You are on page 1of 16

Res Sci Educ (2008) 38:529544

DOI 10.1007/s11165-007-9060-y

Models in Physics, Models for Physics Learning,


and Why the Distinction may Matter in the Case
of Electric Circuits
Christina Hart

Published online: 20 September 2007


# Springer Science + Business Media B.V. 2007

Abstract Models are important both in the development of physics itself and in teaching
physics. Historically, the consensus models of physics have come to embody particular
ontological assumptions and epistemological commitments. Educators have generally
assumed that the consensus models of physics, which have stood the test of time, will also
work well as teaching models, and for many topics this assumption is at least unproblematic
and in many cases productive. However, in the case of electric circuits the consensus
models are highly abstract and consequently inaccessible to beginning learners. Certain
historically derived analogues for the consensus models are accepted in texts, but these are
demonstrably ineffective for helping learners grasp the fundamental concepts of electric
circuits. While awareness of other models circulates informally in the teaching community,
these are not well documented in the science education literature and rarely referred to in
authoritative texts, possibly because the models do not share the ontological assumptions
and epistemological commitments that characterise consensus models. Consequently these
models have not been subjected to a disciplined critique of their effectiveness for teaching
purposes. In this paper I use criteria drawn from the science education literature to reflect
on why I have found particular models valuable in teaching electric circuits. These criteria
contrast with the epistemological and ontological features that characterise the consensus
models of science, and my reflection leads me to attend explicitly to the ways in which
meanings are created within physics. This suggests that all models, whether consensus
models or not, can be used more knowingly for important educational ends.
Keywords Analogies . Electric circuits . Models . Science learning . Science teaching

Introduction: Models in Physics and Physics Education


It could be said that the business of science is to develop models of real world phenomena. The
consensus models of physics (and of science in general) are those models adopted by the
C. Hart (*)
Faculty of Education, The University of Melbourne, Melbourne, VIC, Australia
e-mail: c.hart@unimelb.edu.au

530

Res Sci Educ (2008) 38:529544

community of scientists as enabling an authoritative explanation of a particular phenomenon;


these models play a powerful role, affording a means to represent and encode understanding
and explanation, and enabling questions to be asked and predictions to be made.
It goes without saying that one of the goals of science education should be to help students
understand those consensus models of science that have been honed over time and which have
proved particularly powerful, and these typically figure prominently in school science curricula.
For example, the curricula for senior physics in the Australian States of New South Wales
(NSW Board of Studies 2002) and Victoria (Victorian Curriculum and Assessment Authority
2003) specify or imply the following consensus models from the discipline of physics:

&
&
&
&
&
&

models for matter: atomic and nuclear models (optional in New South Wales),
wave-particle duality;
models for sound and other waves in material media (optional in Victoria);
models for light: ray, particle, wave, wave-particle duality;
models for motion: Newtons laws, Einsteinian relativity; Newtonian gravitation;
electron transport models for electric circuits, including mathematical representations such as those for power dissipation, current conservation, Ohms law;
field models: electric fields (New South Wales only), magnetic fields, electromagnetic
waves.

The curriculum specifications for the lower secondary years in both States (New South
Wales Board of Studies 2003, Victorian Curriculum and Assessment Authority 2005) also
include some of these models, as well as the kinetic particle model for matter.
Historically, the consensus models of physics have acquired particular epistemological
and ontological features. A high value is placed on plausibility (fit with the data),
parsimony (reliance on a few concepts), coherence (congruence with other explanations of
related phenomena), generalisability (transferability), fruitfulness or power (potential to
promote questions, enable predictions, suggest experiments and observations) (Gilbert et al.
1998a, pp 8788, 1998b, p 195). These epistemological features are well-recognised, but
the underlying ontological assumptions of the models have received less attention. Thus,
the entities involved are generally non-volitional, non-self-determining, and their behaviour
can be described in terms of well-defined, and deterministic or probabilistic interactions
(Gilbert et al. 1998b, p 197). Consequently the models allow precise predictions that can be
matched against actual measurements of the relevant phenomenon. The high degree of
correspondence between the predictions and measurements leads to the presumption that
the entities of physics models, such as fields, forces, electrons, atoms, molecules, are as real
as the phenomena they explain. In this realist ontology, the explanatory entities are taken to
be given in the real world, having been discovered there by physics. Physics, it can be
concluded, is about discovering the truth of the way the world is.
We can thus become distracted from seeing that models are always constructions. A
model is an analogy, and the correspondence with the real world is never complete. The
representations, concepts, relationships and explanatory entities that figure in the model are
not given in the phenomenon itself. They are overlays on reality, produced through the
human activities of striving to understand, predict and control the physical world.
Even where curricula specifically refer to models, the consensus models are typically
presented without the modelling process itself being unpacked, and the underlying
epistemological commitments and ontological assumptions made explicit (for example
New South Wales Board of Studies 2003, Victorian Curriculum and Assessment Authority
2003). Indeed, in some curriculum documents the realist ontology is so taken for granted
that there is little or no meaningful reference to the role of models (for example Victorian

Res Sci Educ (2008) 38:529544

531

Curriculum and Assessment Authority 2005, New South Wales Board of Studies 2002), so
that physics effectively becomes defined as the study of interactions between real (as
opposed, for example, to imagined, spiritual or social) entities.

