), mole concept (Furio et al., 2000; Gorin, 1994;
Nelson, 1991; Schmidt, 1994), solubility and solutions
(Ebenezer and Erickson, 1996; Ebenezer and Fraser,

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

23:24

Style file version June 20th, 2002

Misconceptions in Chemical Bonding


2001; Smith and Metz, 1996), evaporation and condensation (Bar and Gaglili, 1994; Chang, 1999; Tytler,
2000), and the particulate nature of matter (Abraham
et al., 1992; De Vos and Verdonk, 1996; Nakhleh and

Samarapungavan, 1999; Ozmen


et al., 2002; Valanides,
2000).
As mentioned above, there are some topics that
chemistry students find more difficult to understand.
One active area of research on chemistry misconceptions is the topic of chemical bonding. This paper aims
to synthesize students misconceptions found in different studies at all levels.

Misconceptions About Chemical Bonding


Chemical bonding is one of the most important
topics in undergraduate chemistry and the topic
involves the use of a variety of models varying from
simple analogical models to sophisticated abstract
models possessing considerable mathematical complexity (Coll and Taylor, 2002; Coll and Treagust,
2003; Fensham, 1975). It is also a topic that students
commonly find problematic and develop a wide range
of misconceptions. The concepts of electron, ionization energy, electronegativity, bonding, geometry,
molecular structure, and stability are central to much
of chemistry, from reactivity in organic chemistry to
spectroscopy in analytical chemistry (Nicoll, 2001).
And also, it is important for students to grasp these
concepts in understanding why and how chemical
bonds occur. Chemical bonding has been classified

149
into a series of three target systems; metallic, ionic,
and covalent bonding. In the science education literature, there have been numerous studies to determine
students understanding and misconceptions about
metallic, ionic, and covalent bonding. These studies
have revealed prevalent and consistent misconceptions across a range of ages and cultural settings.
Butts and Smith (1987) reported that students
were confused about covalent and ionic bonds. Some
of the students they studied conceptualized the
sodium and chlorine atoms as being held together by
covalent bonds.
Peterson et al. (1989) investigated Grade-11
and Grade-12 students misconceptions of covalent
bonding and structure. They found that these students did not acquire a satisfactory understanding of
covalent bonding. Specifically, 33% of Grade-11 and
23% of Grade-12 held misconceptions regarding the
unequal sharing and position of an electron pair in a
covalent bond. These students seem to relate electron
sharing to covalent bonding, yet did not consider
the influence of electronegativity and the resultant
unequal electron sharing. As a result of the analysis
of the students responses, some misconceptions were
identified. These misconceptions were discussed under the categories of bond polarity, molecular shape,
polarity of molecules, intermolecular forces, the octet
rule, and lattices. The misconceptions identified are
depicted in Table I.
In another study, Goh et al. (1993) have investigated students misconceptions including chemical bonding in chemistry and revealed that students

Table I. The Most Common Misconceptions of Covalent Bonding and Structure Held by Grade-11 and Grade-12 Students
Bond polarity
Equal sharing of the electron pair occurs in all covalent bonds.
The polarity of a bond is dependent on the number of valence electrons in each atom involved in the bond.
Ionic charge determines the polarity of the bond.
Molecular shape
The shape of a molecule is due to the repulsion between the bonds.
The V-shape in a molecule is due to the repulsion between the nonbonding electron pairs.
Bond polarity determines the shape of a molecule.
Polarity of molecules
Nonpolar molecules form when the atoms in the molecule have similar electronegativities.
Molecules of the type OF2 are polar as the nonbonding electrons on the oxygen form a partial negative charge.
Intermolecular forces
Intermolecular forces are the forces within a molecule.
Strong intermolecular forces exist in a continuous covalent solid.
Covalent bonds are broken when a substance changes shape.
Octet rule
Nitrogen atoms can share five electron pairs in bonding.
Lattices
High viscosity of some molecular solids is due to strong bonds in the continuous covalent lattice.

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

23:24

Style file version June 20th, 2002

Ozmen

150
believed intermolecular bonding was stronger than
intramolecular bonding. This was consistent with
Peterson et al.s (1989) findings.
One case study conducted by Taber (1995) has investigated students understanding of some very basic
bonding concept and found misconceptions dealing
with covalent bonding, metallic bonding, resonance
structure, coordinate bonding, hydrogen bonding, and
van der Walls forces. For example, it is claimed that
learners invoke intramolecular bonding in ionic compounds. And also, it is stated that there is some
evidence that learners appreciated the relationship
between intermolecular bonding and physical properties such as boiling point. These results are consistent
with Peterson and Treagusts (1989), Peterson et al.s
(1989), and Tabers (1998) findings.
Taber (1997) has also investigated students
misconceptions dealing with ionic bonding. In the
study, a small-scale survey was used to investigate
how widespread were misconceptions of the ionic
bond and he established that students had difficulty
understanding ionic bonding. He stated that many
chemistry students understanding of ionic bonding:
(i) overemphasizes the process of electron transfer,
(ii) explicitly uses the notion of ion-pairs as molecules,
(iii) is constrained by an appropriate consideration
of valency, (iv) pays heed to an irrelevant electron
history, (v) distinguishes between what are actually
equivalent interactions between ions.
A later misconception, reported by Boo (1998),
is that some students believe that a chemical bond is
a physical entity. Boo suggests that this means that
students believed that bond breaking releases energy
and bond making involves energy input.
Robinson (1998) has outlined some of the general misconceptions related to chemical bonding.
These misconceptions are listed in Table II.
Birk and Kurtz (1999) designed a study to
diagnose student misconception over a large range of

chemical experience from high school to faculty and


to determine if and when the misconceptions disappear. The test developed to collect data consisted of
questions that examine understanding in six areas:
bond polarity, molecular shape, polarity of molecules,
lattices, intermolecular forces, and the octet rule. The
most important misconception identified is that students believe that it is absent in polar molecular substances such as water. Students misconceptions identified from the exam papers are depicted in Table III.
Barker (2000) has investigated students understanding of chemical bonding and thermodynamics.
She found that although basic ideas about covalent
and hydrogen bonding appear to be learned by a
majority of students, ions and ionic bonding continue
to cause difficulties. Some students seem to imagine
ionic compounds exist as discrete molecules like as
covalent compounds and therefore think of ionic
bonds as unidirectional and subject to the same rules
of behavior as covalent bonds. Students think that
covalent bonds are weak compared to ionic bonds
and so break more easily. In addition, students reason
that hydrogen chloride exists as discrete molecules
in acid solution and when metal is added a bond is
being formed between the metal and the chlorine
atom, swapping patterns with the hydrogen.
In a study reported by Nicoll (2001), it was
described the types of misconceptions related to electronegativity, bonding, geometry, and microscopic
representations that undergraduate chemistry students hold. According to results, while students may
have appeared to know about the concept of polarity,
they did not associate it at all with electronegativity.
For example, when a junior student, Janet, was asked
to define polarity, she stated, Polarity is like a polar
substance is something thats neither ionic nor it is
covalent. In another misconception on bonding, it
was seen that several students appeared to confuse
the definitions of ionic and covalent bonding. For

