14
Some Student Misconceptions in Chemistry: A Literature Review of Chemical Bonding Author(s): Haluk Özmen Source: Journal of Science Education and Technology, Vol. 13, No. 2 (Jun., 2004), pp. 147-159 Published by: Springer Stable URL: http://www.jstor.org/stable/40188778 . Accessed: 20/04/2014 08:50 Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at . http://www.jstor.org/page/info/about/policies/terms.jsp . JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms of scholarship. For more information about JSTOR, please contact [email protected]. . Springer is collaborating with JSTOR to digitize, preserve and extend access to Journal of Science Education and Technology. http://www.jstor.org This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AM All use subject to JSTOR Terms and Conditions

Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

Embed Size (px)

Citation preview

Page 1: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

Some Student Misconceptions in Chemistry: A Literature Review of Chemical BondingAuthor(s): Haluk ÖzmenSource: Journal of Science Education and Technology, Vol. 13, No. 2 (Jun., 2004), pp. 147-159Published by: SpringerStable URL: http://www.jstor.org/stable/40188778 .

Accessed: 20/04/2014 08:50

Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .http://www.jstor.org/page/info/about/policies/terms.jsp

.JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range ofcontent in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new formsof scholarship. For more information about JSTOR, please contact [email protected].

.

Springer is collaborating with JSTOR to digitize, preserve and extend access to Journal of Science Educationand Technology.

http://www.jstor.org

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 2: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

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

Some Student Misconceptions in Chemistry: A Literature Review of Chemical Bonding

Haluk Ozmen1

Students' misconceptions before or after formal instruction have become a major concern among researchers in science education because they influence how students learn new sci- entific 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 pre- sented from past to day historically. On the basis of the results some suggestions for teaching were made.

KEY WORDS: chemistry; misconception; chemical bonding.

INTRODUCTION

Learning science is a cumulative process and each new piece of information is added to what stu- dents already know about the topic at hand. Research has shown that children bring to lessons a lot of preex- isting (alternative) conceptions about scientific phe- nomena 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 suggest- ing 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 re- sult 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

found after teaching has taken place. This construc- tivist view is the dominant paradigm of learning in sci- ence. 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 informa- tion in terms of existing cognitive structures (Taber and Watts, 1997). The particular knowledge that is constructed by an individual will be affected by the learner's prior knowledge and experience and the so- cial context in which learning takes place (Grayson et al, 2001; Von Glasersfeld, 1992).

Students preexisting beliefs influence how stu- dents learn new scientific knowledge and play an es- sential role in subsequent learning (Arnaudin and Mintez, 1985; Boujaoude, 1991; Driver and Oldham, 1986; Shuell, 1987; Tsai, 1996). Hunt and Minstrell (1997) stated that children's difficulties in science oc- cur because students' conceptions before teaching are not taken into account and therefore communica- tion 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 differ- ent from accepted scientific viewpoints and may not be in conformity with the true or the scientific ex- planation (Osborne, 1982; Schoon and Boone, 1998).

department of Science Education, Fatih Faculty of Education, Karadeniz Technical University, 61335 Sogutlu-Trabzon, Turkey; e-mail: [email protected] and [email protected]

147

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

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 3: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

148 Ozmen

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 conven- tional 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 stu- dents' "hard-core" beliefs. Students' conceptions that are different from those accepted by the scientific community are variously labeled in the science edu- cation 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), naive beliefs (Caramazza et aL, 1981), naive theories (Resnik, 1983), naive conceptions (Champagne et aL, 1983), children's scientific in- tuitions (Sutton, 1980), conceptual frameworks (Southerland et aL, 2001), children's 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), com- mon alternative science conceptions (Gonzalez, 1997), students' intuitive theories (Boujaoude, 1992), presci- entific 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 children's er- roneous conceptions as viewed by science educators. Although Wandersee et aL (1994) presented an anal- ysis of the subtle distinctions in the usage of these terms, no consensus has been reached on the term of

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 under- standing of the term.

