You are on page 1of 45

Teaching and Learning the Concept of chemical bonding

Tami Levy Nahum1, Rachel Mamlok-Naaman1, Avi Hofstein1 and Keith Taber2*

1. Department of Science Teaching, The Weizmann Institute of Science, Rehovot,


Israel

2. Faculty of Education, University of Cambridge, England

1. Dr. Tami Levy Nahum

Postal address: Department of Science Teaching, Weizmann Institute of Science,


Rehovoth 76100, Israel

E-mail: Tami.Levy@Weizmann.ac.il

Tel number: 972-8-9342497

Fax number: 972-8-9344174

2. Dr. Rachel Mamlok-Naaman

Postal address: Department of Science Teaching, Weizmann Institute of Science,


Rehovoth 76100, Israel

E-mail: Rachel.Mamlok@Weizmann.ac.il

Tel number: 972-8-93422446

Fax number: 972-8-9344115

3. Prof. Avi Hofstein

Postal address: Department of Science Teaching, Weizmann Institute of Science,


Rehovoth 76100, Israel

E-mail: Avi.Hofstein@Weizmann.ac.il

Tel number: 972-8-9343811

Fax number: 972-8-9344115

4. Dr. Keith Taber * Corresponding author

Postal address: Science Education Centre, University of Cambridge Faculty of


Education, 184 Hills Road, Cambridge CB2 8PQ, United Kingdom

E-mail: kst24@cam.ac.uk

Tel number: +44 (0) 1223 330569

Fax number: +44 (0) 1223 767602

1
Word count (including reference list etc): 13 565

Author biographical notes:

Dr. Tami Levy Nahum is Senior Intern in the Department of Science Teaching, at
the Weizmann Institute of Science in Israel; Her research interests include inquiry-
based laboratories, students' misconceptions and pseudo-conceptions, alternative
assessment methodologies, teaching and learning chemical bonding, and students'
evaluative thinking capabilities in the Israeli multicultural context. She was involved
in the development of a new chemistry curriculum and in scores of teachers'
professional developments courses. Currently, she is also involved in a National
Project focusing at the professional development of science teachers in Junior High
School.

Dr. Rachel Mamlok-Naaman is Senior Staff scientist and coordinator of the


Chemistry Group at the Department of Science Teaching at the Weizmann Institute of
Science. In addition, head of the National Centre for Chemistry Teachers at the
Weizmann Institute of Science, and a senior member of the Science and Technology-
for-All Group. Engaged in development, implementation and evaluation of new
curricular materials, and research on students' perceptions of chemistry concepts.
Publications in the areas of scientific and technological literacy, teachers' professional
development, cognitive aspects of students' learning, assessment and curriculum
development. In the last three years she has been involved in an educational project in
the framework of the European Union.

Prof. Avi Hofstein is head of the Chemistry Group at the Department of Science
Teaching at the Weizmann Institute of Science in Israel. His scientific activities focus
on all aspects of the curricular process in the context of chemistry education programs
as well as in the program "Science for All". This includes curriculum development,
implementation, professional development of both regular teachers and leadaing
teacher, and research of various aspects related to teaching and learning science
(chemistry). In recent years he was involved in several EU projects on technology
education and continious professional development of science teachers.

Dr. Keith Taber is Senior Lecturer in Science Education in the Faculty of Education
at the University of Cambridge, UK. After completing his first degree in chemistry, he
trained as a teacher of chemistry and physics, and taught in secondary schools and a
further education college. After completing his doctorate on students’ developing
understanding of chemical bonding he moved to Cambridge, initially focusing on
teacher education, but increasingly working with higher degree students. He was the
Royal Society of Chemistry’s Teacher Fellow in 2000-2001. His main research
interests relate to aspects of conceptual learning in science – including issues of
conceptual integration; learning about the nature of science and its relationship with
religion; and challenging the most able (‘gifted’) learners.

2
Teaching and learning the Concept of chemical bonding
Number of words:

Abstract:

Chemical Bonding is one of the key and basic concepts in chemistry. Clearly, many of
the concepts taught in chemistry in both secondary schools as well as in the colleges
are highly based on understating the fundamental ideas related to this concept.
Nevertheless, the concept is perceived both by the teachers as well as by learners as
difficult to teach and as causing many misconceptions regarding the students. Many
of these misconceptions result from over-simplified models used in text books, by the
use of traditional pedagogy that present a rather limited and sometimes incorrect
picture of the issues related to chemical bonding and by assessment of students'
achievement that influence the way the topic is taught. In addition, there are
discrepancies between scientists regarding the definition of the topic and regarding
models to teach it. In this review paper we tried to provide science educators,
curricula developers and pre-service and in-service professional development
providers an up-to-date picture regarding the research and developments of the
chemical bonding. We reviewed the external and internal variables that might cause
these misconceptions and the problematic issue of using limited teaching/learning
models. Finally we review the approaches to teaching the concept that might
overcome some of these misconceptions.

Keywords: chemical bonding, misleading factors, misconceptions, alternative


frameworks, fundamental principles, models

3
Teaching and Learning the Concept of chemical bonding
Tami Levy Nahum1, Rachel Mamlok-Naaman1, Avi Hofstein1 and Keith Taber2

1. Department of Science Teaching, The Weizmann Institute of Science, Rehovot, Israel

2. Faculty of Education, University of Cambridge, England

Introduction

A chemical bond is the physical process responsible for the


interactions between atoms and molecules, and that which confers
stability to bounded particles. The explanation of these
interactions, which are basically electrical forces, is a complex
area that is described by the laws of the quantum mechanics theory
(Wikipedia, 2010).

In school science teaching, ideas need to be presented in ways that are both authentic
representations of the scientific concepts, and yet simple enough to be meaningfully
understood by the learners. Yet chemical bonding is inherently an abstract and
complex concept. When the authors of this review approached prominent scientists,
worldwide, regarding this issue, they got a variety of complicated definitions.

Very often, when considering the teaching of chemical bonding, the arguments are
focused on pedagogical issues rather than scientifically related disagreements.
Needless to say, seeking a core understanding of chemical bonding becomes more
problematic when we try to translate key ideas to the classroom practice. This review
sets out to integrate research-based knowledge regarding the teaching and learning of
chemical bonding.

Bonding: The nature of the concept

There is no doubt that chemical bonding is one of the key concepts in


chemistry and one of the most fundamental ones. It is also one of the areas in the
physical sciences where understanding is developed through diverse models - which
are in turn built upon range of physical principles - and when learners are expected to
interpret a disparate range of symbolic representations standing for chemical bonds
(Taber & Coll, 2002). The concepts associated with chemical bonding and structure,
such as covalent bonds, molecules, ions, giant lattices, and hydrogen bonds are highly

4
abstract and in order to fully understand these concepts, students must be familiar
with mathematical and physical concepts and laws that are associated with the
bonding concept, such as orbital, electro-negativity, electron repulsions, polarity, and
Coulomb's law. In addition, learning about chemical bonding allows the learner to
make predictions, and give explanations, about physical and chemical properties of
substances.

Sánchez Gómez & Martín (2003) have discussed how the most advanced models
available to chemists for understanding the structure of matter, that is those that we
might judge our best approximations to ‘reality’ given the current state of knowledge,
derive from quantum chemistry. However, Sánchez Gómez and Martín also suggest
that the majority of chemists are quite content to work with models that largely pre-
date developments in quantum chemistry, and which have to be considered to be less
well supported by our current best understanding of matter. As this means that most
chemists are using a set of models that are now understood to be limited
representations of the structure of matter, they describe this popular if flawed
conceptual toolkit as the folk molecular theory (FMT). The chemical concepts of
FMT are then retained as useful conceptual tools, and for many chemists continue to
provide a set of explanations of chemical phenomena (that is molar scale phenomena
actually experienced in the laboratory, or in nature) in terms of an understanding of
the submicroscopic composition and structure of matter. Atoms, molecules, ions,
bonds and so forth are therefore available components of explanations that help to
systematize chemical knowledge. In each of the following examples of high
school/college level chemistry explanations, material behaviour that can be directly
investigated is explained in terms of theoretical entities that are part of a conceptual
‘toolkit’:

• copper conducts electricity because it has metallic bonding


with delocalised electrons
• diamond has a high melting temperature because its atoms
are strongly bound into a lattice by covalent bonds
• sodium chloride is soluble in water, but not in benzene,
because the crystal is held together by strong ionic bonds,
but the ions are also able to form bonds with water
molecules when hydrated
• water evaporates readily because it comprises of small
covalent molecules

5
• water expands on freezing because it forms a lattice of
hydrogen bonded molecules
• a stream of water can be deflected by a charged rod because
the molecules have a dipole moment due to the polar bonds

In each of these cases the phenomenon to be explained is something that can be


directly demonstrated on the bench (e.g. copper conducts electricity), and the
explanation refers to entities that are discussed as if real objects (e.g. metallic
bonding, delocalised electrons), but which are theoretical constructs such as types of
bond between atoms.

Understanding particle theory: Quanticles

The notion of chemical bonding is then part of an extensive explanatory framework


that chemists use to make sense of molar scale phenomena in terms of a conjectured
sub-microscopic level of material structure. A key feature of these models is that
matter is not considered to be homogenous and continuous, but rather at a small
enough scale it comprises of myriad components – which are considered to be the
fundamental particles from which macroscopic structures are built.

Particles at the submicroscopic level – atoms, ions, electrons, molecules – are not like
more familiar particles such as salt or sugar grains. These particles of the molecular
world are fuzzy packets of fields without surfaces or definitive volumes, which extend
indefinitely and which can often interpenetrate each other. To use the term ‘particles’
can mislead learners. They are something else – one of us has commonly called them
‘quanticles’ (Taber, 2002b) to emphasise this distinction.