Table II. The General Misconceptions Related to Chemical Bonding


Chemical bonds form in order to produce filled shells rather than filled shells being the consequence of the formation
of many covalent bonds.
Atoms need filled shells.
A covalent bond holds atoms together because the bond is sharing electrons.
Molecules form from isolated atoms.
There are only two kinds of bonds: covalent bonds and ionic bonds. Anything else is just a force, not a proper bond.
Ionic bonds are the transfer of electrons, rather than the attractions of the ions that result from the transfer of
electrons. The reason electrons are transferred is to achieve a full shell.
An ionic bond only occurs between the atoms involved in the electron transfer. Thus, sodium ion forms one ionic bond
to a chloride ion in solid sodium chloride and is involved in five forces with the other adjacent chloride ions.
Na+ and other ions are stable because they have a filled outer shell.

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

23:24

Style file version June 20th, 2002

Misconceptions in Chemical Bonding

151

Table III. Students Misconceptions Identified From the Exam Papers


Molecular shape
The shape of the molecules is due only to the repulsion between bonding pairs.
The shape of the molecules is due only to the repulsion between nonbonding electron pairs.
Bond polarity determines the shape of a molecule.
Bond polarity
Equal sharing of the electron pair occurs in all covalent bonds.
The polarity of the bond is dependent on the number of valence electrons in each atom involved in the bond.
Ionic charge determines the polarity of the bond.
Nonbonding electron pairs influence the position of the shared pair and determine the polarity of the bond
The largest atom exerts the greatest control over the shared electron pair.
Electrons have a positive charge.
Polarity of molecules
Nonpolar molecules form only when atoms in the molecule have similar electronegativities.
Molecules of the type OF2 are polar as the nonbonding electrons on the oxygen form a partial negative charge.
A molecule is polar because it has polar bonds.

example, when he was asked to explain what covalent


bonding was, a junior student, Duane, stated that I
just think of it as attractions between the negative and
positive ends of an atom. During the interview, students were asked to explain why molecules adopted
the geometries that they did. Students would mention
incorrect reasons. For example, when he was asked
to explain why water adopts a bent geometry, a freshman student, Bridgette, stated that Its because the
two lone pairs of electrons have higher energy levels
or they are like stronger. They want more space and
so they push the bonded pairs down because bonds
are less energy, they are happy and they do not need
that much space. When Casey, a senior student, was
asked to pretend that she could see one molecule of
water and to describe what see would see, she replied,
If you saw the electrons, they would be touching.
In a study conducted by Coll and Taylor (2001),
it was aimed to determination of misconceptions of

chemical bonding held by upper secondary and tertiary students. At the end of the study, some 20 misconceptions were revealed, the most common being
belief that continuous ionic and metallic lattices were
molecular in nature, and confusion over ionic size
and charge. Students misconceptions identified in the
study are depicted in Table IV.
Coll and Treagust (2001) investigated year-12 undergraduate and postgraduate Australian students
mental models for chemical bonding using semistructured interviews comprising a three-phase interview
protocol. In the study, students were presented with
samples of metallic, ionic, and covalent substances,
and asked to describe the bonding in the substance.
Students responses revealed that students use simple,
realistic mental models for chemical bonding. In contrast, other studies reveal that learners mental models
of bonding become sophisticated and complex models they were exposed to during instruction (Coll and

Table IV. Students Misconceptions for Chemical Bonding


Metallic bonding is weak bonding.
Intramolecular covalent bonding is weak
bonding.
Ionic bonding is weak bonding.
Continuous metallic or ionic lattices are
molecular in nature.
The bonding in metals and ionic compounds
involves intermolecular bonding.
The ionic radius of the sodium ion is greater than
the chloride ion.
The ionic radius of the lithium ion is greater than
the sodium ion.
Polar covalent compounds contain charged
species.
Molecular iodine contains 1 minus ions.
The charged species in metallic lattices are nuclei
rather than ions.

Metallic lattices contain neutral atoms.


Electronegativity comprises attraction for a
single electron.
Molecular iodine is metallic in nature.
Ionic bonding comprises sharing of electrons.
Ionic and metallic bondings contain an element
of directionality.
Ions in close-packed metal lattices possess other
than eight nearest neighbors.
Metal to nonmetal bonding in alloys is
electrostatic in nature.
Ionic shape and packing is influenced by pressure.
Intermolecular forces are influences by gravity.
Glass is an ionic crystalline substance.

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

152
Taylor, 2002; Coll and Treagust, 2002, 2003). And also,
they struggle to use their mental models to explain the
physical properties of covalently bonded substances.

CONCLUSION AND IMPLICATIONS


FOR TEACHING
Bonding is the key to molecular structure,
and structure is intimately related to the physical and chemical properties of a compound. An
understanding of the concept of bonding is fundamental to subsequent learning of various topics in
chemistry, including chemical equilibrium, thermodynamics, molecular structure, and chemical reactions.
Therefore, an understanding of molecular structure
based on atomic structure and bonding is crucial to
subsequent understanding of chemical reactions. In a
chemical reaction, there is a change in the bonding
of the atoms, from the bonding in the reactants to the
bonding in the products. Since the concepts of molecular structure and chemical bonding are built upon the
fundamental principles of atomic structure, this understanding of chemical behavior at the atomic level
appears important in understanding subsequent concepts in chemistry. But, although the students at each
level have begun to learn this concept from earlier
stages of their schooling, as mentioned above, there
are a lot of studies reported that students have some
difficulties in understanding chemical bonding and
hold several misconceptions about it. These misconceptions appear to be resistant to attempts to change
them over time, despite increased chemistry education. Students pass from grade to grade without fully
grasping the underlying concepts of bonding.
There may be a lot of reasons in generating misconceptions. In the classroom teaching, teachers generally use ball and stick models to represent chemical
bonds. According to Butts and Smith (1987), the ubiquitous use of ball and stick models used to model ionic
lattices may be instrumental in the generation of this
misconception because learners mistake sticks for individual chemical bonds.
A considerable amount of research has pointed
out that the process of knowledge construction involves the replacement or reorganization of the conceptual framework. But for several concepts, such as
chemical bonding, chemical equilibrium, acids and
bases, students have difficulty changing their initial
perceptions of the concepts. Especially, abstract concepts encountered in the study of chemistry provide
increased opportunity for the development of formal