Of course, chemistry is one of the most impor- tant 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 lan- guage of chemistry (Ayas and Demirba§, 1997), there are two major reasons for students having difficul- ties in these areas; firstly, the topics are very abstract (Ben-Zvi et aL, 1988), and secondly, words from ev- eryday language are used but with different mean- ings (Bergquist and Heikkinen, 1990). Because stu- dents' misconceptions in school sciences at all levels constitute a major problem of concern to science ed- ucators, scientist-researchers, teachers, and students (Johnstone and Kellett, 1980; Nussbaum, 1981), the identification of the students' understandings and mis- conceptions 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 Demirba§, 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), chem- ical 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), chemi- cal 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,

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 4: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

Misconceptions in Chemical Bonding 149

2001; Smith and Metz, 1996), evaporation and con- densation (Bar and Gaglili, 1994; Chang, 1999; Tytler, 2000), and the paniculate 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 misconcep- tions is the topic of chemical bonding. This paper aims to synthesize students' misconceptions found in differ- ent 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 com- plexity (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, ioniza- tion 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

into a series of three target systems; metallic, ionic, and covalent bonding. In the science education litera- ture, there have been numerous studies to determine students' understanding and misconceptions about metallic, ionic, and covalent bonding. These studies have revealed prevalent and consistent misconcep- tions 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 stu- dents 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 un- der 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 in- vestigated students' misconceptions including chem- ical 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.

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 5: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

150 Ozmen

believed intermolecular bonding was stronger than intramolecular bonding. This was consistent with Peterson et aUs (1989) findings.

One case study conducted by Taber (1995) has in- vestigated 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 com- pounds. And also, it is stated that there is some evidence that learners appreciated the relationship between intermolecular bonding and physical proper- ties such as boiling point. These results are consistent with Peterson and Treagust's (1989), Peterson et aVs (1989), and Taber's (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 gen- eral 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 disap- pear. 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 stu- dents believe that it is absent in polar molecular sub- stances such as water. Students' misconceptions iden- tified from the exam papers are depicted in Table III.

Barker (2000) has investigated students' under- standing 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 elec- tronegativity, bonding, geometry, and microscopic representations that undergraduate chemistry stu- dents 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 that's 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 n. 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.

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 6: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

Misconceptions in Chemical Bonding 151

Table PI. 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, stu- dents 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 fresh- man student, Bridgette, stated that "It's 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 ter- tiary students. At the end of the study, some 20 mis- conceptions 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 un- dergraduate and postgraduate Australian students' mental models for chemical bonding using semistruc- tured 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 con- trast, other studies reveal that learners' mental models of bonding become sophisticated and complex mod- els they were exposed to during instruction (Coll and

Table IV. Students' Misconceptions for Chemical Bonding • Metallic bonding is weak bonding. • Metallic lattices contain neutral atoms. • Intramolecular covalent bonding is weak • Electronegativity comprises attraction for a

bonding. single electron. • Ionic bonding is weak bonding. • Molecular iodine is metallic in nature. • Continuous metallic or ionic lattices are • Ionic bonding comprises sharing of electrons,

molecular in nature. • The bonding in metals and ionic compounds • Ionic and metallic bondings contain an element

involves intermolecular bonding. of directionality. • The ionic radius of the sodium ion is greater than • Ions in close-packed metal lattices possess other

the chloride ion. than eight nearest neighbors. • The ionic radius of the lithium ion is greater than • Metal to nonmetal bonding in alloys is

the sodium ion. electrostatic in nature. • Polar covalent compounds contain charged • Ionic shape and packing is influenced by pressure,

species. • Molecular iodine contains 1 minus ions. • Intermolecular forces are influences by gravity. • The charged species in metallic lattices are nuclei • Glass is an ionic crystalline substance.

rather than ions.

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 7: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

152 Ozmen

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 physi- cal and chemical properties of a compound. An understanding of the concept of bonding is funda- mental to subsequent learning of various topics in chemistry, including chemical equilibrium, thermody- namics, 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 molecu- lar structure and chemical bonding are built upon the fundamental principles of atomic structure, this un- derstanding of chemical behavior at the atomic level appears important in understanding subsequent con- cepts 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 miscon- ceptions appear to be resistant to attempts to change them over time, despite increased chemistry educa- tion. Students pass from grade to grade without fully grasping the underlying concepts of bonding.

There may be a lot of reasons in generating mis- conceptions. In the classroom teaching, teachers gen- erally use ball and stick models to represent chemical bonds. According to Butts and Smith (1987), the ubiq- uitous 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 in- dividual chemical bonds.

A considerable amount of research has pointed out that the process of knowledge construction in- volves the replacement or reorganization of the con- ceptual 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 con- cepts encountered in the study of chemistry provide increased opportunity for the development of formal

misconceptions. Although students at each level take several science classes during their schooling in or- der to learn various science concepts including chem- ical bonding, the presence of misconceptions in their explanation indicates their fragmented understand- ing of these abstract concepts. Sometimes students have such strong misconceptions that even after learn- ing the correct concepts in the classrooms, they resist modifying their preexisting ideas. Instead, they try to interpret the new acquired knowledge using their pre- conceptions (Khalid, 2003).