Precisely what level of quanticle is considered has shifted historically. The most
salient notion here is surely that of the atom, which offers a strong link with a
macroscopic notion of fundamentality in matter. That is, all substances and materials
can be understood as derived from a limited number of chemical elements (the
macroscopic level); and each element is composed of (to a first approximation) one
type of atom.

The notions of atoms derives from Democritus, and although the modern chemical
notion of the atom rather different to this ancient notion (modern atoms are not
actually atomos, and are not hard volumes of material with discrete edges), the

6
teaching model of atoms introduced to school students (e.g. ‘everything is made of
tiny particles called atoms’) has been considered to retain something of the flavour of
the ancient idea (Taber, 2003b).

Since chemists understand substances in terms of clusters of sub-microscopic


particles, the chemical bonds between those particles are used to explain many of the
chemical and physical properties of substances and chemical phenomena (Hurst,
2002; Levy Nahum, Hofstein, Mamlok-Naaman & Bar-Dov, 2004). Since bonding is
a central concept in teaching chemistry, a thorough appreciation of its nature and
characteristics is essential for understanding almost every other topic in chemistry,
such as carbon compounds, proteins, polymers, acids and bases, chemical
thermodynamics, proteins, carbohydrates, and polymers (Fensham, 1975; Gillespie,
1997; Hurst, 2002; Levy Nahum et al 2004). Gillespie (1997) in his essay titled: Great
ideas in chemistry, included chemical bonding (i.e., what holds atoms together in
molecules and crystals) as one of the six most important key concept that should be
included in every high school and introductory college chemistry courses. In addition,
the concept is very much related to the understanding of many important and
fundamental biological aspects such as molecular biology (e.g. DNA and RNA – i.e.
nucleic acids). However, based on the literature, bonding is considered by teachers,
students, and chemists to be a very complicated concept (Gabel, 1996; Levy Nahum,
Mamlok-Naaman, Hofstein, & Krajcik, 2007; Robinson, 2003; Taber, 1998, 2001,
2002a; Tsaparlis, 1997).

This review is targeted at 10-12th grade (upper-secondary school) and the


beginning of college and undergraduate. Thus, it is mainly focused on the nature of
the ‘chemical bond’, assuming that students at this stage, have already learned about
the periodic table, the particulate nature of matter, elements, compounds, and other
basic concepts that are usually taught in the lower-secondary school. Therefore these
concepts are not included in the current review which is focused on issues related to
the teaching and learning of the chemical bonding concept.

How is bonding being taught?

The problematic traditional approach for teaching the concept

7
One of the goals of the chemistry teaching community is to develop more effective,
pedagogically and scientifically sound, strategies to teach high-school students the
concept of chemical bonding. According to Teichert and Stacy (2002), this aim is
motivated by many studies conducted worldwide that clearly revealed that the
traditional approach to teaching bonding is problematic, and misconceived. More
specifically, during the last two decades researchers have found that students
commonly lack a deep conceptual understanding of the key ideas regarding the
bonding concept and often fail to integrate their mental models into a coherent
conceptual framework (Bodner & Domin, 1998; Griffiths & Preston, 1992; Herron,
1996; Peterson & Treagust, 1989; Taber, 2001).

The problematic approach through which this concept is presented in many chemistry
textbooks worldwide has been examined extensively in the last two decades by
researchers of chemistry teaching and learning (Ashkenazi & Kosloff, 2006; Atzmon,
1991; Hurst, 2002; Justi & Gilbert, 2002; Taber, 1998). In these textbooks, elements
are conveniently classified as metals or-nonmetals (although sometimes semimetals
are also mentioned). Very often this dichotomy among elements leads to a
dichotomous classification of bonding related to compounds, namely covalent
(existing between non-metallic elements) or ionic (existing between metallic and non-
metallic elements). Hurst (2002) suggested that a common approach used by
chemistry curriculum developers (and also by textbooks writers) is to present four
different groups of substances: the ionic lattice, the metallic lattice, the molecular
lattice, and the covalent lattice. According to Hurst, these divisions and
characterizations are then (wrongly) used to differentiate between substances.
Furthermore, many chemistry textbooks do not refer to the hydrogen and Van-der-
Walls interactions as bonds, they refer to them just as ‘forces’ presumably since they
do not meet the ‘octet rule’ (the problematic aspect of using this ‘rule’ for molecules
is discussed below). This traditional pedagogical approach (Figure 1) which is
commonly used worldwide for studying bonding can be found not only in high-school
textbooks but also in general chemistry textbooks intended for college freshmen
(Hurst, 2002).

___________________________________

Insert Figure 1 at about here

8
___________________________________

Levy Nahum et al (2008) focus on two main problems. Firstly, the presentation of
each bond type as a different entity that belongs to a specific category does not foster
a deeper understanding of chemical bonds. Specifically, it may obscure the important
notion of a unified rationalization of all chemical bonds based on underlying
principles. Secondly, over-emphasis of the four ‘ideal’ bonding categories (mentioned
above) is misleading and may actually hinder the learning process. While the ‘ideal’
classification is not without merit, it is now known (Naaman & Kronik, 2004) that
many important groups of modern materials simply cannot be ‘forced’ into one of the
rigid categories.

In the following section, several limitations of the rigid traditional approach,


which were mentioned above are illustrated, with specific examples from literature.

(1) Covalent versus ionic bonding. Typically, covalent and ionic bonds are
presented dichotomously, as ‘electron sharing’ or ‘electron transferring’ bonds,
respectively. However, in hetero-atomic systems, bonding is better described in terms
of a continuum of a covalent-ionic dimension or ‘scale’ (Levy Nahum et al, 2007;
Sproul, 2005; Taber & Coll, 2002). Furthermore, bonds between two identical atoms
are purely covalent, but purely ionic bonds actually do not exist at all. The
dichotomous presentation impedes the understanding of the more subtle ‘scale’.

(2) Electronegativity and bond polarity. Because within the traditional approach
bond polarity is essentially viewed as an additional characteristic of covalent bonds
(Hurst, 2002; Taber & Coll, 2002), the important concept of electronegativity (EN) is
only introduced in the context of covalent bonding and not as an integral part of bond-
polarity concepts (Weinhold & Landis, 2005). EN differences between atoms are then
used as an indication of whether compounds should be classified as ionic or covalent.
However, EN differences are not the ultimate measure for predicting bond type
(Sproul, 2005). Indeed, cases of bonds between atoms with large EN differences that
nevertheless possess a significant covalent nature are known experimentally (Woicik
et al, 2005).

9
(3) The octet ‘rule’. Because it is simple for the learners to visualize and use, the
octet ‘rule’ is often presented as an obligatory condition for ‘proper"’ bonding. Thus,
students often adopt the anthropomorphic notion of atoms ‘wanting’ to possess
‘octets’ or ‘full outer shells’ and consider that chemical reactions occur in order to
‘allow" atoms to achieve this ‘natural desire’ (Taber, 1998; Taber & Watts, 1996).
But this causes some students to have difficulties in accepting anything that is not
clearly explicable in ‘octet’ terms, e.g., hydrogen bonds or even covalent bonds or
transition metal bonding not leading to ‘octets’ (Weinhold & Landis, 2005). It is now
known that many molecules and complexes do not ‘obey’ this rule (e.g. the ‘1998
Nobel Prize molecule’ NO is a good example; Culotta & Koshland, (1992). The octet
rule is certainly a time-honored useful guideline and shall remain so. However, it is
not a valid explanation for bond formation. For these reasons, Taber (2001a), Hurst
(2002) and Taber and Coll (2002) suggested that an over-emphasized ‘octet
framework’ may actually impede higher-level learning process.

(4) Metallic bonds. The traditional discussion of metals and metallic bonds may
involve both an over-generalization and an over-simplification. Over-generalization
occurs because in many general chemistry textbooks metals are characterized by a set
of common physical and chemical properties e.g., malleability, ductility, low
ionization potential, etc (Ben-Zvi, Eylon, & Silberstein, 1986). But the fact is that
there is not even one chemical property common to all metals, and there is a great
variability of parameters in any other quantity, e.g., brittleness, conductivity, boiling
point, etc. Over-simplification occurs because many textbooks introduce a metallic
bond as ‘metal ions floating in a sea of electrons’. This analogy is problematic
because it presents the metallic bond as a bonding entity that is entirely different from
the covalent one, whereas a more modern description views both types of bonds as
involving ‘electron sharing’. The difference is again explained in terms of a
continuum scale, this time involving the degree of electron delocalization (Myers,
1979).

(5) Hydrogen and van der Waals bonds. In most textbooks, covalent and ionic
bonds are described as ‘real’ chemical bonds, whereas hydrogen and van der Waals
bonds are often presented as ‘just forces’ (Taber, 1998). Again this distinction is far
too rigid and misleading. While the relative strengths of different types of bonds are,

10
of course, very important, even ‘weak’ bonds do indeed bond together different
chemical units and sub-units and can have profound chemical consequences (Levy
Nahum et al, 2007; Taber & Coll, 2002), e.g., hydrogen bonding in biochemistry (for
example in the double helix structure of DNA). Therefore, a continuum scale is a
more appropriate scientific description. Two related over-simplifications are the
classification of hydrogen bonds as strictly inter-molecular, whereas they are often
intra-molecular as well (in proteins for instance), and the discussion of such bonds
only when N, O, or F atoms are involved, whereas hydrogen bonds, albeit weaker or
non-conventional, may occur with other atoms or groups as well (e.g., Naaman &
Vager, 1999; Nakanaga & Buchhold, 2002).