23:24

Style file version June 20th, 2002

Ozmen
misconceptions. Although students at each level take
several science classes during their schooling in order to learn various science concepts including chemical bonding, the presence of misconceptions in their
explanation indicates their fragmented understanding of these abstract concepts. Sometimes students
have such strong misconceptions that even after learning the correct concepts in the classrooms, they resist
modifying their preexisting ideas. Instead, they try to
interpret the new acquired knowledge using their preconceptions (Khalid, 2003).
It is obvious that why misconceptions exist is an
important question in science education and in other
disciplines. Although incorrect, imprecise, or incomplete teaching may play an important role, according
to Tsaparlis (1997), there must be a more fundamental cause that results in one or more of the following:
i) the inability of most or many students to employ formal operations, ii) the lack of the proper knowledge
corpus which is a prerequisite for meaningful learning, iii) the absence of the relevant concepts from
long term memory. If someone thinks what can be
done to improve student understanding of the basic
chemistry concepts and to remediate their misconceptions, a starting point may be to remove some of the
content from the first-year course and spend more
time for fundamental concepts before moving onto
more abstract ones, because it is also well-known that
school curricula are very intensive. For this reason,
some reform may be necessary in the chemistry curriculum at all levels to facilitate students conceptual
understanding of bonding topics. Driver and Oldham
(1986) suggested a reduction in content at all levels of
education in order to allow children time to construct
concepts for themselves. And also, Nicoll (2001) suggests that teachers need to emphasize the transitions
between the symbolic, macroscopic, and microscopic
worlds so that students will develop their own mental
models of bonding on these three levels.
Misconceptions arise not only from students
contacts with the physical and social world and from
textbooks (Cho et al., 1985), but also as a result of
interaction with teachers (Gilbert and Zylberstajn,
1985). Teachers should also discuss the abstract
concepts in their classrooms in order to eliminate
students misconceptions regarding these concepts.
When the teachers were less knowledgeable, they
were more likely to rely upon low-level questions
and to give their students less opportunities to
speak (Valanides, 2000). According to Bergquist and
Heikkinen (1990), it is critical to provide students
with opportunity to verbalize their ideas to promote

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

Misconceptions in Chemical Bonding


concept building and remediate misconceptions. Only
then will deep-seated misunderstandings be identified, diagnosed, and addressed.
In addition, researchers indicate that students
difficulties and misconceptions in learning science
concepts are due in part to the teachers lack of knowledge regarding students prior understanding and
knowledge of concepts under study (Krishnan and
Howe, 1994). The identification of possible sources of
misconceptions is important because the instructional
strategies which ultimately might prove effective in
combating misconceptions might differ according to
the type or source of misconception. One of the most
fruitful outcomes of the studies on childrens misconceptions is to alert teachers to students difficulties
in conceptualizing science knowledge and hence suggest more effective strategies for improving classroom
instruction. Before teaching a concept, such as chemical bonding, redox, chemical equilibrium, acids and
bases, teachers should be able to check the literature
to find out what is known about misconceptions that
students may bring to class and which teaching methods are the best in correcting these misconceptions.
Such an approach would provide to teachers a chance
to design better learning environments that help to
develop concepts scientifically. But, unfortunately, in
practice many chemistry teachers continue to teach
their subjects as if none of these researches were undertaken and, as a result of this, there becomes a
gap between research and teaching, and students pass
from grade to grade without fully grasping these concepts and having extra misconceptions. The constructivist literature emphasizes that the teacher always
has to teach from where the students are rather than
where the teacher would like them to be, or where
the curriculum suggests they should be (Taber, 2001).
It is therefore recommended that at the start of the
teaching sequence, students ideas need to be made
explicit to teacher and students (Driver and Oldham,
1986). The key problem here is that teachers expect
research to be presented to them in a form they can
readily apply because they are too busy doing their
job to read the research literature (De Jong, 2000).
For this reason, to explore and use research findings to
improve chemistry learning, it is important to develop
diagnostic instruments as well as improving curricular
resources and teaching approaches.
In the literature, there are several techniques
and instruments, such as concept mapping, interview
about instances and events, interview about concepts,
predictionobservationexplanation, drawings, word
association, pencil-and-paper diagnostic instruments

23:24

Style file version June 20th, 2002

153
based on multiple-choice items, two-tier multiplechoice tests (Peterson et al., 1989; Schmidt, 1997;
White and Gunstone, 1992), that can be used by teachers in their classroom environment in identifying misconceptions of science phenomena. Of these many
approaches, interviews, and multiple-choice diagnostic tests are most common methodologies and have acquired strong support as a viable approach (Osborne
and Gilbert, 1980; Peterson et al., 1989). But according
to Treagust (1988), conventional multiple-choice tests
do not adequately assess student understanding. Although multiple-choice tests have been used to evaluate students content knowledge, they have some limitations with determining students reasoning behind
their choices. However, many instructors agree that
one of the best ways to measure student understanding is to assess how well they can explain a concept to
someone else (Teichert and Stacy, 2002). Therefore,
multiple-choice questions can be validated by asking
students to give reasons for their answers. In addition,
two-tier multiple-choice items to question based on
student reasoning, including known misconceptions,
appear to provide a feasible approach for evaluating students understanding, and for identifying commonly held misconceptions (Peterson and Treagust,
1989). The items in two-tier multiple-choice diagnostic instruments are specifically designed to identify
students misconceptions and misunderstandings in a
limited content area. The first part of each item consists of a multiple-choice content question having two
or three choices. The second part of each item contains
a set of four or five possible reasons for the answer
to the first part. Incorrect reasons are derived from
actual students misconceptions gathered form literature, interviews, and free response tests (Tan et al.,
2002). In addition, this type of test is more readily administered and scored than the other methods, and
are useful for classroom teachers (Tan and Treagust,
1999). But on the other hand, objectively scored twotier tests also have disadvantage of detecting far fewer
conceptions than students may actually possess within
a content domain. By contrast, open-ended two-tier
tests allow teachers to explore each students reasoning patterns and supporting conceptions (Voska and
Heikkinen, 2000). In the literature, although there are
a few diagnostic instruments that teachers can use in
the classroom regarding chemical bonding (Birk and
Kurtz, 1999; Goh et al., 1993; Peterson et al., 1989;
Peterson and Treagust, 1989; Tan and Treagust, 1999;
Treagust, 1988), most reported strategies involve a
combination of multiple-choice tests, interviews, or
other tasks. Simple and objectively scored diagnostic