It is obvious that "why misconceptions exist" is an important question in science education and in other disciplines. Although incorrect, imprecise, or incom- plete teaching may play an important role, according to Tsaparlis (1997), there must be a more fundamen- tal cause that results in one or more of the following: i) the inability of most or many students to employ for- mal operations, ii) the lack of the proper knowledge corpus which is a prerequisite for meaningful learn- ing, 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 misconcep- tions, 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 cur- riculum 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) sug- gests 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

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 8: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

Misconceptions in Chemical Bonding 153

concept building and remediate misconceptions. Only then will deep-seated misunderstandings be identi- fied, 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 knowl- edge 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 children's miscon- ceptions is to alert teachers to students' difficulties in conceptualizing science knowledge and hence sug- gest more effective strategies for improving classroom instruction. Before teaching a concept, such as chem- ical 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 meth- ods 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 un- dertaken 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 con- cepts and having extra misconceptions. The construc- tivist 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, prediction-observation-explanation, drawings, word association, pencil-and-paper diagnostic instruments

based on multiple-choice items, two-tier multiple- choice tests (Peterson et aL9 1989; Schmidt, 1997; White and Gunstone, 1992), that can be used by teach- ers in their classroom environment in identifying mis- conceptions of science phenomena. Of these many approaches, interviews, and multiple-choice diagnos- tic tests are most common methodologies and have ac- quired strong support as a viable approach (Osborne and Gilbert, 1980; Peterson etaL, 1989). But according to Treagust (1988), conventional multiple-choice tests do not adequately assess student understanding. Al- though multiple-choice tests have been used to evalu- ate students' content knowledge, they have some lim- itations with determining students' reasoning behind their choices. However, many instructors agree that one of the best ways to measure student understand- ing 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 evaluat- ing students' understanding, and for identifying com- monly held misconceptions (Peterson and Treagust, 1989). The items in two-tier multiple-choice diagnos- tic instruments are specifically designed to identify students' misconceptions and misunderstandings in a limited content area. The first part of each item con- sists 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 litera- ture, interviews, and free response tests (Tan et al, 2002). In addition, this type of test is more readily ad- ministered and scored than the other methods, and are useful for classroom teachers (Tan and Treagust, 1999). But on the other hand, objectively scored two- tier 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 student's reason- ing 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 ai, 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

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 9: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

154 Ozmen

assessment tests that can be used in the classrooms should be also developed by teachers to determine the level of students' understanding and misconcep- tions. 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 identify- ing their own students' conceptions and implement- ing teaching approaches that promote conceptual un- derstanding 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 infor- mation to prior knowledge, to integrate information for one subject area into another, and to relate class- room information to everyday experiences to help those students become meaningful learners who are better able to retain and use information in novel situ- ations (Prawat, 1989). A majority of teachers and even professors use teacher-centered teaching strategies to teach science (Lord, 1999; Yip, 2001). To be success- ful in examinations, pupils are trained to be good at retrieving factual information and the rote applica- tion of algorithms. These traditional teaching strate- gies 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 ineffec- tive 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 chem- istry 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 ac- tively construct their own knowledge by using ex- isting knowledge to make sense of newly gained ex- periences. Taber (2000) has stated that the first step in a constructivist learning approach is to make the teacher and student aware of the learner's current ideas. Teaching can then be planned that challenges misconceptions, and provides students with the op-

portunities and rationale for conceptual restructur- ing. 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 understand- ing 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 stu- dents' 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 concept's meaning has been completely re- moved and replaced by something else that is incom- parable to the existing meaning, it would be consid- ered a conceptual change (Chiu et ai, 2002). Within this perspective, learning is depicted as a process of conceptual change. This approach represents an al- ternative 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 expe- riences 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 experi- ences can be structured by the new concept, (iii) the new conception must appear plausible; any new con- cept 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 instruc- tional strategies have been proposed to help students change their misconceptions. Among these, refuta- tional texts and conceptual change texts have be- come popular for the last two decades. As stated by Chambers and Andre (1997), the major difference between the refutational text model and the concep- tual change text involves whether students are asked to make a prediction about a situation. In the con- ceptual change model, students are asked to predict what would happen in a situation before being pre- sented with information that demonstrates the incon- sistency between common misconceptions and the scientific conception. In the refutational text model, common misconceptions are contrasted to scientific

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 10: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

Misconceptions in Chemical Bonding 155

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 sci- ence teachers are aware of conceptual-change teach- ing techniques at present time (Hesse and Anderson, 1992). That the teachers should be informed about us- ing 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 in- correct information (Soyibo, 1995). Therefore, text- books 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 dis- semination of misconceptions on important chemistry topics (Haidar, 1997), it is sensible and adequate to ini- tiate 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: 105-120.