Despite this, the traditional general chemistry curriculum as a whole has by and
large been taken for granted by science educators for over a century and only little
was done to develop valid teaching models and pedagogical remedies to overcome
these problematic approaches (de Vos & Pilot, 2001).

In the following sections these problematic issues are being reviewed and elaborated
with specific research data as examples.

Chemical Bond: Learning difficulties, misconceptions, and alternative


conceptions

The nature of chemical bonds

The curricular models of bonds presented in school science are based on the fact that
there are physical forces between different quanticles. At school level these forces are
normally understood as being electrical in nature.

Wightman, Green and Scott (1986) described a UK lower secondary class learning
about the particle model of matter. They found that the children considered the
linkage between particles in material terms: to be like elastic, or based on shapes
interlocking. When the interactions between particles were compared to a magnetic
attraction (i.e. a type of force familiar to the pupils), pupils’ conceptions tended to be
vague and to retain material aspects: e.g. there was “like string between the atoms sort
of holding it all together” (p.291). Part of the children’s difficulties seemed to be
related to finding a model of bonding which allowed melting and vaporization, yet

11
still enabled bonds to form when particles were in close contact. As one of the
students commented, “the point is, how do we get the bonding back?” (p.292). It was
suggested that when the particles slowed down the bonding was then able to “get to
grips” with the particles as it is “a bit easier to keep slower things together” (p.292).
In this context, it has been suggested by Papageorgiou and Johnson (2005) that in
introducing the particle model of matter the notion of bonding might be introduced
“through the idea of the particles having an ‘ability to hold on to each other’, rather
than the more commonly employed talk of attraction”.

Griffiths and Preston (1992) asked grade-12 students (16-18 years) in Canada to
sketch molecules of ice. They report that typical diagrams showed the molecules
touching each other without spaces between, and they conjectured for these learners
the concept of bonding might have little to do with forces of attraction. For some
students in their sample, molecules were not bound due to inherent interactions
between atoms, but were held together by “something external to the molecules”.

In a French study of first year undergraduate science students Cros and colleagues
(1986) carried out 40 unstructured interviews; 50 semi-structured interviews;
followed by a survey of 400 students at two Universities before starting their lecture
courses. They found that the interactions between atoms in molecules were often
unknown or poorly known and that often students were not even aware that such
interactions existed:

A study conducted in the US by Nicoll (2001) was a first attempt to elucidate and
detail the types of misconceptions relating to bonding, electronegativity, electrons,
and molecular geometry that undergraduate chemistry majors hold. Nicoll reported
that quite a few students tried to explain chemical bonding in terms of electrons
attracting one another. She reported a category of incorrect explanations for bonding
phenomena or incorrect explanations for why bonding occurs and quoted one student
who explained “When you have two electrons, they’re negatively charged ions. They
don’t want to come together. But sometimes they do, and that’s a chemical bond”
(p.717).

12
A common alternative conceptual framework for chemical bonding

Taber (1998) undertook a study of students’ thinking about chemical bonding and
related concepts among UK students. The core of the research was an interview-based
study to explore developing understanding of chemical bonding among students
studying chemistry at A level in a further education college (i.e. 16-19 year old
students). Whilst the in-depth nature of the study demonstrated diversity in student
thinking, Taber identified common tendencies and conceptions that were reported as
an alternative conceptual framework – that is a model representing the common
features found across a range of students. Taber characterised students as entering
study at this level holding a notion of chemical bonding deriving from school study,
and tied to the octet rule. For these students bonds would form because atoms wanted
to obtain full electron shells. They would do this by sharing electrons (covalent
bonding) or through an electron transfer (ionic bonding). It was found that students
tended to retain and use the ‘octet rule explanatory principle’ even after they had
studied bond energies and thermodynamic considerations on more advanced courses.
So students would commonly explain the reaction between fluorine and hydrogen in
terms of the atoms wanting to obtain full shells even in response to a probe question
that presented the reaction equation showing that the reactants were already
molecular, and so satisfying the octet rule (Taber, 2002a).

The anthropomorphic nature of many student explanations of bonding, based upon


what atoms 'wanted' or 'needed' to do, seemed to be a key feature for these
explanations seeming authentic to the students themselves (Taber & Watts, 1996).
Taber (2000, 2001b) argued that progression in student thinking during the A level
course could be seen as a shift from understanding bonding in terms of the octet rule
framework to adopting electrostatic and orbital ideas. A detailed study of one
student’s progress demonstrated that this could be a slow process that might not be
complete during the two-year college course. The model students brought from school
level study appeared to be tenacious. In part this seemed to be because the octet rule
explanatory principle was not an isolated conception, but part of a well integrated
framework of ideas.

Taber suggested that in the absence of a viable rationale for bond formation being
offered in school courses, the octet rule was adopted and extended to fill the

13
‘explanatory vacuum’. It was not clear to what extent this was encouraged by teachers
– that is whether in the absence of a suitable curriculum model to explain bonding
teachers’ own explanations suggested the octet rule offered a rationale for bond
formation. However, in a study aimed at exploring New Zealand students’ mental
model of bonding, Coll and Treagust (2003) undertook an examination of curriculum
material and interviews with instructors. They identified “two target models” for ionic
bonding, one of which “termed the electrostatic model is based on the octet rule;
another model, termed the theoretical electrostatic model, is based on the computation
of the alternating attractive and repulsive forces within an ordered ionic lattice”. In
that context they found that “the former model is introduced at secondary school; the
latter model is introduced in the second year of the undergraduate program”. That is,
for school level learners, ionic bonding was taught in terms of a curriculum model
based upon the octet rule.

The covalent bond.

From a test instrument to diagnose senior secondary students' understanding of the


topic covalent bonding and structure, Peterson et al (1986) identified common
misconceptions as those selected by at least 20% of their sample . They listed the 13
misconceptions identified under 6 categories in Table 1.

___________________________________

Insert Table 1 at about here

___________________________________

Some of these misconceptions could be seen as partial understanding or application of


curricular models, as when considering molecular shape to be determined by only the
bonds or only by the non-bonding electron pairs; or when considering the polarity of a
molecule to depend upon bond polarity without consideration of the arrangement of
bonds. The authors acknowledged that the confusion between intermolecular and
intramolecular interactions “may be a case of mistaken terminology rather than a
conceptual misunderstanding” (p.311).

This type of instrument used, offers limited insight into student thinking. A common
feature of work in this area is the necessary trade-off between large samples offering

14
indications representative of student populations, and in-depth studies that can explore
the nuances of students’ reasoning (Taber, 2009a). However, we would identify three
aspects of the study worthy of further note:

• Firstly, although the research concerned ‘understanding of the topic covalent


bonding and structure’ the authors clearly considered that polar bonds fell
within this topic. That is, for the purposes of the study, polar bonds are
treated as a type of covalent bond (a point to which we return later).

• Secondly, one of the misconceptions identified (‘nitrogen atoms can share 5


electron pairs in bonding’, see Table 1) relates to the failure to apply the octet
rule. This misconception is worthy of note, in light of the work reviewed
above that suggests that the octet rule often takes on particular (and
unwarranted) significance in student thinking about bonding. Overall then,
Peterson and colleagues suggest that knowledge of the octet rule increases
from grade-11 to grade-12, but that during grade-12 a new misconception
appears. Although the authors offer no explanation, a feasible possibility is
that during grade-12 students meet compounds of period 3 where elements
such as phosphorus ‘expand their octets’, and they simply apply this to
nitrogen as well.

• Finally, the authors draw attention to the notion of a covalent bond as the
sharing of a pair of electrons, and acknowledge that this linguistic formulation
may be interpreted differently by students than intended by teachers.

Coll and Treagust (2001) undertook an in-depth interviews with six Australian
learners found that these “learners view covalent bonding as the sharing of electrons,
with the secondary school and undergraduate learners relating this specifically to the
octet rule of full shell stability”. They reported how the two postgraduate students
they interviewed tended to supplement their octet rule based model, that “covalent
bonding results from the sharing of electrons” with elements from other models (e.g.
orbital, electron clouds). They reported that all the learners struggled to apply their
model of covalent bonding to explain chemical phenomena.

15
The ionic bond.

Levy Nahum et al (2004) explored misconceptions among Israeli students, based on


their responses in the national matriculation examinations (final examinations set by
the government at the end of the 12th grade) over a period of twelve-years. One of the
common patterns they revealed was that in general students confuse ionic compounds
and molecular compounds. Similar pattern was also found by Butts and Smith (1987)
who undertook research to follow up a survey finding that the difference in properties
between ionic compounds and molecular compounds had been rated as a difficult
topic by 29% of students who were asked. Butts and Smith reported that there were
some who believed that sodium ions and chloride ions were released only when the
solid dissolved in water. In addition, they reported that the understanding of a
common form of a model used to illustrate NaCl structure in chemistry teaching (a
‘ball and stick’ model):

…was confused, many interpreting the six wires attached to each


ball (‘ion‘) as each representing a bend of some sort (’there is one
ionic bond and five physical bonds’ and ‘I would have expected
seven wires not six, because chlorine has seven electrons in its outer
shell’). Very few were able to explicitly state that the wires did not
represent bonds but merely held the wooden balls in place (p.196).

It should be noted that ionic bonding can be considered a description of the overall
interactions between the various ions in an ionic crystal; however, it is also common
to consider that in the NaCl lattice, each ion forms six ionic bonds with surrounding
counter-ions.