The Problematic Case Of Electric circuits


Several authors have argued that models play an important role in teaching and learning
physics (Duit 1991; Harrison and Treagust 1998; Heywood 2002), although it seems that
teachers may make rather limited use of models to support student learning (Treagust et al.
1992). The question of whether the consensus models of physics necessarily make good
models for teaching the relevant concepts has received little attention. Nonetheless, for many
topics, the assumption that the realist consensus models do work well as teaching models is at
least unproblematic and in many cases productive; for example the kinetic particle model
poses no obvious pedagogical dilemmas (Mulhall et al. 2001). However, in the case of
electric circuits, there is considerable uncertainty about the pedagogical purposes and
usefulness of particular models (Mulhall et al. 2001; Stocklmayer and Treagust 1994).
Textbook treatments of electric circuits for beginning students (for example Nardelli
2006, Lofts and Evergreen 2007) commonly assume an electron-transport model,
explaining electric current in terms of the flow of electrons around a circuit. This model
fits comfortably within the realist ontological framework of physics. However, the model is
often not made explicit and, in any case, cannot provide a complete and coherent account of
how electrons are involved in the transport and distribution of energy around the circuit
(Mulhall et al. 2001; Stocklmayer and Treagust 1994). As a consequence, as Gunstone et al.
(2007) have shown, introductory texts may simply introduce the terms energy and/or
voltage with little or no explanation and, in some cases, without even clear definition. Some
introductory texts invoke Ohms law (for example Lofts and Evergreen 2007, p 108) in an
apparent attempt to explain energy transfers, but this is an empirical relationship that, by its
very nature, cannot fulfil an explanatory role in order to address the fundamental limitations
of the electron transport model.
The use of overtly analogical models in textbooks, especially recent texts, appears to be
quite rare, although Nuffield Primary Science (Nuffield-Chelsea Curriculum Trust 1993, p
81) makes use of a bicycle chain model, and Lofts et al. (2004, p 307) use a rope model.
Only two analogies appear to be more widely used: a water-flow model (for example Hewitt
1987, pp 510, 513; Stannard and Williamson 2006, p 60) and a gravitational model (for
example Halliday and Resnick 1988, p 6631; Storen and Martine 2004, p 266), or some
combination of these (for example Ohanian 1994, p 557; Kane and Sternheim 1988, p 407).
It is interesting to speculate on the reasons why physics texts appear comfortable to employ
these models when they avoid other analogues that are potentially pedagogically useful.
The most likely reason for the acceptance of the water-flow model for electric circuits is
that early theories of electric circuits were based on belief in an electric fluid that flowed
like water in a pipe (Spencer 1999). By the time Thomson and Millikan discovered the
electron, the mathematical theory of circuits was already complete, and electrical
engineering was an established discipline and successful practice, all based on the
presumption of an electrical fluid. This situation is similar to that which resulted in the
also overtly analogical ray model for light being incorporated into the cannon. Within
Newtons view of light as a stream of particles, light rays could be seen as real entities and
1

The analogy does not appear in later editions of this text (see Halliday et al. 1997).

532

Res Sci Educ (2008) 38:529544

ray optics enabled the mathematics of lenses, leading to improvements in the design of
telescopes and other optical instruments (Gilbert et al. 1998a, p 85). So, by historical
accident, the water-flow model and ray model became sedimented in the canon of
consensus models at a time when it was believed that the entities in the models did
correspond with actual entities in the real world.
The gravitational model for electric current is promoted, I suspect, for a different reason.
The mathematics of the motion of a mass in a gravitational field and of charged particles in
an electric field is exactly parallel. This displays a prized characteristic of the discipline of
physics: the unity of the conceptual structure in disparate topics. Physicists of a mystical
bent have taken this discovery of underlying unity as evidence that creation displays the
mind of God (Wertheim 1995). A similar unity prevails between the mathematical
expressions for electric circuits and for the flow of water in closed pipes, thus providing
further grounds for incorporating the water-flow model into the cannon of consensus
models. [Gilbert et al. (1998b, p 188) point to an early example of this pursuit of unity,
between the mathematical wave models for sound and light.] Feynmann however, cautions
against reading too much into any appearance of unity in the equations of physics
(Feynmann 1964, p 1212).
The pedagogical limitations of the water-flow model are well known: beginning
students of physics do not have the requisite understanding of water pressure and flow
that might make the analogy productive (Gentner and Gentner 1983, p 126). If the
purpose of the model is to help students make the crucial distinction between electrical
current (or charge) and electrical energy (voltage) then the model most likely fails
because students cannot discriminate the notion of pressure or push of the pump from
the notion of the rate of water flow. The two concepts are too closely intertwined, and the
model is therefore of no help in discriminating the similarly closely intertwined concepts
of electrical current and electrical energy. The potential difficulties with the model are
further accentuated by the propensity of textbook writers to simply imply the model
through the use of words such as push or pressure to describe voltage, without
making the model itself explicit (Mulhall et al. 2001, p 580). The gravitational model
similarly offers little support to students if their understanding of the energy transformations that may be accomplished by the forces of gravity and friction are themselves
rudimentary, as is likely the case for beginning students.
Perhaps the inadequacy of these models as teaching models goes some way towards
explaining why it is that teachers have difficulty articulating useful models for teaching
electric circuits: this was a difficulty for even the outstanding teachers that Mulhall and
her colleagues worked with (Mulhall et al. 2001, p 5789). As Mulhall et al. note, this
group of teachers did not have the same difficulty with articulating their teaching models in
the topic of Newtonian mechanics. I am suggesting that the Newtonian models inherited
from the discipline can be elaborated into good teaching models for mechanics, whereas the
inherited models for electric circuits supply a meagre resource for the challenges of
teaching this difficult topic.
Some educators (Mulhall et al. 2001, p 578; Stocklmayer and Treagust 1996) have argued
that only an electric field model can enable the holistic reasoning about electric circuits that is
necessary, not only to account for a.c. circuits, but even to address the common student
question about how the current/charge knows about the presence of a second globe in a
d.c. series circuit (Mulhall et al. 2001, p 578). Textbooks for advanced students do usually
provide a formal mathematical definition of electrical potential in terms of work done in an
electrostatic field (for example Weicek et al. 2005, Storen and Martine 2004, Halliday et al.
1997, Serway and Faughn 2000), and this formal mathematical definition is then applied to