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

154
assessment tests that can be used in the classrooms
should be also developed by teachers to determine
the level of students understanding and misconceptions. And also, teachers should be informed about
determining and alleviating of misconceptions and
using appropriate teaching strategies with in-service
training courses. In another words, teachers should be
equipped with the necessary capabilities of identifying their own students conceptions and implementing teaching approaches that promote conceptual understanding among their students. In parallel to this,
teacher education department of universities should
give special attention in this regard.
Training should help students to relate new information to prior knowledge, to integrate information
for one subject area into another, and to relate classroom information to everyday experiences to help
those students become meaningful learners who are
better able to retain and use information in novel situations (Prawat, 1989). A majority of teachers and even
professors use teacher-centered teaching strategies to
teach science (Lord, 1999; Yip, 2001). To be successful in examinations, pupils are trained to be good at
retrieving factual information and the rote application of algorithms. These traditional teaching strategies provide conceptual information to the students
who learn the material, memorize it, and reproduce
it on the day of examination (Khalid, 2003). It is well
known that traditional teaching strategies are ineffective to help students with a complete understanding
of the abstract concepts such as chemical bonding,
chemical equilibrium, the mole concept, chemical
kinetics, acids and bases, atoms and molecules, to
build correct conceptions, to alleviate misconceptions,
and to promote conceptual change (Westbrook and
Marek, 1991). As students learn more about chemistry their cognitive structure is expected to develop
in at least three ways: the range of their concepts will
increase, the level of sophistication of their concepts
will deepen, and their concepts will become better
integrated with each other (Taber and Watts, 1997).
Therefore, teaching methods used in classrooms by
teachers should support these expectations.
According to constructivist view of learning,
meaningful learning occurs when the learners actively construct their own knowledge by using existing knowledge to make sense of newly gained experiences. Taber (2000) has stated that the first step
in a constructivist learning approach is to make the
teacher and student aware of the learners current
ideas. Teaching can then be planned that challenges
misconceptions, and provides students with the op-

23:24

Style file version June 20th, 2002

Ozmen
portunities and rationale for conceptual restructuring. In this situation, teachers can play an important
role in teaching chemistry concepts. Teachers can help
students eliminate their misconceptions by providing
an adequate knowledge base and clear understanding of these concepts. This view highlights the impact
of learners preconceptions and misconceptions on
the process of developing new knowledge. Because
misconceptions affect subsequent learning negatively
(Bodner, 1986), the correction or remediation of students misconceptions is as important as identification
of them. In the literature, there are several methods
used in remediation of the misconceptions. Among
these, conceptual change approach has a large usage
area in science education (Posner et al., 1982; Sanger,
2000). If a concepts meaning has been completely removed and replaced by something else that is incomparable to the existing meaning, it would be considered a conceptual change (Chiu et al., 2002). Within
this perspective, learning is depicted as a process of
conceptual change. This approach represents an alternative approach designed to encourage students
to alter misconceptions. This approach suggests that
four conditions must exist before a conceptual change
is likely to occur (Chambers and Andre, 1997; Posner
et al., 1982): (i) students must become dissatisfied with
their existing conceptions; students must have experiences which lead them to lose faith in the ability
of their current conceptions to solve problems, (ii)
the new conceptions must be intelligible; the student
must be able to understand sufficiently how experiences can be structured by the new concept, (iii) the
new conception must appear plausible; any new concept adopted must at least appear to have the ability to
solve the problems generated by its predecessors, (iv)
the new conception must be fruitful; it should have
the capacity to open up new areas of inquiry.
On the basis of this model, many specific instructional strategies have been proposed to help students
change their misconceptions. Among these, refutational texts and conceptual change texts have become popular for the last two decades. As stated by
Chambers and Andre (1997), the major difference
between the refutational text model and the conceptual change text involves whether students are asked
to make a prediction about a situation. In the conceptual change model, students are asked to predict
what would happen in a situation before being presented with information that demonstrates the inconsistency between common misconceptions and the
scientific conception. In the refutational text model,
common misconceptions are contrasted to scientific

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

Misconceptions in Chemical Bonding


conceptions, but the students is not asked first to
make a prediction about a common situation before
the refutation is given. Although these strategies are
well-known and most useful strategies, very few science teachers are aware of conceptual-change teaching techniques at present time (Hesse and Anderson,
1992). That the teachers should be informed about using of these strategies may be very useful for them to
help students change their misconceptions.
Among many instructional materials, textbooks
are most important information sources for students.
Many research studies have found that the textbooks
used in schools have inadequate or sometimes incorrect information (Soyibo, 1995). Therefore, textbooks authors should help teachers become aware
of the common misconceptions students bring to the
chemistry classroom. And also, taking into account
the apparent preeminence of textbooks in shaping
curricula, and the contribution of instruction for dissemination of misconceptions on important chemistry
topics (Haidar, 1997), it is sensible and adequate to initiate a coherent programme beginning with textbooks
followed by programmes with chemistry teachers
(Pedrosa and Dias, 2000). In parallel to textbooks,
guide materials and new teaching materials that may
help to remedy students misconceptions should be
devised and presented to teachers usage.

ACKNOWLEDGMENTS
I am very grateful to Dr Mansoor Niaz for his
proofreading, correction, and comments.