In a UK interview study, Taber (1994) found that 16-19 year old students
demonstrated ideas about ionic bonding similar to those found by Butts and Smith.
Taber proposed an alternative conceptual framework for making sense of ionic
bonding (the ‘molecular’ framework), that as well as including molecular or pseudo-
molecular entities, encompassed three particular alternative conceptions (Taber,
1994).

• that an ionic bond only existed where there had been an


electron transfer between atoms to form ions;
• that electrovalency (charge) limited the number of ionic
bonds formed;

16
• that the ionic lattice comprised of ions that were ionically
bonded to some counter-ions, and attracted to other by ‘just
forces’.

Based on the above, regarding NaCl, a student would consider a Na+ ion to be bonded
to a single Cl- ion by an ionic bond, and then attracted to another 5 because of the
electrical forces between oppositely charged ions. That is, the ionic bond was
considered to be something more than the electrical attraction, and this was because it
was associated in the student’s mind with a conjectured electron transfer event during
which ions were formed to allow the atoms to obtain octet structures/full outer shells
(that is, the molecular framework for thinking about ionic bonding was subsumed
within the broader octet framework for understanding bonding phenomena). For the
students the focus of ionic bonding is the electron transfer event. Indeed students
asked about precipitation reactions have been found to explain ionic charge in a
precipitate as due to electron transfer between the species, even after acknowledging
those same ions were already present in the reagent solutions (Taber, 2002a).

A survey of students near the end of school science (14-16 year olds) and undertaking
A level study (16-19 year olds) suggested that students were commonly supportive of
statements reflecting the molecular framework, but often also agreed with statements
reflecting the curriculum model (i.e. that all adjacent oppositely charged ions are
ionically bonded together). Taber suggested that students often operated with a
mixture of the alternative conceptual framework and the curriculum model (Taber,
1997).

The survey instrument has recently been used in translation to survey samples of
undergraduate students in Greece and Turkey, and it was found that students in these
countries also commonly agree with statements based on the alternative conceptions
of ionic bonding elicited from UK students (Nakiboğlu, Tsaparlis & Taber, 2009).

Coll and Treagust (2003) reported on a study where students from New Zealand at
three educational levels - secondary school, undergraduate and graduate - were
interviewed about their mental models of bonding. They reported that “the secondary
school learners identified ionic bonding as an attraction between charged species” but
“related their mental models directly to the octet rule of full-shell stability” (p. 472).
So although the bond was understood in terms of “attraction between oppositely

17
charged species”, again, the octet rule was seen as the “sole driving force” for
bonding, leading to a transfer of electrons as an “automatic consequence” (p.478).
Coll and Treagust reported that the undergraduate students placed greater emphasis on
lattice structure and saw the attraction between oppositely charged species as the
principal driving force for formation of ionic bonding. However, even the graduates
related the formation of ions to the octet rule, and they quoted one of these graduate
students describing in anthropomorphic terms how “the sodium prefers to have a one
plus and the chlorine prefers to have a one minus” (p.474). At this level, Coll and
Treagust found that the octet rule was “routinely mentioned, but not seen as principal
driving force” (p.478).

The metallic bond

Taber reported that students commencing their A level studies in the UK generally
had limited knowledge of metallic bonding (Taber, 1997b, 2003a). He suggested that
their school level learning had focused on covalent bonding (as electron sharing) and
ionic bonding (in terms of electron transfer), and metallic bonding seemed to have
been given little attention. He found that students adopted the model of metallic
bonding in terms of a ‘sea of electrons’, but this seemed to often be learnt as a
‘slogan’ without deep understanding (student diagrams suggesting that the ‘sea’
metaphor implied a vast excess of electrons despite the lack of charge conservation
this implied).

For these students a chemical bond was understood to be a means of allowing atoms
to fill their shells, and students interpreted metallic bonding in terms that fitted this
framework: either considering metallic bonds not to be ‘proper’ bonds but just forces,
or as some variation on, or hybrid of, the ionic and covalent cases. The delocalised
electrons could be seen as shared around to fill electron shells, or as donated to the
lattice to allow ion formation.

Acar and Tarhan (2008) undertook a comparative study of learning about metallic
bonding in two classes of 15 year old students in Turkey. One class was taught by
traditional teacher and text-book led instruction; the other by cooperative group-work
guided by enquiry questions and supported by a teacher said to be acting as a
‘facilitator’. Distracters on the multiple choice post-test that referred to the bonding in

18
metals as being ionic, as being a type of ionic bonding, or looking like ionic bonding
were selected by fair proportions of the control group (approximately two-fifths, half,
and three-fifths respectively). Almost a third of this group selected a statement
referring to the interactions as ‘weak forces’. It should be noted that the students in
Acar and Tarhan’s research studied ionic and covalent bonding prior to being
introduced to metallic bonding. In this regard, Taber (2001a) has argued that students
will tend to interpret new classes of bonding in terms of previously learnt ones. Acar
and Tarhan report interview comments that support their view that a large proportion
of the students confused metallic bonding and ionic bonding in this study. These
comments suggest that it may be difficult to distinguish between students confusing
ionic and metallic bonding, and simply making sense of the new family of bonding in
terms of the more familiar category.

Polar bonding

It is clearly agreed by scientists that most bonds are polar, in the sense that the
electron density is neither equally shared (as in the covalent model) nor considered to
reside entirely around one atomic core (as in the ionic model).

Taber (1998) found in his interview study that when students learnt about polar bonds
they tended to consider them in terms of their existing model of covalent (sharing
between two atoms of non metallic elements) and ionic bonding (electron transfer
from atom of metallic element to atom of non-metallic element). Learning about
degrees of electronegativity should break down the metal/non-metal dichotomy of
element type that provides the basis for the covalent/ionic dichotomy of bond type.
However, as students understood the covalent and ionic bond types in terms of
distinct ways to obtain full outer shells, they did not readily adopt polar bonding
(which could not readily be explained in terms of the octet explanatory principle) as a
primary category of bond. Instead they tended to see polarisation as a secondary effect
imposed upon the more familiar pattern of covalent and ionic bonds. Indeed, although
it is possible to conceptualise polar bonds as distortions in either covalent or ionic
bonding, Taber found that his student generally saw polar bonds as examples of
covalent bonds where sharing was uneven. Although they learnt explanations of bond
polarity in terms of electrical interactions, this was superimposed upon their well-

19
established notion of a covalent bond as being the sharing of electrons to allow atoms
to have full shells.

Intermolecular bonding.

Most of the research studies that have been conducted regarding inter- and intra-
molecular forces clearly showed that in general students tend to confuse these two
categories. In his research Taber (1988) operated from a more inclusive notion of the
chemical bond, such that hydrogen bonds, van der Waals’ forces, solvation
interactions etc were all considered to be types of chemical bonding. He accordingly
considered student thinking deficient when they excluded these types of interactions
from their notion of the chemical bond. Once again, Taber found that students’
school-level learning about bonding in terms of ways of allowing atoms to fill their
outer electron shells acted as a barrier to accepting forms of bonding that could not be
explained in these terms. Taber characterized students as seeing interactions such as
van der Waals’ forces as not being ‘proper’ bonds.

Henderleiter, Smart, Anderson and Elian (2001) interviewed 22 undergraduate


students in a US public university who had completed the second organic chemistry
course of a two-semester sequence. They reported how “four students confused
hydrogen bonding with a covalent bond that exists between hydrogen and some other
atom”; five students thought that “hydrogen bonds can be induced”; and that five
students “confused intramolecular hydrogen bonding with a chemical reaction” and
thought that “intramolecular hydrogen bonding resulting in formation of new covalent
bonds”. They commented that “these results illustrate that some students completing
what is typically their second year of college-level chemistry still possess
misconceptions found in younger, less experienced students. Similar findings were
found by Levy Nahum et al, (2004), who reported that students in Israel confused
inter-molecular and intra-molecular bonding.

External sources that might cause learning difficulties and misconceptions


regarding learning the chemical bonding concept.

According to Gilbert (1982) and to Kind (2005), a key perspective will be that it is
often less useful to think about the problems mentioned above in terms of curriculum
science (the representations of scientific knowledge prescribed in the curriculum) as

20
being ‘correct’ or ‘true’, and alternative ideas presented by students as simply being
‘incorrect’ or ‘false’. Rather, when considering understanding of chemical bonding, it
is important to recognize that curriculum science comprises a set of models which are
intended to provide an authentic representation of the models used by scientists, at a
level accessible to learners (Gilbert, 2004). It is suggested that this complicates
judgments of the correctness of students’ conceptions, and that both the nature of the
scientific models, and the way the topic is taught, contribute to student learning
difficulties in the topic.

There are several reasons for dissatisfaction among the chemistry teaching community
regarding the current teaching and learning of this concept; we will refer to two main
components based on the literature: (1) the traditional pedagogical approach, as it
appears mainly in chemistry textbooks worldwide, and (2) the assessment methods
used worldwide (high stakes testing). For example, the Israeli Matriculation
Examination (ME) influences teachers' instruction and students' learning regarding the
bonding concept, as the teachers' main objective is in preparing their students for the
ME questions and answers, leading to superficial teaching which results in
misconceptions and pseudo-conceptions (Levy Nahum et al, 2004, 2007). This is
shown in Figure 2.

___________________________________

Insert Figure 2 at about here

___________________________________

Gilbert, Osborne and Fensham (1982) pointed out the importance of avoiding
confusion between what they called ‘children’s science’ and ‘scientist’s science’, or
for that matter ‘teacher’s science’. Certainly in terms of school science teaching, it
would often be quite inappropriate to consider current ‘state of the art’ scientific
models as suitable target knowledge to inform the planning of teaching and the
evaluation of student learning. Rather there is at least a two-stage process of
transformation between ‘scientific knowledge’ and the learning material set out for
students to learn in classes.