Res Sci Educ (2008) 38:529544

533

electrical potential in electric circuits. However the field model itself is not usually applied
explicitly to electric circuits (exceptions are Storen and Martine 2004, p 2667; Halliday et al.
1997, p 679), and so students are unlikely to perceive its value as a tool for holistic reasoning
about circuits. For beginning students, no matter what their age, formal mathematical
definitions are unhelpful, even alienating, and the electric field would seem to be inevitably
inaccessible. Abstract by its very nature, the concept of an electric field has no satisfactory
analogue indeed Dave Heywood cites Richard Feynmann as claiming that it is pointless
even to seek such an analogy (Heywood 2002, p 236).
Alternative Teaching Models for Electric Circuits
Nonetheless, classroom practitioners do have a repertoire of analogues for the electrontransport model (see, for example, Heywood 2002, p 238). The repertoire includes
variations on Dedre Gentner and Donald Gentners (1983) moving-crowds model: people
moving through corridors, cars on freeways can effectively model resistance and its effect
on the whole circuit. Elaborated versions of such models that use people carrying Smarties
or jelly beans, loaded coal trucks or shopping trolleys, can help make the distinction
between charge and energy and can model the role of batteries. Then the bicycle-chain
model (Nuffield Chelsea Curriculum Trust 1993, p 81) has the advantage of both avoiding
the representation of energy as a substance and showing that energy can be moved without
the actual movement of a material carrier.
But while awareness of these models circulates informally in the teaching community,
they are not well documented in the science education literature and, in my experience,
rarely referred to in authoritative physics texts. They seem to have a rather clandestine
existence, and consequently are easily dismissed as idiosyncratic knowledge of individuals
(Mulhall et al. 2001, p 581). Perhaps this is partly because they do not embody the
ontological assumptions and epistemological commitments that characterise consensus
models. In one sense this is just a particular manifestation of the general problem we have
with distilling, validating and communicating teachers craft knowledge, that knowledge
that experienced practitioners accumulate over time, of how to effectively structure and
guide the learning of particular concepts and topics (van Driel et al. 1998). However, one
consequence of this neglect is that the use of such models is not subjected to the kind of
critical analysis that could clarify their role in supporting meaningful learning.
In my own teaching I attempt to help students construct an understanding of simple
electric circuits that incorporates the following conceptual elements:
1. why a complete circuit is necessary for electrons to move in a simple circuit;
2. how the movement of negative charges can appear indistinguishable from the
movement of positive charges;
3. why charges move and what causes resistance;
4. why resistance at one point in the circuit affects the current through the whole circuit;
5. the role of the battery is to supply energy, not create charges;
6. how current and energy are distinguished;
7. how energy is distributed around the circuit, including how electrons know what is in
the circuit and so where to deposit different amounts of energy.
With both general science and senior physics classes, I call on a succession of different
analogical models. I was first introduced to these models through the work of the McClintock
Collective, a network of teachers in Victoria who, in the 1980s and early 1990s worked to
increase the participation of girls in school science (see for example Gianello et al. 1988;

534

Res Sci Educ (2008) 38:529544

McClintock Collective and various authors 1989). Some are variants of the moving crowds
model, others were the result of original ideas of Collective members.
As a teacher in school, my impressions of the value of these models depended on
feedback from students during my interactions with them in the classroom, and the
understanding they demonstrated in assessment tasks. I became increasingly convinced of
the pedagogical value of the models I employ, even as I wondered about their neglect by
physics textbook writers and education researchers. But I had little concrete data of a kind
that could support the disciplined critique I felt was needed, or illuminate the specific ways
in which the various models might or might not support students developing
understanding of electric circuits concepts.