Andersson, B. (1990). Pupils' conceptions of matter and its trans- formations (age 12-16). Studies in Science Education 18: 53-85.

Arnaudin, M. W, and Mintez, J. J. (1985). Students' alternative con- ceptions of the human circulatory system: A cross-age study. Science Education 69: 721-733.

Astudillo, L. R., and Niaz, M. (1996). Reasoning strategies used by students to solve stoichiometry problems and its relation- ship 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: 518-521.

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 8-10, Gazimagusa, TUrkish Republic of Northern Cyprus.

tfanerjee, A. U, and rower, c. IN. (lyyi). ihe development ot mod- ules for the teaching of chemical equilibrium. International Journal of Science Education 13: 355-362.

Bar, V., and Gaglili, I. (1994). Stages of children's views about evap- oration. International Journal of Science Education 16: 157- 174.

Barker, V. (2000). Students' reasoning about basic chemical ther- modynamics and chemical bonding: What changes occur dur- ing a context-based post-16 chemistry course? International Journal of Science Education 22: 1171-1200.

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' visu- alization of a chemical reaction. Education in Chemistry 24: 117-120.

Ben-Zvi, R., Eylon, B., and Silberstein, J. (1988). Theories, princi- ples and laws. Education in Chemistry 25: 89-92.

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: 124-128.

Bodner, u. (1956). Constructivism: A theory ot knowledge. Journal of Chemical Education 63: 873-878.

Boo, H. K. (1998). Students' understanding of chemical bonds and the energetic of chemical reactions. Journal of Research in Sci- ence Teaching 35: 569-581.

Boo, H. K., and Watson, J. R. (2001). Progression m high school students' (aged 16-18) conceptualizations about chemical re- actions in solution. Science Education 85: 568-585.

Boujaoude, S. B. (1991). A study of the nature of students' under- standing about the concept of burning. Journal of Research in Science Teaching 28: 689-704.

Boujaoude, S. B. (1992). The relationship between students' learn- ing strategies and the change in their misunderstandings during a high school chemistry course. Journal of Research in Science Teaching 29: 687-699.

Bradley, J. D., and Mosimege, M. D. (1998). Misconceptions in acids and bases: A comparative study of student teachers with dif- ferent chemistry backgrounds. South African Journal of Chem- istry 51: 137-147.

Brown, D. E. (1992). Using examples and analogies to remedi- ate misconceptions in physics: Factors influencing concep- tual change. Journal of Research in Science Teaching 29: 17- 34.

Butts, B., and Smith, R. (1987). HSC chemistry students' under- standing of the structure and properties of molecular and ionic compounds. Research in Science Education 17: 192-201.

Caramazza, A., McCloskey, M., and Green, B. (1981). Naive beliefs in "sophisticated" subjects: Misconceptions about trajectories of objects. Cognition 9: 117-123.

Carter, C. S., and Brickhouse, N. W. (1989). What makes chemistry difficult? Alternate perceptions. Journal of Chemical Educa- tion 66: 223-225.

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: 107-123.

Champagne, A., Gunstone, R., and Klopfer, L. (1983). Naive knowledge and science learning. Research in Science and Tech- nological Education 1: 173-183.

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 11: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

156 Ozmen

Champagne, A. B., Klopfer, L. E., and Gunstone, R. F. (1982). Cognitive research and the design of science instruction. Ed- ucational Psychologist 17: 31-53.

Chang, J. Y. (1999). Teacher collage students' conception about evaporation, condensation, and boiling. Science Education 83: 511-526.

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: 688-712.

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: 707-719.

Clement, J. (1982). Students' preconceptions in introductory me- chanics. American Journal of Physics 50: 66-71.

Coll, R. K., and Taylor, N. (2001). Alternative conceptions of chem- ical bonding held by upper secondary and tertiary students. Research in Science and Technological Education 19: 171-191.