The first stage of transformation generally occurs at a system-wide level, beyond


particular schools or classrooms – at least in educational systems where there is a

21
prescribed curriculum, or a set of reference ‘standards’. In this process the models of
science inform the development of curricula models: representations of scientific
knowledge considered to be suitable for teaching and learning for a particular age
range. These curricula models are intended to be authentic representations of the
science, but at a suitable level of simplification. This process of curricula model
development is commonly undertaken by committees, which are often likely to be
dominated by educationalists who are not themselves the scientists who develop and
apply the scientific models in their professional work. There is inevitably a process of
interpretation therefore which may lead to aspects of the scientific models being
misconceiving and distorted.

An example of the type of argument that informs the development of curriculum


science is provided by Gillespie (1997) suggesting treatment of chemical bonding in
introductory general chemistry courses (at undergraduate level):

All chemical bonds are formed by electrostatic attractions between


positively charged cores and negatively charged valence electrons.
Electrostatic forces are the only important force in chemistry.
Bonds are not formed by the overlap of orbitals, as we read not
infrequently; this is just a model—admittedly a very useful one and
essential for the chemistry major, but I don’t think it is essential
for students at the introductory level. We can obtain a very good
understanding of chemistry without it; indeed, many chemists
make little use of it. It distracts attention from the real reason for
bond formation: the electrostatic attraction between electrons and
nuclei. … We can simply describe ionic bonds as resulting from
the electrostatic attraction between ions, and covalent bonds as
resulting from a shared pair of electrons’ attraction for the two
atomic cores. The corresponding Lewis structures tell us how
many bonds an atom will form. In my opinion, these concepts are
all we need to discuss chemical bonding at the introductory level
(p.862).

In this passage Gillespie rightly labels the notion of bonds as overlap of orbitals as
‘just a model’. However, the notion of bonds as ‘the electrostatic attraction between
electrons and nuclei’ is presented as thought this description is something more than a
model - the way things really are – where it might be better to consider this also as a
model. Whilst Gillespie’s argument about overcomplicating chemistry for novices has
merits, it is important to realize that it is not possible to explain all we want students
to understand purely in electrostatic terms. So to “simply describe …covalent bonds
as resulting from a shared pair of electrons’ attraction for the two atomic cores” asks

22
students to simply accept that negatively charged electrons can often be considered to
occur in pairs (something a purely electrostatic approach would suggest is unlikely),
and the adoption of Lewis models relies upon the heuristic of the octet rule, unless
quantum considerations are acknowledged to offer a theoretical basis for why
electrons appear to be found in somewhat discrete ‘shells’ of specific capacity (2, 8,
18 etc). This issue is revisited below.

Coll and Treagust (2001) reported an in-depth study of six Australian students’
preferred mental models of bonding, based upon the target knowledge according to
their “examination of curriculum material and interviews with instructors” (p.360).
They reported across three levels of learning (secondary school, undergraduate and
graduate). They wrote:

This inquiry resulted in the identification of what we have termed target models for
each of the target systems… For metallic bonding there are two target models – the
sea of electrons model and the band theory… For ionic bonding there are two
models – a model based in part on the octet rule of full shell stability termed the
electrostatic model, and a theoretical model based on the calculation of the forces
present in an ionic lattice, termed the theoretical electrostatic model… For
covalent bonding, a total of four models were identified – the octet rule, the
valence bond approach, the molecular orbital theory, and the ligand field theory
(pp.360-361).

Pedagogical content knowledge (PCK), since its inception as teacher-specific


professional knowledge, has been researched extensively (Kind, 2009). As shown in
Figure 2, the combination of the traditional pedagogical approach of curriculum
developers worldwide and the demands imposed by the common ME questions
generated a growing body of PCK with regard to this concept, which is overly
simplistic; it is not aligned with the up-to-date scientific knowledge and fails in
developing conceptual understanding. According to Magnusson, Krajcik and Borko
(1999):

PCK is a teacher's understanding of how to help students


understand specific subject matter. It includes knowledge of
how particular subject matter topics, problems, and issues
can be organized, represented, and adapted to the diverse
interests and abilities of learners, and then presented for
instruction (p. 96).

The traditional PCK regarding the bonding concept (also termed content-specific
pedagogical knowledge, by Magnusson et al. (1999) has been developed during the

23
last decades within the high-school chemistry education community. This content-
specific pedagogical knowledge guides many teachers in their classrooms. According
to Taber (2005), one of the professional capabilities of the teacher is to find ways to
make complex ideas seem accessible, but this must be balanced by the need to present
material in a way that is scientifically valid, and provides a suitable platform for
future learning. In other words, the teacher needs to find the "optimal level of
simplification": simplifying sufficiently to suit the learners’ present purposes, but not
oversimplifying to undermine their future needs. Over the years, the traditional
pedagogical approach became increasingly simplistic and adopted clear-cut
definitions in order to facilitate students' learning. Unfortunately, the consequent
superficial teaching results in meaningless learning (see Figure 2); students often do
not understand these concepts and this is reflected in their misconceptions and in their
pseudo-conceptions (Vinner, 1997). In discussing the teaching of the chemical
bonding topic in Israel, Levy Nahum et al., (2004) wrote that:

In Israel, chemistry is taught and learned based on the same syllabus and
curricular materials. The syllabus and the materials are centrally developed by the
Ministry of Education. Thus, all the students study the same curriculum and
teachers teach using the same books.

Moreover, they argue that:

the way that this topic is taught is highly based on the definitions of the key
concepts. Based on the syllabus and the textbooks, and according to the
Matriculation Examinations demands, teachers tend to define these concepts in
certain ways and to use “rules” and classifications that are not in alignment with
the scientific theoretical models (p.315).

The traditional approach is characterized by clear-cut definitions and rigid


distinctions and is an insufficient basis for rationalization of current chemical
knowledge. This problem might be amplified by traditional assessment methods, in
which the superficial study of classification and ‘rules’ by rote is rewarding to both
students and teachers because it allows for an efficient evaluation process based on
clear-cut answers to well-defined questions (Levy Nahum et al., 2004; Levy Nahum et
al., 2007).

24
New directions for teaching the chemical bonding concept

A new bottom-up framework

Based on a long-term collaboration between prominent scientists, researchers in


chemistry education, and expert teachers, an innovative program aimed at teaching
the chemical bonding concept, which follows a holistic approach to curriculum (Levy
Nahum et al., 2007), was developed and implemented in 11th-grade chemistry classes
in Israel. Its general approach relies on basic concepts such as columbic forces and
energy at the atomic level to build a coherent and consistent perspective for dealing
with all types of chemical bonds.

As described by Levy Nahum et al. (2008, p. 1680): “It is possible to show how
this diversity [of bond types] arises from a small number of fundamental principles
instead of presenting it as a large number of disparate concepts”. The framework
proposed by Levy Nahum et. al (2008) (see Figure 3) introduces the elemental
principles of an isolated atom (stage 1); it then follows by discussions of general
principles of chemical bonding between two atoms (stage 2), and the general
principles are then used to present the different traditional categories of chemical
bonding as extreme cases of various continuum scales (stage 3). Equipped with this
knowledge, students can then construct a coherent understanding of different
molecular structures (stage 4) and properties (stage 5).

__________________________________

Insert Figure 3 at about here

___________________________________

The primary purpose of the first stage is to provide a qualitative description that is
conceptually consistent with quantum mechanics but gives a very clear, intuitive
answer to the question which puzzles many students, ‘what really causes atoms to
interact and form a chemical bond?’

In order to provide evidence to students that there is nothing ‘mysterious’ about


chemical bond formation, this stage begins by introducing the concepts of energy and
force and the interrelation between them. The understanding that nuclei are held
together because of nucleus-electron attraction, which is a simple consequence of

25
Coulomb's law, is the first step towards a rational view of chemistry which is not
based on rules of thumb, anthropomorphic concepts, etc. A crucial concept is that
stability, in general, is obtained by minimizing energy. The above principles are best
explained by considering the energy curve for any two isolated atoms (figure 4).

Once this is understood, all chemical bonds, of any type, can be rationalized in
terms of energy stabilization (i.e., bond energy) and all equilibrium inter-atomic
distances (i.e., bond lengths) reflect positions where there is no net force on the
nuclei, i.e., attraction balances repulsion.

___________________________________

Insert Figure 4 at about here

___________________________________

Importantly, Figure. 4 is general. It describes the relation between energy and


inter-nuclear distance for the H2 dimer as well as it describes the Li2 dimer, the NaCl
hetero-dimer, or even the He2 dimer! Obviously, there is very much that separates H2
and He2. One of the key goals of the proposed framework is to emphasize that a
continuum scale exists between extreme cases of qualitatively different bonding
scenarios. Having understood, in stage 2, the common denominator of all bonds, some
distinct bonding categories can now be rationalized, as shown in Fig. 5. The example
of a chemical bond in a diatomic molecule may be the most instructive because of the
focus on one bonding entity only. In this context, the concept of electronegativity can
be introduced, naturally, as one way of quantifying the covalence/ionicity balance.
This continuum follows Pauling (1967), who recognized that bonds between unlike
atoms typically have greater bond energy than that of the average of the
corresponding homo-atomic bonds. However, bond strength is not only a function of
the degree of ionicity but also a function of atomic size and other factors.