REFERENCES
Abimbola, I. O. (1988). The problem of terminology in the study
of students conceptions in science. Science Education 72: 175
184.
Abraham, M. R., Grzybowski, E. B., Renner, J. W., and Marek,
E. A. (1992). Understandings and misunderstandings of eighth
graders of five chemistry concepts found in textbooks. Journal
of Research in Science Teaching 29: 105120.
Andersson, B. (1990). Pupils conceptions of matter and its transformations (age 1216). Studies in Science Education 18: 5385.
Arnaudin, M. W., and Mintez, J. J. (1985). Students alternative conceptions of the human circulatory system: A cross-age study.
Science Education 69: 721733.
Astudillo, L. R., and Niaz, M. (1996). Reasoning strategies used
by students to solve stoichiometry problems and its relationship to alternative conceptions, prior knowledge, and cognitive
variables. Journal of Science Education and Technology 5: 131
140.
Ayas, A., and Demirbas, A. (1997). Turkish secondary students
conceptions of introductory chemistry concepts. Journal of
Chemical Education 74: 518521.

23:24

Style file version June 20th, 2002

155

Ayas, A., and Ozmen,


H. (2002). Students misconceptions about
chemical reactions at secondary level. Paper presented the
First International Education Conference on Changing Times,
Changing Needs, Eastern Mediterranean University Faculty
of Education, May 810, Gazimagusa, Turkish Republic of
Northern Cyprus.
Banerjee, A. C., and Power, C. N. (1991). The development of modules for the teaching of chemical equilibrium. International
Journal of Science Education 13: 355362.
Bar, V., and Gaglili, I. (1994). Stages of childrens views about evaporation. International Journal of Science Education 16: 157
174.
Barker, V. (2000). Students reasoning about basic chemical thermodynamics and chemical bonding: What changes occur during a context-based post-16 chemistry course? International
Journal of Science Education 22: 11711200.
Ben-Zvi, R., Eylon, B., and Silberstein, J. (1986). Is an atom of
copper malleable? Journal of Chemical Education 63: 64
66.
Ben-Zvi, R., Eylon, B., and Silberstein, J. (1987). Students visualization of a chemical reaction. Education in Chemistry 24:
117120.
Ben-Zvi, R., Eylon, B., and Silberstein, J. (1988). Theories, principles and laws. Education in Chemistry 25: 8992.
Bergquist, W., and Heikkinen, H. (1990). Student ideas regarding
chemical equilibrium. Journal of Chemical Education 67: 1000
1003.
Birk, J. P., and Kurtz, M. J. (1999). Effect of experience on retention
and elimination of misconceptions about molecular structure
and bonding. Journal of Chemical Education 76: 124128.
Bodner, G. (1986). Constructivism: A theory of knowledge. Journal
of Chemical Education 63: 873878.
Boo, H. K. (1998). Students understanding of chemical bonds and
the energetic of chemical reactions. Journal of Research in Science Teaching 35: 569581.
Boo, H. K., and Watson, J. R. (2001). Progression in high school
students (aged 1618) conceptualizations about chemical reactions in solution. Science Education 85: 568585.
Boujaoude, S. B. (1991). A study of the nature of students understanding about the concept of burning. Journal of Research in
Science Teaching 28: 689704.
Boujaoude, S. B. (1992). The relationship between students learning strategies and the change in their misunderstandings during
a high school chemistry course. Journal of Research in Science
Teaching 29: 687699.
Bradley, J. D., and Mosimege, M. D. (1998). Misconceptions in acids
and bases: A comparative study of student teachers with different chemistry backgrounds. South African Journal of Chemistry 51: 137147.
Brown, D. E. (1992). Using examples and analogies to remediate misconceptions in physics: Factors influencing conceptual change. Journal of Research in Science Teaching 29: 17
34.
Butts, B., and Smith, R. (1987). HSC chemistry students understanding of the structure and properties of molecular and ionic
compounds. Research in Science Education 17: 192201.
Caramazza, A., McCloskey, M., and Green, B. (1981). Naive beliefs
in sophisticated subjects: Misconceptions about trajectories
of objects. Cognition 9: 117123.
Carter, C. S., and Brickhouse, N. W. (1989). What makes chemistry
difficult? Alternate perceptions. Journal of Chemical Education 66: 223225.
Chambers, S. K., and Andre, T. (1997). Gender, prior knowledge,
interest, and experiences in electricity and conceptual change
text manipulations in learning about direct current. Journal of
Research in Science Teaching 34: 107123.
Champagne, A., Gunstone, R., and Klopfer, L. (1983). Naive
knowledge and science learning. Research in Science and Technological Education 1: 173183.

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

156
Champagne, A. B., Klopfer, L. E., and Gunstone, R. F. (1982).
Cognitive research and the design of science instruction. Educational Psychologist 17: 3153.
Chang, J. Y. (1999). Teacher collage students conception about
evaporation, condensation, and boiling. Science Education 83:
511526.
Chiu, M. H., Chou, C. C., and Liu, C. J. (2002). Dynamic processes of
conceptual change: Analysis if constructing mental models of
chemical equilibrium. Journal of Research in Science Teaching
39: 688712.
Cho, H., Kahle, J. B., and Nordland, F. H. (1985). An investigation
of high school biology textbooks as sources of misconceptions
and difficulties in genetics and some suggestions for teaching
genetics. Science Education 69: 707719.
Clement, J. (1982). Students preconceptions in introductory mechanics. American Journal of Physics 50: 6671.
Coll, R. K., and Taylor, N. (2001). Alternative conceptions of chemical bonding held by upper secondary and tertiary students.
Research in Science and Technological Education 19: 171191.
Coll, R. K., and Taylor, N. (2002). Mental models in chemistry:
Senior chemistry students mental models of chemical bonding. Chemistry Education: Research and Practice in Europe 3:
175184.
Coll, R. K., and Treagust, D. F. (2001). Learners mental models of
chemical bonding. Research in Science Education 31: 357382.
Coll, R. K., and Treagust, D. F. (2002). Exploring tertiary students
understanding of covalent bonding. Research in Science and
Technological Education 20: 241267.
Coll, R. K., and Treagust, D. F. (2003). Investigation of secondary
school, undergraduate, and graduate learners mental models
of ionic bonding. Journal of Research in Science Teaching 40:
464486.
De Jong, O. (2000). Crossing the borders: Chemical education research and teaching practice. University Chemistry Education
4: 2932.
De Vos, W., and Verdonk, A. H. (1996). The particulate nature of
matter in science education and in science. Journal of Research
in Science Teaching 33: 657664.
Din, Y. (1998). Childrens misconceptions on reproduction and implications for teaching. Journal of Biological Education 33: 21.
Driver, R. (1981). Pupils alternative frameworks in science. European Journal of Science Education 3: 93101.
Driver, R., and Easley, J. (1978). Pupils and paradigms: A review
of literature related the concept development in adolescent
science students. Studies in Science Education 5: 6184.
Driver, R., and Erickson, G. (1983). Theories-in-action: Some theoretical and empirical issues in the study of students conceptual
frameworks in science. Studies in Science Education 10: 3760.
Driver, R., and Oldham, V. (1986). A constructivist approach to curriculum development in science. Studies in Science Education
13: 105122.
Ebenezer, J. V., and Erickson, L. G. (1996). Chemistry students
conception of solubility: A phenomenography. Science Education 80: 181201.
Ebenezer, J. V., and Fraser, M. D. (2001). First year chemical engineering students conceptions of energy in solution process:
Phenomenographic categories for common knowledge construction. Science Education 85: 509535.
Fensham, P. (1975). Concept formation. In Daniels, D. J. (Ed.), New
Movements in the Study and Teaching of Chemistry, Temple
Smith, London, pp. 199217.
Fleer, M. (1999). Childrens alternative views: Alternative to what?
International Journal of Science Education 21: 119135.
Furio, C., Azcona, R., Guisasola, J., and Ratcliffe, M., (2000). Difficulties in teaching the concept of amount of substance and
mole. International Journal of Science Education 22: 1285
1304.
Garnett, P. J., Garnett, P. J., and Hackling, M. W. (1995). Students
alternative conceptions in chemistry: A review of research and