Coll, R. K., and Taylor, N. (2002). Mental models in chemistry: Senior chemistry students' mental models of chemical bond- ing. Chemistry Education: Research and Practice in Europe 3: 175-184.

Coll, R. K., and Treagust, D. F. (2001). Learners' mental models of chemical bonding. Research in Science Education 31: 357-382.

Coll, R. K., and Treagust, D. F. (2002). Exploring tertiary students' understanding of covalent bonding. Research in Science and Technological Education 20: 241-267.

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: 464-486.

De Jong, O. (2000). Crossing the borders: Chemical education re- search and teaching practice. University Chemistry Education 4: 29-32.

De Vos, W., and Verdonk, A. H. (1996). The participate nature of matter in science education and in science. Journal of Research in Science Teaching 33: 657-664.

Din, Y. (1998). Children's misconceptions on reproduction and im- plications for teaching. Journal of Biological Education 33: 21.

Driver, R. (1981). Pupils' alternative frameworks in science. Euro- pean Journal of Science Education 3: 93-101.

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: 61-84.

Driver, R., and Enckson, G. (1983). Theories-in-action: Some theo- retical and empirical issues in the study of students' conceptual frameworks in science. Studies in Science Education 10: 37-60.

Driver, R., and Oldham, V. (1986). A construed vist approach to cur- riculum development in science. Studies in Science Education 13: 105-122.

Ebenezer, J. V., and Erickson, L. G. (1996). Chemistry students' conception of solubility: A phenomenography. Science Edu- cation 80: 181-201.

Ebenezer, J. V., and Fraser, M. D. (2001). First year chemical en- gineering students' conceptions of energy in solution process: Phenomenographic categories for common knowledge con- struction. Science Education 85: 509-535.

Fensham, P. (1975). Concept formation. In Daniels, D. J. (Ed.), New Movements in the Study and Teaching of Chemistry , Temple Smith, London, pp. 199-217.

Fleer, M. (1999). Children's alternative views: Alternative to what? International Journal of Science Education 21: 119-135.

huno, (J., Azcona, R., Ciuisasola, J., and Ratclirle, M., (2000). Diffi- culties in teaching the concept of amount of substance and mole. International Journal of Science Education 22: 1285- 1304.

Garnett, P. J., Garnett, P. I, and Hackling, M. W. (1995). Students' alternative conceptions in chemistry: A review of research and

implications for teaching and learning. Studies in Science Ed- ucation 25: 69-95.

Gilbert, I, and Swift, D. (1985). Towards a Lakatosian analysis of the Piagetian and alternative conceptions research programs. Science Education 69: 681-696.

Gilbert, J. K., Osborne, R. J., and Fensham, P. J. (1982). Children's science and its consequences for teaching. Science Education 66: 623-633.

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: 107-120.

Goh, N. K., Khoo, L. E., and Chia, L. S. (1993). Some miscon- ceptions in chemistry: A cross-cultural comparison, and impli- cations for teaching. Australian Science Teachers Journal 39: 65-68.

Gonzalez, F. M. (1997). Diagnosis of Spanish primary school stu- dents' 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: 114-116.

Gorodetsky, M., and Gussarsky, E. (1986). Misconceptualization of the chemical equilibrium concept as revealed by different evaluation methods. European Journal of Science Education 8: 427-441.

Grayson, D. I, Anderson, T. R., and Crossley, L. G. (2001). A four- level framework for identifying and classifying student concep- tual and reasoning difficulties. International Journal of Science Education 23: 611-622.

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 In- ternational Council of Association for Science Education Pub- lications, pp. 70-99.

Griffiths, A. K., and Preston, K. R. (1992). Grade-12 students' mis- conceptions relating to fundamental characteristics of atoms and molecules. Journal of Research in Science Teaching 29: 611-628.

Griffiths, A. K., Thomey, K., Cooke, B., and Normore, G (1988). Re- mediation of student-specific misconceptions relating to three science concepts. Journal of Research in Science Teaching 25: 709-719.

Gussarsky, E., and Gorodetsky, M. (1988). On the chemical equilib- rium concept: Constrained word associations and conception. Journal of Research in Science Teaching 25: 319-333.

Gussarsky, E., and Gorodetsky, M. (1990). On the concept "chem- ical equilibrium": The associative framework. Journal of Re- search in Science Teaching 27: 197-204.

Guzzetti, B. J. (2000). Learning counter intuitive science concepts: What have we learned from over a decade of research? Read- ing, Writing, Quarterly 16: 89-95.