__________________________________

Insert Figure 5 at about here

___________________________________

Similar arguments are also given and discussed in relation to hydrogen and van-
der Walls interactions (for more details see Levy Nahum et al, 2008). For instance,

26
the He2 dimer - a large-bond-length, very weakly bound molecule – can be used for
introducing the concept of induced-dipole induced-dipole interactions, which can be
rationalized as yet another manifestation of Coulomb’ law.

In the preliminary results of her research, Levy Nahum (2007) found several
practical advantages of the new framework. Generally, the proposed new framework
overcomes a major difficulty in the traditional ‘top-down’ approach by removing the
artificial division between different types of bonding. Instead, a variety of bonds are
introduced to the students from a continuum point of view. Furthermore, a gradual
exposure of the main concepts and ideas, in five stages, allow overcoming the
dichotomous classification without falling into the trap of over-simplifications and
over-generalizations. Most importantly, such a framework facilitates the attainment of
two important objectives: First, preventing pedagogical impediments for further
studies. Second, fostering the understanding that molecular species and bonding
scenarios that textbooks often designate as ‘exceptions’ can, in fact, be rationalized by
the same small number of principles used to rationalize the ‘regular cases’. Levy
Nahum et al report that at present, this unit has been tested in ten 11th-grades classes
and more research is clearly needed. The following quotes strongly enhance Levy
Nahum’s faith in the presented approach, suggesting that the new approach
encourages teachers and students to ask relevant questions and be interested of
understanding bonding principles, whereas the consequence of the traditional
presentation reduces students' reasoning and questioning, which is essential for a
deeper understanding of the main ideas. Both students and teachers acquired a much
deeper understanding of the underlying key concepts regarding bonding.

Teacher: I loved the idea that I can teach and explain all bonds based on a uniform
model. Starting from sub-microscopic ideas and moving up to the material world improved
the students' learning and thinking.

Student: The continuum scale of bonds helped me to understand…last year the teacher
said: 'it's one of the two' (covalent/ionic).

In the academic year 2010, the new program will be implemented in all 11th-grade
chemistry classes in Israel, which is possible because the educational system is
centralized. In due course, a full scale procedure, assessing both the teaching and
learning will be conducted, aiming at reporting broader and statistically sound field

27
results. In this system framework, building off a knowledge-in-pieces perspective,
Yayon, Mamlok-Naaman and Fortus (2010) describe the development and testing of a
matrix that represents a systematic organization of the canonical knowledge on
chemical bonding required at high-school level and a tool for representing students’
knowledge of bonding. The matrix contain three strands: the structure of matter at the
nano-scopic level, electrostatic interactions between charged entities, and energy
aspects related to bonding.

However, Levy Nahum (2007) is also aware of the weakness of such a


framework – using abstract theoretical ideas right from the start may prove difficult
for some students. According to Loucks-Horsley and Matsumoto (1999), in the
framework of reform in science education an extensive, dynamic, and long-term
professional development of science teachers should take place. Teachers need to
receive guidance and support throughout the various teaching and implementation
stages involving changes in the curriculum. It has been noted by Joyce and Showers
(1983) that teachers, in general, are excellent learners, and are interested in trying to
teach a new curriculum, as well as in improving and enriching their teaching methods.
However, in the context of introducing and implementing of this new unit, the
psychological process that teachers have to undergo is unique due to two main
reasons: (1) the importance of the topic - many chemistry teachers (and chemists) feel
that bonding is the "heart of chemistry", and (2) in Israel, this topic was taught for
decades in a very definite way, and, since the new curriculum will be implemented
starting from 2010, all the teachers have to adopt it and apply it.

Nevertheless, the teachers understood that the oversimplified teaching


approach they used routinely, for so many years, encouraged their students to learn by
rote-memorization rather than by using their thinking skills and scientific tools.

Teaching models: Finding ways to help students access curriculum science

A further stage of interpretation occurs when an individual teacher, or textbook


author, uses the curricular models as the starting point for developing teaching.
Teachers, in particular, will need to consider how they help particular groups of
students access the curricular models. They will often use further simplifications and
various forms of teaching models, designed to link with students’ current knowledge,

28
understanding and interests, and inevitably giving opportunities further potential for
distorting the original science. Again, the transformation will be undertaken by
someone distant from the development and application of the original scientific
models. Clearly, through these transformations processes there will be distortions of
the original scientific models: some intentional and principled, but possibly also due
to the limitations of the scientific understanding of those in charge with shaping
curricular and classroom knowledge. One particular concern raised by Justi and
Gilbert (2000) was the adoption of ‘hybrid’ models as curricular models. During the
historical development of scientific models, there may be many discrete models
proposed. Each of these will represent a certain stage of knowledge (the response to
combining the strengths of previous models with new features that add further
explanatory power in the light of anomalies or new data or a desire to expand the
range of application of a model). Within such a context, a scientific model will
usually be consistent and logical: but appreciating the features of the model may be
difficult without awareness of that context. According to Justi and Gilbert, who
considered the case of representation in the curriculum of models of the atom,
curriculum models can often be ‘hybrid models’ that are true to no specific historical
model but are rather “composites drawn from several distinct historical models”
(p.993). Such hybrids cannot readily be understood in terms of logical development of
the subject.

This characteristic of curricular models adds to the already significant burden for
students who may well be introduced to a range of models (some curricula models
intended as target knowledge; some teaching models intended as scaffolding towards
such knowledge; perhaps some awareness of formal scientific models from the media
or personal reading) with different but overlapping ranges of application. To the
teacher, there may be a logical sequence to support developing understanding (as in
the different models of oxidation or acidity), but for the students there is a very real
risk of ‘model confusion’ (Carr, 1984).

An example of this would seem to be basing teaching about chemical reactions and
bonding around the octet rule. Taber characterised students as entering study at
college/senior high school level holding a notion of chemical bonding deriving from
school study, and tied to the octet rule. He also identified common tendencies and

29
conceptions which were reported as an alternative conceptual framework (see the
section on students’ conceptions of bonding, above) – that is a model representing the
common features found across a range of students (Taber, 1998).

Instructional techniques

Chemical structure and bonding is a topic in which understanding is developed


through diverse models, which, in turn, are built upon range of physical principles;
and where students are expected to interpret a disparate range of symbolic
representations standing for chemical bonds (Taber & Coll, 2002). According to
Johnstone (1991) and Gabel (1996), matter can be represented on three levels,
macroscopic (physical phenomena), microscopic (particles), and symbolic-
representational (chemical and mathematical language). The symbolic level can be
seen as having the key role of mediating between the phenomenological-descriptive
level of what students can directly perceive and the abstract conceptual level of
theoretical entities such as quanticles (e.g. ‘H2’ can stand for both the substance and
the molecule, and so acts as means of linking one to the other). Yet in the teaching of
chemistry, teachers repeatedly (‘instinctively’) move from one level to another in their
teaching without making this explicit to their students each time they are doing it:
missing an opportunity to help the students integrate the levels and potentially
confusing them in terms of the current referent for a symbol (Taber, 2009b).

It is assumed by both curriculum developers and those involved in research on


learning that the models used can enhance scientific learning (Lunetta & Hofstein,
1981). It is however suggested that students are poorer at modeling than commonly
expected by teachers. Moreover, Gilbert (1998) argues that students tend to think that
models are scale replicas or incomplete copies of actual objects, and therefore they do
not look for ideas or seek purposes in the model form. Harrison and Treagust (2000)
however claim that models are more than communicative tools: rather that they are
also important links in the methods and product of science. Moreover they suggest
that students who participated in negotiating the shared and unshared attributes of
common analogical models for atoms, molecules, and chemical bonds actually used
these models more consistently.

30
The use of visual models for teaching the bonding concept

In recent years computer technologies in general, and web-based teaching and


learning in particular, have gained momentum in teaching and learning the sciences.
Several studies have noted the benefits of web-based learning and its vast potential to
empower learning and teaching in terms of its visualization, accessibility, and
dynamicity (Capri, 2001; Clark, 2004; Linn, Clark, & Slotta, 2003; Mistler-Jackson,
2000). More specifically, in alignment with the idea of visualization to support
students' learning the chemical bonding concept, Clark, (2004) noted the importance
of integrating computer-based visualizations in learning abstract concept and
phenomena. Kozma and Russel (2005) suggested that molecular models, simulations,
and animations have the potential to contribute to the learning of chemistry in general
and to better understanding of the chemical bonding concept in particular. Gilbert and
Justi (2002) wrote that the use of computerized models has been advocated as a way
to improve students' understanding of chemical phenomena by translating information
expressed in different representations (see also Kozma & Russel, 1997). Moreover (as
Justi & Gilbert (2002) wrote), studies on students' use of computer-based models have
shown that they can also improve students' visualization in chemistry (Barnea & Dori,
2000).

A recent study conducted in Israel by Frailich, Kesner, and Hofstein (2009)


investigated the effectiveness of a web-based environment in enhancing 10th grade
students’ understanding of the concept chemical bonding. Computer-based visual
models were developed and implemented for the purpose of demonstrating bonding
and the structure of matter. These were substantially based on student-centered
learning. Drawing on a combined quantitative and qualitative research study Frailich
et al. (2009) were able to conclude that the web-based learning activities which
integrated visualization tools with active cooperative learning strategies provided
students with opportunities to construct their knowledge regarding the abstract aspects
concept of chemical bonding.

31
Summary

In this review we have considered teaching and learning of one of the central concepts
in chemistry, bonding. We have considered the nature of the concept, and the reasons
why it might present a challenge when included in the upper secondary/senior high
school or college curriculum. We have suggested that some aspects of traditional
ways of teaching the topic may have contributed to the learning difficulties faced by
many learners. This was reflected in the literature discussing learners' conceptions of
chemical bonding where some key issues (over-reliance on the covalent-ionic bonding
dichotomy; adopting the octet rule as the basis of explanations of bonding; seeing
bonds as ontologically quite different from physical forces) seem to reflect common
teaching approaches.