23:24

Style file version June 20th, 2002

Ozmen
implications for teaching and learning. Studies in Science Education 25: 6995.
Gilbert, J., and Swift, D. (1985). Towards a Lakatosian analysis of
the Piagetian and alternative conceptions research programs.
Science Education 69: 681696.
Gilbert, J. K., Osborne, R. J., and Fensham, P. J. (1982). Childrens
science and its consequences for teaching. Science Education
66: 623633.
Gilbert, J. K., and Zylberstajn, A. (1985). A conceptual framework
for science education: The case study of force and movement.
European Journal of Science Education 7: 107120.
Goh, N. K., Khoo, L. E., and Chia, L. S. (1993). Some misconceptions in chemistry: A cross-cultural comparison, and implications for teaching. Australian Science Teachers Journal 39:
6568.
Gonzalez, F. M. (1997). Diagnosis of Spanish primary school students common alternative science conceptions. School Science
and Mathematics 97: 68.
Good, R. (1991). Editorial. Journal of Research in Science Teaching
28: 387.
Gorin, G. (1994). Mole and chemical amount. Journal of Chemical
Education 71: 114116.
Gorodetsky, M., and Gussarsky, E. (1986). Misconceptualization
of the chemical equilibrium concept as revealed by different
evaluation methods. European Journal of Science Education
8: 427441.
Grayson, D. J., Anderson, T. R., and Crossley, L. G. (2001). A fourlevel framework for identifying and classifying student conceptual and reasoning difficulties. International Journal of Science
Education 23: 611622.
Griffiths, A. K. (1994). A critical analysis and synthesis of research
on students chemistry misconceptions. In Schmidt, H.-J. (Ed.),
Proceedings of the 1994 International Symposium on Problem
Solving and Misconceptions in Chemistry and Physics, The International Council of Association for Science Education Publications, pp. 7099.
Griffiths, A. K., and Preston, K. R. (1992). Grade-12 students misconceptions relating to fundamental characteristics of atoms
and molecules. Journal of Research in Science Teaching 29:
611628.
Griffiths, A. K., Thomey, K., Cooke, B., and Normore, G. (1988). Remediation of student-specific misconceptions relating to three
science concepts. Journal of Research in Science Teaching 25:
709719.
Gussarsky, E., and Gorodetsky, M. (1988). On the chemical equilibrium concept: Constrained word associations and conception.
Journal of Research in Science Teaching 25: 319333.
Gussarsky, E., and Gorodetsky, M. (1990). On the concept chemical equilibrium: The associative framework. Journal of Research in Science Teaching 27: 197204.
Guzzetti, B. J. (2000). Learning counter intuitive science concepts:
What have we learned from over a decade of research? Reading, Writing, Quarterly 16: 8995.
Hackling, M. W., and Garnett, P. J. (1985). Misconceptions of chemical equilibrium. European Journal of Science Education 7: 205
214.
Haidar, A. H. (1997). Prospective chemistry teachers conceptions
of the conservation of matter and related concepts. Journal of
Research in Science Teaching 34: 181197.
Halloun, I. A., and Hestenes, D. (1985a). The initial knowledge
state of college physics students. American Journal of Physics
53: 10431055.
Halloun, I. A., and Hestenes, D. (1985b). Common sense concepts about motion. American Journal of Physics 53: 1056
1065.
Hameed, H., Hackling, M. W., and Garnett, P. J. (1993). Facilitating conceptual change in chemical equilibrium using a CAI
strategy. International Journal of Science Education 15: 221
230.