Criteria for a Teaching Model


Nonetheless, my instincts have been informed by my encounters with constructivist views
of learning, which have highlighted the learning demand associated with particular concepts
and the complexity of the task facing the learner. However, while constructivist theories
have advanced our understanding of science learning, they have been of more limited use in
guiding us to improved classroom teaching approaches (Leach and Scott 2003), and this
limitation is more apparent for some topics such as electric circuits than others. Sociocultural theory (for example Wells 1999), which has gained prominence more recently, has
addressed some aspects of learning that remained problematic within constructivist theories:
in particular, the status of science as a body of knowledge and the role of the classroom
teacher in making that body of knowledge available to students (Leach and Scott 2003).
Within the socio-cultural view, learning entails becoming competent in using the tools that
are available within a particular culture. We develop competence by participating, perhaps
initially by imitation, in the kinds of social activities that makes the tool itself useful. The
activities of science involve several types of tools: physical tools (equipment), procedures
(skills), and linguistic tools (concepts, models). Learners become competent with these
tools by participating in the activities of explaining, predicting and controlling the physical
world. This view of learning, with its emphasis on the acquisition of skill in using linguistic
tools, highlights the importance of classroom discourse, including the teachers role of first
talking into existence the entities of science, then helping students to appropriate the
linguistic and other tools of science for their own purposes, and giving them practice at
talking science (Lemke 1990).
Together, constructivist and socio-cultural theories suggest a set of criteria that might
apply if a model is to be used for teaching purposes. These criteria provide a possible
starting point for undertaking the kind of critique that I consider is needed if we are to better
understand what makes a model pedagogically useful.
1. The model must be initially intelligible to students, and must then come to appear first
plausible and, finally, fruitful (Strike and Posner 1985). For the model to be intelligible,
its ontological basis must first be intelligible: in other words, at some level it must be
understood why an explanation is given in terms of certain entities and not others. The
model must be plausible and fruitful in that it can account for observed phenomena,
particularly phenomena that the students themselves find puzzling, and suggest further
questions, experiments or observations.
2. The causal mechanisms that the model supplies must be meaningful to the students, so
that students can think about the model in their own terms. (Heywood 2002, p 234)

Res Sci Educ (2008) 38:529544

535

3. The model must allow common conceptual difficulties and misconceptions to be


articulated and addressed. (Gilbert et al. 1998b, p 192)
4. The model must engage students imaginations and intellects, in order to promote a
rich classroom discourse (Heywood 2002) which students can freely participate in and
contribute to. The teachers role is to guide the flow of discourse (Mortimer and Scott
2003), so that meaningful understanding is socially constructed in the classroom,
misconceptions are addressed and conceptual confusions and difficulties clarified.
5. The model must enable students to move towards an understanding of the relevant
consensus models of science. This should be viewed as a long term goal: in some
topics, including specifically electric circuits, it may be counter-productive to envisage
that students will find the consensus models of science either intelligible, plausible or
fruitful early in their science learning experiences.
6. The model should be overtly presented (Gilbert et al. 1998b, p 192). It should also
provide insight into: the nature and purpose of models in scientific thinking; the
epistemological commitments and ontological assumptions that underlie consensus
models; the fact that all models have limits and their application must not be pushed
beyond those limits, and that a variety of models may need to be flexibly employed in
order to construct a complete explanation of the phenomenon (see Harrison and
Treagust 2000; Heywood 2002).
These criteria are rather different from the features that have historically determined
whether or not a model was accepted as a consensus model within the discipline of physics.
In particular, it is explicitly not a requirement of teaching models that the entities in the
model refer to entities that are postulated to actually exist in the real world. Neither is it a
requirement that the entities involved are non-volitional, non-self-determining, or that their
behaviour can be described in determinist terms. Rather, as Dave Heywood has argued, the
requirement is that the model has hermeneutic potential, that there is the possibility of
interpretation, or multiple readings, of the analogy (Heywood 2002). In hermeneutic
theory there is a need to recognise the powerful constraint and possibilities afforded in
analogical reasoning. ... the constraint of language in analogical reasoning ... is a necessary
condition of interpretation, for without such constraint there would be no recognition that
there was something there to interpret. (Heywood 2002, p 240 [emphases in the original])

Data Collection and Analysis


In 2001 I used these models in a series of workshops for a group of 26 experienced
teachers, who did not have a strong background in physics, enrolled in a Post-Graduate
Certificate (PGC) Course in Science Education (Physics). Typically, these teachers were
biology specialists who were teaching physics as part of a general science program at
years 7 to 10 (lower secondary, age 1116 years). They were rather apprehensive about
their own understanding of the electric circuits and their ability to teach the topic effectively
as part of a general science curriculum. They therefore brought a critical pedagogical
perspective to their learning but, on their own admission, experienced the conceptual
challenges from a position that was rather close to that of their students.
As one of the assessment requirements, the participants completed a written reflection on
their experiences of learning two topics that they selected from those covered in the course.
Electric circuits and forces were the most commonly selected topics. I had previously
obtained ethical approval to seek students consent to retain copies of assignments for