A key theme here is the central involvement of modeling at all stages. Chemists use
models of chemical bonding, which are considered useful in supporting their work. It
has been suggested that for many chemists, the conceptual bonding ‘tools’ they
regularly apply may be somewhat different from the most advanced models being
developed by those chemists who actually research into ways of understanding
bonding itself. This can be understood as a pragmatic matter, which many chemists
behaving as instrumentalists, being content to apply bonding models that are ‘fit for
purpose’ in their own areas of research, despite having been shown to be deficient in
terms of not giving a full and authentic account of bonding in some contexts.

Traditional science curricula have included accounts of bonding models selected for a
mixture of their perceived importance in the subject and their judged suitability for
the students concerned. However, such an approach is based on the assumption that
school science and general cognitive development will allow learners to meet and
mater a sufficient sequence of curriculum models to reach the desired end-point at the
end of their upper secondary / senior high school courses. However, based on the
research evidence discussed in this review paper, for students this might be an
unrealistic assumption. Matters are even more complicated in such cases in which
chemistry (physical science) teachers may come to consider, or at least to teach as if
they consider, the curriculum models, which themselves are simplifications of the
chemical concepts, as something more than limited models that are partial
representations of the ‘real’ chemistry. If what is meant to be an education transition

32
state is understood instead as the desired pedagogic product, it becomes clear why
students commonly ascribe limited ideas (bonding dichotomies, the significance of
octets or full electron shells) the status of the way chemists have found the world to
be. Something that was meant to have transitional epistemological value is instead
understood as being absolute ontological reality.

We have shown in this review that the research literature is rather rich in information
resulting from diagnostic studies that were conducted over the years in many
countries. However, we have observed that although vast amount of information
exists it has had only limited influence on curriculum development (development of
learning materials and pedagogical remedies), teachers' professional development,
and classroom practice. In other words there is a huge gap between what we know
about learning the chemical bonding concept and what has been implemented in
educational systems. We have shown that the traditional approach to the teaching of
the concept is strongly embedded in the teachers' knowledge and in the various
curricular approaches around the world.

The chemical bond is one of the key concepts that should be taught in every upper
secondary school, college and undergraduate chemistry course. As an epilogue to our
review, we sincerely hope that the conglomerate of information presented in this
paper will serve as a vehicle for new, more scientifically and pedagogically sound,
curricula that will reach the teachers in the school system.

References

Ashkenazi, G., & Kosloff, R. (2006). The uncertainty principle and covalent bonding.
The Chemical Educator, 11, 66-76.

Atzmon, A. (1991). Natural science curriculum as designer of the image of


science using rhetorical patterns within the socio-political system.
Unpublished Doctoral dissertation, Hebrew University, Jerusalem (in
Hebrew).

Barnea, N., & Dori, Y.J. (2000). Computerized molecular modeling: The new
technology for enhance model perception among chemistry educators and
learners. Chemistry Education: Research and Practice in Europe, 1, 109–120.

Ben-Zvi, R., Bat-Sheva, E., & Silberstein, J. (1986). Is an atom of copper malleable?
Journal of Chemical Education, 63(1), 64-66.

33
Bodner, G., & Domin, D. (1998). Mental models: The role of representations in
problem solving in chemistry. International Council for Association in Science
Education, Summer Symposium, Proceedings.

Brook, A. (1986). Children’s understanding of ideas about energy: a review of the


literature,. In R. Driver & R. Millar (Eds.), Energy Matters - Proceedings of
an invited conference: teaching about energy within the secondary science
curriculum, (pp. 33-45). Leeds: Centre for Studies in Science and
Mathematics Education.

Butts, B., & Smith, R. (1987). HSC Chemistry Students’ Understanding of the
Structure and Properties of Molecular and Ionic Compounds. Research in
Science Education, 17, 192-201.

Carpi, A. (2001). Improvements in undergraduate science education using web-based


instructional modules: The natural science pages. Journal of Chemical
Education, 78, 1709–1712.

Carr, M. (1984). Model confusion in chemistry. Research in Science Education, 14,


97-103.

Clark, D. (2004). Hands-on investigation in Internet environments: Teaching thermal


equilibrium. In: M.C. Linn, E.A. Davis, & P. Bell (Eds.), Internet
environments for science education. (pp. 175–200). Mahwah, NJ: Erlbaum.

Coll, R. K., & Treagust, D. F. (2001). Learners' mental models of chemical bonding.
Research in Science Education, 31(3), 357-382.

Coll, R. K., & Treagust, D. F. (2003). Investigation of Secondary School,


Undergraduate, and Graduate Learners’ Mental Models of Ionic Bonding.
Journal of Research in Science Teaching, 40(5), 464-486.

Cros, D. l., Chastrette, M., & Fayol, M. (1988). Conceptions of second year university
students of some fundamental notions in chemistry. International Journal of
Science Education, 10(3), 331-336.

Cros, D. l., Maurin, M., Amouroux, R., Chastrette, M., Leber, J., & Fayol, M. (1986).
Conceptions of first year university students of the constitution of matter and
the notions of acids and bases. European Journal of Science Education, 8(3),
305-313.

Culotta, E. & Koshland DE, Jr. (1992). The molecule of the year: NO news is good
news. Science 258, 1862-1865.

de Vos, W. & Pilot, A. (2001). Acids and Bases in Layers: The Stratal Structure of an
Ancient Topic. Journal of Chemical Education, 78, 494–499.

Fensham, P. (1975). Concept formation. In D. J. Daniels (Ed.), New movements in the


study and teaching of chemistry (pp. 199-217). London: Temple Smith.

34
Frailich, M., Kesner, M. & Hofstein, A. (2009). Enhancing students’ understanding of
the concept of chemical bonding by using activities provided on an interactive
website. Journal of Research in Science Education, 46(3), 289–310.

Gabel, D. (1996, July). The complexity of chemistry: Research for teaching in the 21st
century. Paper presented at the 14th International Conference on Chemical
Education, Brisbane, Australia.

Gilbert, J. (2004). Models and Modelling: Routes to More Authentic Science


Education. International Journal of Science and Mathematics Education, 2, 115-130.

Gilbert, J. K., Osborne, R. J., & Fensham, P. J. (1982). Children’s science and its
consequences for teaching. Science Education, 66(4), 623-633.

Gillespie, R. J. (1997). The great ideas of chemistry. Journal of Chemical Education,


74(7), 862-864.

Glasersfeld, E. v. (1989). Cognition, Construction of Knowledge, and Teaching


[Electronic Version]. Synthese, 80, 121–140, from
http://www.univie.ac.at/constructivism/EvG/papers/117.pdf 80-1-1

Griffiths, A. K., & Preston, K. R. (1992). Grade-12 students’ misconceptions relating


to fundamental characteristics of atoms and molecules. Journal of Research in
Science Teaching, 29(6), 611-628.

Hapkiewicz, A. (1991). Clarifying chemical bonding. The Science Teacher, 58(3), 24-
27.

Harrison, A. G. & Treagust, D. F. (1998). Modelling in science lessons: Are there


better ways to learn with models? School Science and Mathematics, 98, 420-
429.

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

Hawkes, S. J. (1996). Salts are mostly not ionised. Journal of Chemical Education,
73(5), 421-423.

Henderleiter, J., Smart, R., Anderson, J., & Elian, O. (2001). How Do Organic
Chemistry Students Understand and Apply Hydrogen Bonding? Journal of
Chemical Education, 78(8), 1126-1130.

Herron, J. D. (1996). The chemistry classroom: Formulas for successful teaching,


Washington, DC: American Chemical Society.

Holman, J. (1986). Teaching about energy - the chemical perspective. In R. Driver &
R. Millar (Eds.), Energy Matters - Proceedings of an invited conference:
Teaching about energy within the secondary science curriculum (pp. 47-52).
Leeds: Centre for Studies in Science and Mathematics Education.

35
Hurst, O. (2002). How we teach molecular structure to freshmen, Journal of Chemical
Education, 79(6), 763-764.

Johnstone, A. H. (1991). Why is science difficult to learn? Things are seldom what
they seem. Journal of Computer Assisted Learning, 7, 75-83.

Joyce, B., & Showers, B. (1983). Power and staff development through research on
training. Alexandria, VA: Association for Supervision and Curriculum
Development.

Justi, R., & Gilbert, J. K. (2000). History and philosophy of science through models:
some challenges in the case of ‘the atom’. International Journal of Science
Education, 22(9), 993-1009.

Justi, R., & Gilbert, J. (2002). Models and modeling in chemical education. In J. K.
Gilbert, O. D. Jong, R. Justy, D. F., Treagust, & J. H. Van Driel (Eds.),
Chemical education: Towards research-based practice (pp. 47-68).
Dordrecht: Kluwer.

Kind, V. (2009). Pedagogical content knowledge in science education: perspectives


and potential for progress, Studies in Science Education, 45(2),169-204.

Kozma, R., & Russell, J. (2005). Modelling students becoming chemists: Developing
representational competence. In: J.K. Gilbert (Ed.), Visualization in science
education. (pp. 121–145). Dordrecht: Academic Publishers.

Levy Nahum, T., Hofstein, A., Mamlok-Naaman, R., & Bar-Dov, Z. (2004). Can final
examinations amplify students’ misconceptions in chemistry. Chemistry
Education: Research & Practice, 5(3), 301-325.