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

Misconceptions in Chemical Bonding


Hand, B., and Treagust, D. (1991). Student achievement and science curriculum development using a constructivist framework. School Science and Mathematics 91: 172176.
Harrison, A. G., and Treagust, D. (2000). Learning about atoms,
molecules, and chemical bonds: A case study of multiplemodel use in grade 11 chemistry. Science Education 84: 352
381.
Hashweh, M. Z. (1988). Descriptive studies of students conceptions in science. Journal of Research in Science Teaching 25:
121134.
Helm, H. (1980). Misconceptions in physics amongst South African
students. Physics Education 15: 9197.
Hesse, J. J., and Anderson, C. W. (1992). Students conceptions of
chemical change. Journal of Research in Science Teaching 29:
277299.
Hewson, P. W., and Hewson, M. G. (1984). The role of conceptual conflict in conceptual change and the design of science
instruction. Instructional Science 13: 113.
Hills, G. (1983). Misconceptions misconceived? Using conceptual
change to understand some of the problems pupils have in
learning in science. In Proceedings of the International Seminar
on Misconceptions in Science and Mathematics, June Cornell
University, New York, pp. 245256.
Huddle, P. A., and Pillay, A. E. (1996). An in-depth study of misconceptions in stoichiometry and chemical equilibrium at a South
African University. Journal of Research in Science Teaching 33:
6577.
Hunt, E., and Minstrell, J. (1997). Effective instruction in science and mathematics: Psychological principles and social constraints. Issues in Education: Contributions from Educational
Psychology.
Johnstone, A. H., and Kellett, N. C. (1980). Learning difficulties
in school science-toward a working hypothesis. International
Journal of Science Education 2: 171181.
Khalid, T. (2003). Pre-service high school teachers perceptions of
three environmental phenomena. Environmental Education
Research 9: 3550.
Krishnan, S. R., and Howe, A. C. (1994). The mole concept: Developing an instrument to assess conceptual understanding.
Journal of Chemical Education 71(8): 653658.
Kuiper, J. (1994). Student ideas of science concepts: alternative
frameworks? International Journal of Science Education 16:
279292.
Lawson, A. E., and Thompson, L. D. (1988). Formal reasoning ability and misconceptions concerning genetics and natural selection. Journal of Research in Science Teaching 25: 733746.
Lee, O., Eichinger, D. C., Anderson, C. W., and Berkheimer, G. D.
(1993). Changing middle school students conceptions of matter and molecules. Journal of Research in Science Teaching 30:
249270.
Lee, Y., and Law, N. (2001). Explorations in promoting conceptual
change in electrical concepts via ontological category shift.
International Journal of Science Education 23: 111149.
Lord, T. R. (1999). A comparison between traditional and constructivist teaching in environmental science. Journal of Environmental Education 30: 2228.
Maskill, R., and Cachapuz, A. F. C. (1989). Learning about the
chemistry topic of equilibrium: The use of word association
tests to detect developing conceptualizations. International
Journal of Science Education 11: 5769.
Michael, J. (2002). Misconceptionswhat students think they
know? Advances in Physiology Education 26: 56.
Nakhleh, M. B. (1992). Why some students dont learn chemistry?
Chemical misconceptions. Journal of Chemical Education 69:
191196.
Nakhleh, M. B., and Krajcik, J. S. (1994). Influence of levels of information as presented by different technologies on students
understanding of acids, base and pH concepts. Journal of Research in Science Teaching 34: 10771096.

23:24

Style file version June 20th, 2002

157
Nakhleh, M. B., and Samarapungavan, A. (1999). Elementary
school childrens beliefs about matter. Journal of Research in
Science Teaching 36: 777805.
Nelson, P. G. (1991). The elusive mole. Education and Chemistry
28: 103104.
Niaz, M. (1995). Relationship between student performance on
conceptual and computational problems of chemical equilibrium. International Journal of Science Education 17: 343355.
Niaz, M. (1998). A Lakatosian conceptual change teaching strategy
based on student ability to build models with varying degrees
of conceptual understanding of chemical equilibrium. Science
and Education 7: 107127.
Niaz, M. (2001a). Response to contradiction: Conflict resolution
strategies used by students in solving problems of chemical
equilibrium. Journal of Science Education and Technology 10:
205211.
Niaz, M. (2001b). A rational reconstruction of the origin of the
covalent bond and its implications for general chemistry textbooks. International Journal of Science Education 23: 623
641.
Nicoll, G. (2001). A report of undergraduates bonding misconceptions. International Journal of Science Education 23: 707
730.
Novak, J. D. (1977). Theory of Education, Cornell University Press,
Ithaca, NY.
Nussbaum, J. (1981). Towards a diagnosis by science teachers of
pupils misconceptions: An exercise with student teachers. International Journal of Science Education 3: 159169.
Osborne, R. J. (1982). Science education: Where do we start? The
Australian Science Teachers Journal 28: 2130.
Osborne, R. J., Bell, B. F., and Gilbert, J. K. (1983). Science teaching
and childrens views of the world. European Journal of Science
Education 5: 114.
Osborne, R. J., and Cosgrove, M. M. (1983). Childrens conceptions
of the changes of state of water. Journal of Research in Science
Teaching 20: 825838.
Osborne, R. J., and Gilbert, J. K. (1980). A method for investigating
concept understanding in science. European Journal of Science
Education 2: 311321.
Osborne, R. J., and Wittrock, M. C. (1983). Learning science: A
generative process. Science Education 67: 489508.

Ozmen,
H., and Ayas, A. (2003). Students difficulties in understanding of the conservation of the matter in open and closedsystem chemical reactions. Chemistry Education: Research and
Practice 4: 279290.

Ozmen,
H., Ayas, A., and Costu, B. (2002). Determination of the
science student teachers understanding level and misunderstandings about the particulate nature of the matter. Educational Sciences: Theory and Practice 2: 507529.
Palmer, D. (1999). Exploring to link between students scientific and nonscientific conceptions. Science Education 83: 639
653.
Palmer, D. (2001). Students alternative conceptions and scientifically acceptable conceptions about gravity. International Journal of Science Education 23: 691706.
Papageorgiou, G., and Sakka, D. (2000). Primary school teachers
views of fundamental chemical concepts. Chemistry Education: Research and Practice in Europe 1: 237247.
Pedrosa, M. A., and Dias, M. H. (2000). Chemistry textbook
approaches to chemical equilibrium and student alternative
conceptions. Chemistry Education: Research and Practice in
Europe 1: 227236.
Peterson, R., and Treagust, D. F. (1989). Grade-12 students misconceptions of covalent bonding and structure. Journal of Chemical Education 66: 459460.
Peterson, R., Treagust, D. F., and Garnett, P. (1986). Identification of secondary students misconceptions of covalent bonding and the structure concepts using a diagnostic instrument.
Research in Science Education 16: 4048.