536

Res Sci Educ (2008) 38:529544

research purposes, and eight participants gave permission to use reflections related to
electric circuits. All of these participants were female, a fact that reflected the
predominantly female enrolment in the course. At the time of seeking ethical clearance I
was just acting on a hunch that the reflections would provide me with useful data for better
understanding my own teaching and students experiences of learning. I did not have
specific research questions in mind. However, I was both affirmed and challenged as I
pondered the reflections of these eight students. I realised that they provided some quite
specific insights into the role that the models played in helping them develop a personally
meaningful understanding of concepts that had previously appeared intractably abstract. I
began to see that that here was some data that could support at a preliminary level at least
a critical analysis of the models I was using, providing some evidence for the models
pedagogical effectiveness and suggesting how that effectiveness comes about.
Only one of the teachers found the models generally unhelpful: her reflection was quite
different to the other seven and gave me considerable food for thought. I do not have the space
here to address the specific difficulties which this student experienced but, in my opinion, those
difficulties do not detract from the responses of the remaining seven students to the models. For
my present purposes I will focus on the participants experiences of one particular model in my
sequence. This is a version of Gentner and Gentners (1983) moving crowds model, which I
refer to as the Smarties model. I employ this model specifically to address points 5, 6 and 7 in
the above list of conceptual elements. Five of the teachers (Teachers B, C, D, E, F) refer
specifically to this model, and the relevant sections from their reflections are presented below
in full: they have been minimally edited for clarity and brevity. Teachers A and G do not
specifically mention the Smarties model in their reflections: they chose to focus on other
models in my sequence. However, a short extract from Teacher As reflection is included here
to help set the scene for the narratives the others present. Teacher G does refer to the
discussion, which occurs when the Smarties model (and consequently the electron transport
model) breaks down, and this section is also included below.

The Teachers/Learners Voices


Teacher A
Until the workshop I dont think I had ever really considered how the electrons actually
moved to produce the current at an atomic level. When I did think about it early in the first
workshop, I imagined that one individual electron had to go right around the circuit before
a current was produced and that this electron was followed in turn by many other electrons.
In our workshop the initial activity was to make a globe light up using a battery, globe,
wires and wire strippers. This started us thinking about electric circuits. We then recorded
on the board questions people had about the circuit we had made. This activity made me
think about what I thought happened when electric current flows in a circuit. I realized I
couldnt answer many of the questions posed. As we talked through the questions posed I
discovered that my view of an electric current was incorrect on an atomic level. After the
first workshop I was really confused. ...
Teacher B
... The main key terms and concepts I will focus on are: voltage, current and charge. The
experience that really made the concepts real for me was the role play that we did in one of

Res Sci Educ (2008) 38:529544

537

(the workshop) sessions. The people walking around the circuit were the charges
(electrons). The Smarties represented the energy that the electrons collected at the battery
and then gave to the light globe. The first time we did the role play each electron had one
Smartie. We then did it again to show that the amount of energy each electron has can
change (this time each electron carried two Smarties). In real life this would be due to a
higher voltage battery being in the circuit. The demonstration clearly illustrated for me the
difference between voltage and current. Previously my understanding of these two concepts
was very simple and clinical. Rather than an understanding it was probably more a
definition that I had remembered.
... The largest conceptual change I experienced ... was related to voltage. Prior to (the
workshop) session on electric circuits I had always thought that the voltage was a measure
of the push from the battery that is driving the electrons around the circuit. The higher the
voltage is, the higher the pressure of the push from the battery (or source of energy for the
circuit). According to (one textbook): The battery acts like a pump, forcing the electric
current around the circuit. The amount of pumping force is measured in volts. This
electrical push is called the voltage of the battery. The Year 7 textbook keeps the
explanation simple and I had never learnt (or maybe I learnt it, but did not retain) the more
complicated scenario of the electrons carrying energy around the circuit. ... (One text)
didnt even define current or voltage but regularly used these words in their discussion of
how transformers work, where electricity comes from (power station) and how generators
work. Students in Year 10 using this textbook would be completely lost with these difficult
concepts without an explanation and some revision of the concepts current and voltage.
... Role playing an electric circuit was a very useful and meaningful way to demonstrate what
was happening. Electricity is a difficult concept to understand because you cannot directly
observe the electrons moving. By acting out the scenario I had a real, concrete experience I
could turn back to. In the past I have just tried to remember definitions when learning this unit.
... I have previously underestimated the learning that can take place while acting things out. This
is a strategy that I will be able to take into my classroom and adapt to suit various units.
Teacher C
Electric circuits used to be the part of (the) electricity course that I quickly glossed over. I
considered that if I was able to get a year eight student to connect globes up in series and
parallel then I was doing a good job. I really knew that I was not teaching the concepts well
but I was teaching the things I was comfortable with.
Then ... I was a part of (the) electric circuit/Smartie dramatic performance. I can honestly
say that being an electron picking up and dropping off energy/Smarties has changed my
understanding of certain beliefs that I once held.
Resistance My understanding before was that resistance was some figure that you
calculated using a fairly simple formula but the number didnt really mean anything to
me. ... After walking through the circuit I now can see (a vision in my mind) that as
electrons move through a circuit then there are going to be variations in the rate at which
they flow because of the different materials that they have to flow through. I can also
visualise the difference between having the appliances connected in series or parallel and
how much quicker the electrons can flow in a parallel circuit because they dont all have to
take the same path. Now I would explain resistance in terms of the material which the
electrons have to flow through and that some materials are easier to flow through than
others, depending on the width of the wire and the properties of the material.

538

Res Sci Educ (2008) 38:529544

Volts and amps Again my previous understanding was that these two units were measuring
something electrical but what I was unsure of, and if I was asked the difference between the
two, I really had no idea. An amp is the measurement of the number of electrons passing a
particular point at any time and a volt is the amount of energy that each electron carries. I
was able to remember these statements from our experiences in class but again they were
just learnt by rote. ... After walking through the electric circuit I now know the difference
between the two and something about what they are measuring. ...