Hackling, M. W., and Garnett, P. J. (1985). Misconceptions of chemi- cal 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: 181-197.

Halloun, I. A., and Hestenes, D. (1985a). The initial knowledge state of college physics students. American Journal of Physics 53: 1043-1055.

Halloun, I. A., and Hestenes, D. (1985b). Common sense con- cepts about motion. American Journal of Physics 53: 1056- 1065.

Hameed, H., Hackling, M. W., and Garnett, P. J. (1993). Facilitat- ing conceptual change in chemical equilibrium using a CAI strategy. International Journal of Science Education 15: 221- 230.

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 12: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

Misconceptions in Chemical Bonding 157

Hand, B., and Treagust, D. (1991). Student achievement and sci- ence curriculum development using a constructivist frame- work. School Science and Mathematics 91: 172-176.

Harrison, A. G., and Treagust, D. (2000). Learning about atoms, molecules, and chemical bonds: A case study of multiple- model use in grade 11 chemistry. Science Education 84: 352- 381.

Hashweh, M. Z. (1988). Descriptive studies of students' concep- tions in science. Journal of Research in Science Teaching 25: 121-134.

Helm, H. (1980). Misconceptions in physics amongst South African students. Physics Education 15: 91-97.

Hesse, J. I, and Anderson, C. W. (1992). Students' conceptions of chemical change. Journal of Research in Science Teaching 29: 277-299.

Hewson, P. W., and Hewson, M. G. (1984). The role of concep- tual conflict in conceptual change and the design of science instruction. Instructional Science 13: 1-13.

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. 245-256.

Huddle, P. A., and Pillay, A. E. (1996). An in-depth study of miscon- ceptions in stoichiometry and chemical equilibrium at a South African University. Journal of Research in Science Teaching 33: 65-77.

Hunt, E., and Minstrell, J. (1997). Effective instruction in sci- ence and mathematics: Psychological principles and social con- straints. 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: 171-181.

Khalid, T. (2003). Pre-service high school teachers' perceptions of three environmental phenomena. Environmental Education Research 9: 35-50.

Krishnan, S. R., and Howe, A. C. (1994). The mole concept: De- veloping an instrument to assess conceptual understanding. Journal of Chemical Education 71(8): 653-658.

Kuiper, J. (1994). Student ideas of science concepts: alternative frameworks? International Journal of Science Education 16: 279-292.

Lawson, A. E., and Thompson, L. D. (1988). Formal reasoning abil- ity and misconceptions concerning genetics and natural selec- tion. Journal of Research in Science Teaching 25: 733-746.

Lee, O., Eichinger, D. C, Anderson, C. W., and Berkheimer, G. D. (1993). Changing middle school students' conceptions of mat- ter and molecules. Journal of Research in Science Teaching 30: 249-270.

Lee, Y, and Law, N. (2001). Explorations in promoting conceptual change in electrical concepts via ontological category shift. International Journal of Science Education 23: 111-149.

Lord, T. R. (1999). A comparison between traditional and construc- tivist teaching in environmental science. Journal of Environ- mental Education 30: 22-28.

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: 57-69.

Michael, J. (2002). Misconceptions - what students think they know? Advances in Physiology Education 26: 5-6.

Nakhleh, M. b. (lyyz). Why some students don t learn chemistry? Chemical misconceptions. Journal of Chemical Education 69: 191-196.

Nakhleh, M. B., and Krajcik, J. S. (1994). Influence of levels of in- formation as presented by different technologies on students' understanding of acids, base and pH concepts. Journal of Re- search in Science Teaching 34: 1077-1096.

Nakhleh, M. B., and Samarapungavan, A. (1999). Elementary school children's beliefs about matter. Journal of Research in Science Teaching 36: 777-805.

Nelson, P. G. (1991). The elusive mole. Education and Chemistry 28: 103-104.

Niaz, M. (1995). Relationship between student performance on conceptual and computational problems of chemical equilib- rium. International Journal of Science Education 17: 343-355.

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: 107-127.

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: 205-211.

Niaz, M. (2001b). A rational reconstruction of the origin of the covalent bond and its implications for general chemistry text- books. International Journal of Science Education 23: 623- 641.

Nicoll, G. (2001). A report of undergraduates' bonding miscon- ceptions. 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. In- ternational Journal of Science Education 3: 159-169.

Osborne, R. J. (1982). Science education: Where do we start? The Australian Science Teachers' Journal 2%: 21-30.