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

158
Peterson, R., Treagust, D. F., and Garnett, P. (1989). Development
and application of a diagnostic instrument to evaluate grade-11
and -12 students concepts of covalent bonding and structure
following a course of instruction. Journal of Research in Science
Teaching 26: 301314.
Pines, L., and West, L. (1986). Conceptual understanding and science learning: An interpretation of research within a source
of knowledge framework. Science Education 70: 583604.
Posner, G. J., Strike, K. A., Hewson, P. W., and Gertzog, W. A.
(1982). Accommodation of a scientific conception: Towards a
theory of conceptual change. Science Education 66: 211217.
Prawat, R. (1989). Promoting access to knowledge, strategy, and
disposition of students: A research synthesis. Review of Educational Research 59: 141.
Preece, P. (1984). Intuitive science: Learned and triggered? European Journal of Science Education 6: 710.
Quilez-Pardo, J., and Solaz-Portoles, J. (1995). Students and teachers misapplication of Le Chateliers principle: Implications for
the teaching of chemical equilibrium. Journal of Research in
Science Teaching 32: 939957.
Resnik, L. (1983). Mathematics and science learning: A new conception. Science Education 64: 5984.
Robinson, W. R. (1998). An alternative framework for chemical
bonding. Journal of Chemical Education 75: 10741075.
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.
Schmidt, H.-J. (1994). Stoichiometric problem solving in high
school chemistry. International Journal of Science Education
6: 191200.
Schmidt, H.-J. (1997). Students misconceptions-looking for a pattern. Science Education 81: 123135.
Schoon, J. K., and Boone, J. W. (1998). Self-efficacy and alternative conceptions of science of preservice elementary teachers.
Science Education 82: 553568.
Shuell, T. (1987). Cognitive psychology and conceptual change: Implications for teaching science. Science Education 71: 239250.
Sisovic, D., and Bojovic, S. (2000). Approaching the concepts of
acids and bases by cooperative learning. Chemistry Education:
Research and Practice in Europe 1: 263275.
Skamp, K. (1999). Are atoms and molecules too difficult for primary
children? School Science Review 81: 8796.
Smith, K. J., and Metz, P. A. (1996). Evaluating student understanding of solution chemistry through microscopic representations.
Journal of Chemical Education 73: 233235.
Southerland, S. A., Abrams, E., Cummins, C. L., and Anzelmo,
J. (2001). Understanding students explanations of biological
phenomena: Conceptual frameworks or P-prims? Science Education 85: 328348.
Soyibo, K. (1995). Using concept maps to analyze textbook presentation of respiration. The American Biology Journal 57:
344351.
Stavy, R. (1991). Using analogy to overcome misconceptions about
conservation of matter. Journal of Research in Science Teaching
28: 305313.
Sutton, C. R. (1980). The learners prior knowledge: A critical review of techniques for probing its organization. European Journal of Science Education 2: 107120.
Taber, K. (2000). Chemistry lessons for universities?: A review of
constructivist ideas. University Chemistry Education 4: 6372.
Taber, K. S. (1994). Misunderstanding the ionic bond. Education in
Chemistry 31: 100103.
Taber, K. S. (1995). Development of student understanding: A case
study of stability and lability in cognitive structure. Research
in Science and Technological Education 13: 8999.
Taber, K. S. (1997). Student understanding of ionic bonding: Molecular versus electrostatic framework? School Science Review 78:
8595.

23:24

Style file version June 20th, 2002

Ozmen
Taber, K. S. (1998). An alternative conceptual framework from
chemistry education. International Journal of Science Education 20: 597608.
Taber, K. S. (1999). Alternative frameworks in chemistry. Education
in Chemistry 36: 135137.
Taber, K. S. (2001). Constructing chemical concepts in the classroom?: Using research to inform the practice. Chemistry Education: Research and Practice in Europe 2: 4351.
Taber, K. S., and Watts, M. (1997). Constructivism and concept
learning in chemistry: Perspectives from a case study. Research
in Education 58: 1020.
Tan, K. C., Goh, N. K., Chia, L. S., and Treagust, D. F. (2002). Development and application of a two-tier multiple choice diagnostic instrument to assess high school students understanding of
inorganic chemistry qualitative analysis. Journal of Research
in Science Teaching 39: 283301.
Tan, K. C., and Treagust, D. (1999). Evaluating students understanding of chemical bonding. School Science Review 81: 75
84.
Teichert, M. A., and Stacy, A. M. (2002). Promoting understanding
of chemical bonding and spontaneity through student explanation and integration of ideas. Journal of Research in Science
Teaching 39: 464496.
Treagust, D. F. (1988). Development and use of diagnostic tests
to evaluate students misconceptions in science. International
Journal of Science Education 10: 159169.
Tsai, C.-C. (1996). The Interrelations Between Junior High
School Students Scientific Epistemological Beliefs, Learning Environment Preferences and Cognitive Structure Outcomes, Doctoral dissertation, Teachers College, Columbia
University.
Tsai, C.-C. (1998). The constructivist epistemology: The interplay
between the philosophy of science and students science learning. Curriculum and Teaching 13(1).
Tsaparlis, G. (1997). Atomic and molecular structure in chemical education: A critical analysis from various perspectives
of science education. Journal of Chemical Education 74: 922
925.
Tsaparlis, G., Kousathana, M., and Niaz, M. (1998). Molecularequilibrium problems: Manipulation of logical structure and
of M-demand, and their effect on students performance. Science Education 82: 437454.
Tyson, L., Treagust, D. F., and Bucat, R. B. (1999). The complexity of teaching and learning chemical equilibrium. Journal of
Chemical Equilibrium 76: 554558.
Tytler, R. (2000). A comparison of year 1 and year 6 students conceptions of evaporation and condensation: Dimension of conceptual progression. International Journal of Science Education 22: 447467.
Valanides, N. (2000). Primary student teachers understanding of
the particulate nature of matter and its transformations during dissolving. Chemistry Education: Research and Practice in
Europe 1: 249262.
Van Driel, J. H. (2002). Students corpuscular conceptions the context of chemical equilibrium and chemical kinetics. Chemistry Education: Research and Practice in Europe 3: 201
213.
Von Glasersfeld, E. (1992). A constructivist view of teaching and
learning. In Duit, R., Goldberg, F., and Niedderer, H. (Eds.),
Research in Physics Learning: Theoretical Issues and Empirical
Studies, IPN, Kiel, Germany, pp. 2939.
Voska, K. W., and Heikkinen, H. W. (2000). Identification and analysis of students conceptions used to solve chemical equilibrium problems. Journal of Research in Science Teaching 37:
160176.
Wandersee, H., Mintzes, J. J., and Novak, J. D. (1994). Research
on Alternative Conceptions in Science. In Gabel, D. L. (Ed.),
Handbook of Research on Science Teaching and Learning,
McMillan, New York, pp. 177210.

P1: JLS
Journal of Science Education and Technology

pp1240-jost-488289

May 31, 2004

Misconceptions in Chemical Bonding


Westbrook, S. L., and Marek, E. A. (1991). A cross-age of student
understanding of the concept of diffusion. Journal of Research
in Science Teaching 28: 649660.
Wheeler, A. E., and Kass, H. (1978). Student misconceptions in chemical equilibrium. Science Education 62: 223
232.

23:24

Style file version June 20th, 2002

159
White, R., and Gunstone, R. (1992). Probing Understanding,
Graphicraft, Hong Kong.
Yip, D. Y. (2001). Promoting the development of a conceptual
change model of science instruction in prospective secondary
biology teachers. International Journal of Science Education
23: 755770.

Potrebbero piacerti anche