Teacher D
I think that the greatest change in my understanding is the concept of electron movement in
a circuit. My teaching of circuits has always been for the students to play with the
equipment and see how they can get the globe to work. I have been unable to further this
and put an adequate explanation as to why the globe lights up in certain set-ups and doesnt
for others. I end up talking about electrons flowing around, but due to the students (being
unable) to see this occurring, they dont establish a concrete understanding, and their test
results reflect this. The students end up with a half-understanding and cant explain what is
occurring with the electrons.
The hands-on activities we participated in during the workshop helped shape my
conceptual understanding and give me a way to better teach electron movement in circuits.
The use of Smarties is brilliant! I could visualize the electron movement, along with the
globe, battery and resistance. The concept of resistance had never been mentioned in my
class, as I didnt know how to bring it into the mix. Now I have a way that allows visual
learners to see it all occurring so that when I explain the details, such as the filament being
thinner causing more resistance, the students will have the visual representation in their
minds and not just hear my voice droning on! ...
The workshops have made me understand how much I rely on creating a visual image of
concepts in my head to help make things concrete. ...
Teacher E
Why does a light globe light up? An easy question, or is it? Sitting in (the) workshop I
realised that this was not something I had really thought about. I had just taken it for
granted that when you turn on the light switch the light globe will shine. If I had of thought
about it in any way, my answer was simple and clear, based on my understanding of year 7
electric circuits. My answer would have been as follows: when you turn on a light switch
you complete an electric circuit, thus allowing electricity to flow along the wires and give
the light globe the energy to shine. Simple isnt it. But wait, a problem, what is electricity
and what energy is it giving to the light globe? Pondering my prior knowledge of
electricity I attempted to answer this question.
... (The) energy carrying ability of electrons was further highlighted as (our) workshop
group modelled an electric circuit. ... I think I understand! The light globe shines as it gets
energy from the electrons passing through it. My prior knowledge is kicking in again.
The electrons use some energy to pass through the wires, they do not give up all their
energy at the light globe, as they need some energy to return to the battery or power
source in order to complete their circuit. The electrons give up most of their energy at the
light globe because there is a greater resistance as the electrons pass through the
filament of the globe.

Res Sci Educ (2008) 38:529544

539

... Physical models and analogies can be useful in developing understanding but they
need to have explanatory power for the individual concerned and not all models or
analogies will work for all students. Again the analogies used must coincide with a students
past experience of the world.
Teacher F
... One of the main reasons I have difficulty with this topic is its abstract nature. Since one
cant actually see what is happening inside the circuit you have to accept a lot of concepts
on faith! Another problem I have is that some of the concepts dont make any sense to me,
they dont fit logically with what I already know, my prior knowledge. ...
I now accept that when a circuit is closed an electric field is created because of the
positive and negative ends of the battery. This leads to an instantaneous force on all of the
electrons within the circuit: they all feel the push/pull generated and so gain energy. This
was illustrated by getting us to drop our Smarties as soon as the circuit was broken and
quickly pick them up again on closing the switch. As I dont fully understand electric fields
I prefer the explanation of the class member that described the transfer of energy occurring
when the electron pushed out of the battery knocks into the electron next to it which knocks
into the electron next to it setting up a chain reaction around the circuit. It is clear from this
explanation that the initial electron does not have to pass through the light globe for it to
glow. ...
The topic of electricity is a complex one that cannot be learnt by just reading a page of
text in a book. The use of carefully thought out models to promote discussion and help
explain the abstract concepts involved is a valuable tool for any teacher. It enables students
to challenge their own beliefs in a way no page of text ever can, especially ones in the
physics texts Ive read!
Teacher G
Before attending (this workshop) I had always understood that electric current was the
movement of negative charge through the circuit. However, I did not understand the term
conventional current at all. The general definition given by most textbooks describes
conventional current as the movement of positive charge around a circuit, which did not
make sense to me at all. I was under the impression that the electrons released from a
battery themselves had to travel around the circuit. Each electron had a certain amount of
energy that it carried with it.
After the explanation during (the workshop), I came to realise that the electrons
themselves do not move but rather it works as a domino effect: i.e. the electron jumps to its
nearest ion then that ion then becomes crowded and must release an electron of its own
which jumps to the next electron etc. This is how the charge travels along the wire. ...

Applying the Criteria for a Teaching Model to my Experiences of Teaching


Electric Circuits
The teachers reflections show how the criteria posited above as applying to a teaching
model, play out through the Smarties model. In the initial activity participants were given a
battery, torch globe and some lengths of wire, and asked to make the globe light. They were
asked to explain why some connections work and others do not, and then to explain why