Osborne, R. J., Bell, B. E, and Gilbert, J. K. (1983). Science teaching and children's views of the world. European Journal of Science Education 5: 1-14.

Osborne, R. X, and Cosgrove, M. M. (1983). Children's conceptions of the changes of state of water. Journal of Research in Science Teaching 20: 825-838.

Usborne, K. J., and uilbert, J. K. (lysuj. A method tor investigating concept understanding in science. European Journal of Science Education 2: 311-321.

Osborne, R. J., and Wittrock, M. G (1983). Learning science: A generative process. Science Education 67: 489-508.

Ozmen, H., and Ayas, A. (2003). Students' difficulties in under- standing of the conservation of the matter in open and closed- system chemical reactions. Chemistry Education: Research and Practice 4: 279-290.

Ozmen, H., Ayas, A., and Co§tu, B. (2002). Determination of the science student teachers' understanding level and misunder- standings about the participate nature of the matter. Educa- tional Sciences: Theory and Practice 2: 507-529.

Palmer, D. (1999). Exploring to link between students' scien- tific and nonscientific conceptions. Science Education 83: 639- 653.

Palmer, D. (2001). Students' alternative conceptions and scientifi- cally acceptable conceptions about gravity. International Jour- nal of Science Education 23: 691-706.

Papageorgiou, G., and Sakka, D. (2000). Primary school teachers' views of fundamental chemical concepts. Chemistry Educa- tion: Research and Practice in Europe 1: 237-247.

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: 227-236.

Peterson, R., and Treagust, D. F. (1989). Grade-12 students' miscon- ceptions of covalent bonding and structure. Journal of Chem- ical Education 66: 459-460.

Peterson, R., Treagust, D. E, and Garnett, P. (1986). Identifica- tion of secondary students' misconceptions of covalent bond- ing and the structure concepts using a diagnostic instrument. Research in Science Education 16: 40-48.

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 13: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

158 Ozmen

Peterson, R., Treagust, D. E, 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: 301-314.

Pines, L., and West, L. (1986). Conceptual understanding and sci- ence learning: An interpretation of research within a source of knowledge framework. Science Education 70: 583-604.

Posner, G. I, 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: 211-217.

Prawat, R. (1989). Promoting access to knowledge, strategy, and disposition of students: A research synthesis. Review of Edu- cational Research 59: 1-41.

Preece, P. (1984). Intuitive science: Learned and triggered? Euro- pean Journal of Science Education 6: 7-10.

Quilez-Pardo, J., and Solaz-Portoles, J. (1995). Students' and teach- ers' misapplication of Le Chatelier's principle: Implications for the teaching of chemical equilibrium. Journal of Research in Science Teaching 32: 939-957.

Resnik, L. (1983). Mathematics and science learning: A new con- ception. Science Education 64: 59-84.

Robinson, W. R. (1998). An alternative framework for chemical bonding. Journal of Chemical Education 75: 1074-1075.

Sanger, M. J. (2000). Addressing student misconceptions concern- ing electron flow in aqueous solutions with instruction includ- ing computer animations and conceptual change strategies. In- ternational Journal of Science Education 22: 521-537.

Schmidt, H.-J. (1994). Stoichiometric problem solving in high school chemistry. International Journal of Science Education 6: 191-200.

Schmidt, H.-J. (1997). Students' misconceptions-looking for a pat- tern. Science Education 81: 123-135.

Schoon, J. K., and Boone, J. W. (1998). Self-efficacy and alterna- tive conceptions of science of preservice elementary teachers. Science Education 82: 553-568.

Shuell, T. (1987). Cognitive psychology and conceptual change: Im- plications for teaching science. Science Education 71: 239-250.

Sisovic, D., and Bojovic, S. (2000). Approaching the concepts of acids and bases by cooperative learning. Chemistry Education: Research and Practice in Europe 1: 263-275.

Skamp, K. (1999). Are atoms and molecules too difficult for primary children? School Science Review 81: 87-96.

smith, K. J., and Metz, F. A. (1996). Evaluating student understand- ing of solution chemistry through microscopic representations. Journal of Chemical Education 73: 233-235.

boutherland, b. A., Abrams, E., Cummins, C. L., and Anzelmo, J. (2001). Understanding students' explanations of biological phenomena: Conceptual frameworks or P-prims? Science Ed- ucation 85: 328-348.

Soyibo, K. (1995). Using concept maps to analyze textbook pre- sentation of respiration. The American Biology Journal 57: 344-351.