Levy Nahum, T., Mamlok-Naaman, R., Hofstein, A. & Krajcik, J. (2007). Developing
a new teaching approach for the chemical bonding concept aligned with
current scientific and pedagogical knowledge, Science Education, 91, 579-603.

Levy Nahum, T. (2007). Teaching the concept of Chemical Bonding in high-school:


Developing and implementing a new framework based on the analysis of
misleading systemic factors. Unpublished Doctoral Dissertation, The
Weizmann Institute of Science, Rehovot (In Hebrew).

Levy Nahum, T., Mamlok-Naaman, R., Hofstein, A., & Kronik, L. (2008). A new
"bottom-up" framework for teaching chemical bonding. Journal of Chemical
Education, 85, 1680-1685.

Linn, M.C., Clark, D., & Slotta, J.D. (2003). WISE design for knowledge integration.
Science Education, 87, 517–538.

Loucks-Horsley, S., & Matsumoto, C. (1999). Research on professional development


for teachers of mathematics and science: The state of the scene. School
Science and Mathematics, 99, 258-271.

Lunetta, V, N., & Hofstein, A. (1981). Simulation in science education. Science


Education, 65, 252-273.

36
Magnusson, S., Krajcik, J., & Borko, H. (1999). Nature, sources, and development of
pedagogical content knowledge for science teaching. In J. Gess-Newsome and
N. G. Lederman (Eds.), Pedagogical content knowledge and science education
(pp. 95-132). Dordrecht: Kluwer.

Mistler-Jackson, M., & Songer, N.B. (2000). Student motivation and Internet
technology: Are students empowered to learn science? Journal of Research in
Science Education, 37, 459–479.

Myers, R. T. (1979). Physical and chemical properties and bonding of metallic


elements. Journal of Chemical Education, 56, 712–713.

Naaman, R. & Kronik, L. (2004). Personal communication.

Naaman, R. & Vager, Z. J. (1999). Dipole moment of non-rigid molecular systems-


The dipole moment of hydrochloric dimmers. Journal of Chemical Physics,
110, 359–362.

Nakanaga, T., Buchhold, K. & Fumiyuki, I. (2002). Weak hydrogen bond interactions
in the aniline–alkene (CH2=CH(CH2)nH, n=0,1,2) clusters studied by infrared
depletion spectroscopy. Chemical Physics, 277, 171–178.

Nakiboğlu, C., Tsaparlis, G., & Taber, K. S. (2009). Student understanding of ionic
bonding in three countries: molecular versus electrostatic thinking. Paper
presented at the European Science Education Research Association
Conference, Istanbul, Turkey, August 31st - September 4th 2009.

Nicoll, G. (2001). A report of undergraduates’ bonding misconceptions. International


Journal of Science Education, 23(7), 707-730.

Pauling, L. The Nature of the Chemical Bond, 3rd ed.; Cornell University Press:
Ithaca, NY, 1967.

Papageorgiou, G., & Johnson, P. (2005). Do Particle Ideas Help or Hinder Pupils'
Understanding of Phenomena? International Journal of Science Education,
27(11), 1299 - 1317.

Peterson, R. F. & Treagust, D. F. (1989). Grade-12 students' misconceptions of


covalent bonding and structure, Journal of Chemical Education, 66, 459-460.

Peterson, R., Treagust, D., & Garnett, P. (1986). Identification of secondary students’
misconceptions of covalent bonding and structure concepts using a diagnostic
instrument. Research In Science Education, 16, 40-48.

Peterson, R. F., Treagust, D. F., & 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(4), 301-314.

Robinson, W. (2003). Chemistry problem-solving: Symbol, macro, micro, and


process aspects. Journal of Chemical Education, 80, 978-982.

37
Sánchez Gómez, P. J., & Martín, F. (2003). Quantum versus ‘classical’ chemistry in
university chemistry education: A case study of the role of history in thinking
the curriculum. Chemistry Education: Research & Practice, 4(2), 131-148.

Schmidt, H., Kaufmanna, B. & Treagust, D. (2009). Students’ understanding of


boiling points and intermolecular forces. Chemistry Education: Research & Practice,
10, 265–272.

Sproul, G. J. (2001). Electronegativity and bond type: Predicting bond type. Journal
of Chemical Education, 78, 387–390.

Taber, K. S. (1994). Misunderstanding the Ionic Bond. Education in Chemistry, 31(4),


100-103.

Taber, K. S. (1997a). Student understanding of ionic bonding: molecular versus


electrostatic thinking? School Science Review, 78(285), 85-95.

Taber, K. S. (1997b). Understanding Chemical Bonding: The development of A level


students' understanding of the concept of chemical bonding. University of
Surrey.

Taber, K. S. (1998). An alternative conceptual framework from chemistry education.


International Journal of Science Education, 20(5), 597-608.

Taber, K. S. (2000). Multiple frameworks?: Evidence of manifold conceptions in


individual cognitive structure. International Journal of Science Education,
22(4), 399-417.

Taber, K. S. (2001a). Building the structural concepts of chemistry: some


considerations from educational research. Chemistry Education: Research and
Practice in Europe, 2(2), 123-158.

Taber, K. S. (2001b). Shifting sands: a case study of conceptual development as


competition between alternative conceptions. International Journal of Science
Education, 23(7), 731-753.

Taber, K. S. (2002a). Chemical misconceptions - prevention, diagnosis and cure:


Theoretical background (Vol. 1). London: Royal Society of Chemistry.

Taber, K. S. (2002b). Conceptualizing quanta - illuminating the ground state of


student understanding of atomic orbitals. Chemistry Education: Research and
Practice in Europe, 3(2), 145-158.

Taber, K. S. (2003a). Mediating mental models of metals: acknowledging the priority


of the learner’s prior learning. Science Education, 87(732-758).

Taber, K. S. (2003b). The atom in the chemistry curriculum: fundamental concept,


teaching model or epistemological obstacle? Foundations of Chemistry, 5(1),
43-84.

38
Taber, K. S. (2009a). Progressing Science Education: Constructing the scientific
research programme into the contingent nature of learning science.
Dordrecht: Springer.

Taber, K. S. (2009b). Learning at the symbolic level. In J. K. Gilbert & D. F. Treagust


(Eds.), Multiple Representations in Chemical Education (pp. 75-108).
Dordrecht: Springer

Taber, K. S., & Coll, R. (2002). Bonding. In J. K. Gilbert, O. D. Jong, R. Justy, D. F.,
Treagust, & J. H. Van Driel (Eds.), Chemical education: Towards research-
based practice (pp. 213-234). Dordrecht: Kluwer

Taber, K. S., & Watts, M. (1996). The secret life of the chemical bond: students’
anthropomorphic and animistic references to bonding. International Journal of
Science Education, 18(5), 557-568.

Teichert, M. & Stacy, A. (2002). Promoting understanding of chemical bonding and


spontaneity through student explanation and integration of ideas, Journal of
Research in Science Teaching, 39(6), 464-496.

Tsaparlis, G. (1997). Atomic and molecular structure in chemical education. Journal


of Chemical Education, 74, 922-925.

Vinner, S. (1997). The pseudo-conceptual and the pseudo-analytical thought


processes in mathematics learning. Educational Studies in Mathematics, 34, 97
– 129.

Weinhold, F. & Landis, C. R. (2005). Valency and Bonding: A Natural Bond Orbital
Donor–Acceptor Perspective; Cambridge University Press: Cambridge.

Wightman, T., Green, P., & Scott, P. (1986). The Construction of Meaning and
Conceptual Change in Classroom Settings: Case Studies on the Particulate
Nature of Matter. Leeds: Centre for Studies in Science and Mathematics
Education, University of Leeds.

Woicik, J. C., Nelson, E. J., Kronik, L., Jain, M., Chelikowsky, J. R., Heskett, D.,
Berman, L. E. & Herman, G. S. (2002). Hybridization and Bond-Orbital
Components in Site-Specific X-Ray Photoelectron Spectra of Rutile TiO2.
Physical Review Letters, 89, 1–4 (077401).

Yayon, M., Mamlok-Naaman, R., & Fortus, D. (2010). Characterizing and


Representing Students’ Knowledge of Chemical Bonding (submitted).

39
Figure 1: A schematic illustration of the traditional approach for teaching chemical
bonding (Levy Nahum et al 2008)

40
Figure 2: The curricular process and outcomes regarding the traditional teaching of the
bonding concept in Israel

41
Figure 3: A schematic illustration of a new "bottom-up" framework for teaching

chemical bonding

42
Figure 4: A schematic energy curve for any two atoms that interact

43
Figure 5: A schematic continuous scale of bond strengths

44
Category Misconception

Bond polarity: • Equal sharing of electron pairs 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 • The shape of a molecule is due to equal repulsion between the


shape: bonds.

• Bond polarity determines the shape of a molecule.

• The V-shape in a molecule of the type SCl2 is due to repulsion


between the non-bonding electron pairs [only].

Intermolecular • Intermolecular forces are the forces within a molecule.


forces:
• Strong intermolecular forces exist in a continuous covalent solid.

• Covalent bonds are broken when a substance changes shape.

Polarity of • Non-polar molecules [only] form when the atoms in the molecule
molecules: have similar electronegativities.

• Molecules of the type OF2 are polar as the non-bonding electrons


on the oxygen form a partial negative charge.

Octet rule: • Nitrogen atoms can share 5 electron pairs in bonding.

Lattices: • High viscosity of some molecular solids is due to strong bonds in


the continuous covalent lattice.

Table 1: Misconceptions related to covalent bonding reported


by Peterson and colleagues

45

You might also like