540

Res Sci Educ (2008) 38:529544

the filament of the globe glows, but the connecting wires do not. This sets up a puzzle, so
that the explanation provided by the Smarties model can eventually be seen as fruitful
(criterion 1). Teachers A, D and E specifically recognise this to be important.
The entities in the model are concrete, and this appears to free the participants to think
about what is happening in the circuit in their own terms, making crucial connections with
other elements of their knowledge (criterion 2). By contrast the abstract, mainly
mathematical, interpretations they had previously encountered seem to have paralysed
their thinking, both about the concepts per se and about how they might teach them
(Teachers B, C, D, F). In addition, there is something about having participants move
around and physically identify with entities in the model that makes it a powerful tool for
thinking. The questions come as the participants actually enact the circuit. And of course
the Smarties themselves invite engagement! These features, in my experience, make the
Smarties model much more effective pedagogically than conceptually equivalent models of
coal trucks and shopping trolleys. My impression, supported by these teachers comments,
is that the model enables the rich classroom discourse that should be a major consideration
in selecting a teaching model (criterion 4). In the process, the crucial distinction between
the moving charged particles (current) and the energy they distribute around the circuit (emf
and potential difference) is clearly made (criterion 3), addressing the fifth and sixth
conceptual elements in my list above (Teachers B, C and E).
These teachers were not at all concerned that no entities in the real world could be said
to correspond with either people or Smarties. This underlines for me the appropriateness of
having different criteria for judging the effectiveness of a teaching model as compared with
consensus models. The teachers also had no difficulty with the way I invoked several
different models in the course of our attempts to construct a meaningful explanation of the
circuit. Furthermore, as Teacher F confirms, the participants suggested several alternative
analogies when I invited them to do so. Thus they did appear to be able to employ a range
of models flexibly and appropriately according to the problem at hand (criterion 6).
The Smarties model is also intended to begin to address the seventh of the conceptual
element in my list above, but it is here that the model breaks down and, paradoxically, it is
here that the model becomes most powerful. The model breaks down initially when it
attempts to account for how an electron knows that there are two globes in a series circuit.
I deal with this problem by saying, for the moment, that they just do. However, I later
introduce the contradiction that the electron drift speed is very slow, and yet a globe glows
as soon as the circuit is made. We then discuss how we might model the transport of energy
in ways that do not depend on the electrons actually moving from the battery to the light
globe. The realisation that energy transfer does not depend on the physical movement of
electrons is specifically highlighted by Teachers A and G, and Teacher F mentions some of
the groups suggestions for modelling such an energy transfer. I introduce the idea that
when the circuit is completed an electric field is set up in the conducting wires, and the
varying strength of the field determines how the electrons distribute their energy around the
circuit. This electric field model can then be used to address the electron knows question.
I say that the Smarties model is most powerful at the point when it breaks down for
several reasons. Firstly this makes the point that all analogies have their limits (criterion 6).
As Teacher E points out, it is critically important that learners appreciate this point.
Secondly, it introduces the electric field model, the consensus model that, arguably, we
want students to move towards (criterion 6) (Gunstone et al. 2007; Stocklmayer and
Treagust 1996). Thirdly, it makes it clear why the field model is needed (criterion 1). While
Teacher F exercises her prerogative of declining to accept the field explanation, on the
grounds that it is too abstract to be meaningful to her, the discussion has clearly empowered

Res Sci Educ (2008) 38:529544

541

her to construct a model which affords her an explanation that she does find adequate and
acceptable.
Fourthly, in moving to the field model, a switch in ontologies is achieved: instead of having
volitional persons deciding how many Smarties to deposit at different points in the circuit, we
have electrons interacting deterministically with the electric field. In other words, we have an
opportunity to address the ontological basis of scientific models, to explain how meaning is
created in science, why meanings are created in that way, and how the conditions for creating
meaning differ from possible alternatives (criterion 6). This point does not come through
immediately in the teachers reflections, probably because, at the time, I did not recognise and
capitalise on the opportunity here. Its significance has become apparent to me only as I have
subsequently engaged with the teachers accounts of their learning.

Recommendations for Teaching Electric Circuits


While there might well be agreement among physics educators that a field model is the
consensus model we want students to eventually adopt, the particular stage in learning at
which it might be reasonable to expect students to engage with the field model is a matter of
contention. Stocklmayer and Treagust (1996, p 177) argue that it should be adopted from
the earliest stages of learning, but I maintain that the electric field is a sophisticated and
highly abstract concept, and as such is necessarily inaccessible to beginning students. I
would further contend that this problem is not solved by simply postponing the study of
electric circuits until readiness for a field model can be assumed. I suggest that, for all its
inevitable shortcomings, the electron-transport model remains the most appropriate starting
point for beginning students of any age.
On the basis of my own experience, judged in the light of the criteria I have drawn from
the literature, I suggest that it is possible to employ concrete analogues of the electrontransport model to develop a qualitative understanding of dc circuits, encompassing the
first six conceptual elements in my list above. I suspect that discriminating clearly between
electrical charge and electrical energy will, initially, inevitably entail modelling energy as a
consumable substance (whether Smarties, or coal in coal trucks). I maintain that the point is
so important that such shortcomings in the models must be accepted and used knowingly.
At some later stage it must be explicitly acknowledged that the electron-transport model
cannot fully address the seventh conceptual element of the above list, and consequently
cannot address the electron knows problem, or give even an elementary account of ac
circuits. I contend that this explicit recognition should form part of the understanding we
intend developing students to form.
Whatever the pedagogical merits and limitations of preferred models of explanation,
the most critical issue is that all analogies break down under critical scrutiny and
therefore the focus in pedagogy should be less concerned with the search for the holy
grail of analogies to explain the phenomenon of electricity than it is with the reasoning
that such analogies generate when they break down at critical points in explanation.
(Heywood 2002, p 239)
At this point the bicycle chain model (for example Nuffield Primary Science 1993, p 81)
or other conceptually equivalent model (for example: rope model, Lofts et al. 2004, p 307)
becomes meaningful and fruitful, but I consider that such models should never replace the
Smarties-type models in the early stages of learning, despite the potential problems caused
by modelling energy as a consumable substance. This is because, as with the water-flow