Stavy, R. (1991). Using analogy to overcome misconceptions about conservation of matter. Journal of Research in Science Teaching 28: 305-313.

Sutton, C. R. (1980). The learner's prior knowledge: A critical re- view of techniques for probing its organization. European Jour- nal of Science Education 2: 107-120.

Taber, K. (2000). Chemistry lessons for universities?: A review of constructivist ideas. University Chemistry Education 4: 63-72.

Taber, K. S. (1994). Misunderstanding the ionic bond. Education in Chemistry 31: 100-103.

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: 89-99.

Taber, K. S. (1997). Student understanding of ionic bonding: Molec- ular versus electrostatic framework? School Science Review 78: 85-95.

Taber, K. S. (1998). An alternative conceptual framework from chemistry education. International Journal of Science Educa- tion 20: 597-608.

Taber, K. S. (1999). Alternative frameworks in chemistry. Education in Chemistry 36: 135-137.

Taber, K. S. (2001). Constructing chemical concepts in the class- room?: Using research to inform the practice. Chemistry Ed- ucation: Research and Practice in Europe 2: 43-51.

Taber, K. S., and Watts, M. (1997). Constructivism and concept learning in chemistry: Perspectives from a case study. Research in Education 58: 10-20.

Tan, K. C, Goh, N. K., Chia, L. S., and Treagust, D. F. (2002). Devel- opment and application of a two-tier multiple choice diagnos- tic instrument to assess high school students' understanding of inorganic chemistry qualitative analysis. Journal of Research in Science Teaching 39: 283-301.

Tan, K. C, and Treagust, D. (1999). Evaluating students' under- standing 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 expla- nation and integration of ideas. Journal of Research in Science Teaching 39: 464-496.

Treagust, D. E (1988). Development and use of diagnostic tests to evaluate students' misconceptions in science. International Journal of Science Education 10: 159-169.

Tsai, C.-C. (1996). The Interrelations Between Junior High School Students' Scientific Epistemological Beliefs, Learn- ing Environment Preferences and Cognitive Structure Out- comes, Doctoral dissertation, Teachers College, Columbia University.

Tsai, C-C (1998). The constructivist epistemology: The interplay between the philosophy of science and students' science learn- ing. Curriculum and Teaching 13(1).

Tsaparlis, G. (1997). Atomic and molecular structure in chemi- cal 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). Molecular- equilibrium problems: Manipulation of logical structure and of M-demand, and their effect on students' performance. Sci- ence Education 82: 437-454.

Tyson, L., Treagust, D. E, and Bucat, R. B. (1999). The complex- ity of teaching and learning chemical equilibrium. Journal of Chemical Equilibrium 76: 554-558.

Tytler, R. (2000). A comparison of year 1 and year 6 students' con- ceptions of evaporation and condensation: Dimension of con- ceptual progression. International Journal of Science Educa- tion 22: 447-467.

Valanides, N. (2000). Primary student teachers' understanding of the particulate nature of matter and its transformations dur- ing dissolving. Chemistry Education: Research and Practice in Europe 1: 249-262.

Van Driel, J. H. (2002). Students' corpuscular conceptions the con- text of chemical equilibrium and chemical kinetics. Chem- istry Education: Research and Practice in Europe 3: 201- 213.

Von Glasersfeld, E. (1992). A constructivist view of teaching and learning. In Duit, R., Goldberg, E, and Niedderer, H. (Eds.), Research in Physics Learning: Theoretical Issues and Empirical Studies, IPN, Kiel, Germany, pp. 29-39.

Voska, K. W, and Heikkinen, H. W. (2000). Identification and anal- ysis of students' conceptions used to solve chemical equilib- rium problems. Journal of Research in Science Teaching 37: 160-176.

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. 177-210.

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions

Page 14: Some Student Misconceptions in Chemistry: A Literature …science+is... · identification of the students' understandings and mis- ... Some of the conceptual areas in which most studies

Misconceptions in Chemical Bonding 159

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: 649-660.

Wheeler, A. E., and Kass, H. (1978). Student misconcep- tions in chemical equilibrium. Science Education 62: 223- 232.

White, R., and Gunstone, R. (1992). Probing Understanding, Graphicraft, Hong Kong.

Tip, 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: 755-770.

This content downloaded from 163.1.62.81 on Sun, 20 Apr 2014 08:50:25 AMAll use subject to JSTOR Terms and Conditions