542

Res Sci Educ (2008) 38:529544

model, there is a danger that the energy is too closely intertwined with the carrier for
the crucial conceptual distinction to be clear. Eventually students can be introduced to the
electric field model, but only once the need for it has been established, and only if the
students are ready to engage with such a highly abstract construct. The teacher who claims
her right to reject the field model alerts us to the dangers of imposing abstractions on
learners who are not yet ready to accept them.
I suggest that identifying the different models by explicitly naming them (Smarties
model, bicycle chain model, electron-transport model and field model) would help to focus
on the transition from one model to another, and what makes such a transition necessary at
particular stages in the reasoning. In the case of models for light the influential PSSC series
resulted in the terms particle model and wave model becoming part of the lexicon of
physics education, so we do have a precedent.

Using Teaching Models Knowingly


This exploration of the role of models in teaching electric circuits has highlighted for me
the fact that the consensus models of science do not necessarily make good teaching
models. Beyond this, the deeper engagement with these models that my teacher participants
opened up, showed me that models which violate the ontology of science provide an
opportunity for addressing significant questions about how science creates meanings, as
well as the strengths and limitations inherent in those meanings.
I have long shared the concern of many science educators that students neither construct
a robust and meaningful understanding of the concepts of science, nor find their learning
experiences interesting and enjoyable. Through my work with these teachers I have begun
to realise that, while we have directed considerable attention to the conceptual difficulties
that students experience, we may not have paid sufficient attention to their grasp of the
ontological basis of scientific explanations. When students say they find the explanations of
science meaningless, perhaps they are alerting us that the problem is one of meaning, not
simply understanding. Their everyday world is meaningful because at some intuitive level
they grasp the terms in which everyday meanings are created. In the science classroom,
they understand that their ways of creating ordinary meanings seem not to work, but they
do not understand how, or in what terms meanings can be created.
I have already noted that the kinetic particle model of matter is an example of a
consensus model that works reasonably well as a teaching model. However, my learning
from these and other teachers I have worked with, suggests that the relative ease with which
we have appropriated this model for teaching purposes conceals more fundamental issues,
issues that are actually more readily perceived in the problematic teaching context of
electric circuits. Thus, in teaching the kinetic particle model of matter, we teach students
that air consists of tiny particles that are in random motion at high speed, and that between
the particles is totally empty space. Yet we rarely address the key ontological question of
why one particular kind of explanation of matter is acceptable, whereas an explanation in
terms of, for example, nature abhorring a vacuum or a vacuum sucking particles is
now regarded as unacceptable. We do not consider why scientists originally asked such
questions about the nature of matter, let alone what contemporary purpose we might have
for wanting to consider how matter is structured. Neither do we generally consider the
intellectual struggle, which took place over a period of more than 150 years from the time
of Robert Boyle, and eventually resulted in the scientific community accepting the notion of
entirely empty space between the particles of air. Many students today still find this notion

Res Sci Educ (2008) 38:529544

543

preposterously unacceptable: perhaps addressing these questions would help them gain a
truer perspective on their difficulties.
It is part of our task as members of the scientific community to help our students
understand and use the consensus models of science. It is also part of our task as educators
to use the knowledge of our discipline to understand what makes a good teaching model
effective, and to recognise that consensus models do not always make good teaching
models. Going beyond this, we can use models to help students to see into the ontological
heart of science, to appreciate how science differs from other ways of imagining and
understanding the world, and to recognise that those fundamental assumptions are the
source of both the weaknesses and the strengths of our subject.

References
Duit, R. (1991). On the role of analogies and metaphors in learning science. Science Education, 75(6), 649672.
Feynmann, R., et al (1964) The Feynmann Lectures in Physics, Volume II - The Electromagnetic Field.
Bombay, India: Addison-Wesley.
Gentner, D., & Gentner, D. R. (1983). Flowing waters or teeming crowds: Mental models of electricity. In
D. Gentner & A. L. Stevens (Eds.), Mental models (pp. 99129). Hillsdale, NJ: Erlbaum.
Gianello, L., Dick, B., & McClintock Collective (1988). Getting into gear: Gender inclusive teaching
strategies in science. Canberra, ACT: Curriculum Development Centre.
Gilbert, J. K., Boulter, C., & Rutherford, M. (1998a). Models in explanations, part 1: Horses for courses?
International Journal of Science Education, 20(1), 8397.
Gilbert, J. K., Boulter, C., & Rutherford, M. (1998b). Models in explanations, part 2: Whose voice, whose
ears? International Journal of Science Education, 20(2), 187203.
Gunstone, R., McKittrick, B., & Mulhall, P. (2007). Physics teachers perceptions of the difficulty of teaching of
electricity. In press.
Halliday, D., & Resnick R. (1988). Fundamentals of physics (3rd ed.). New York: Wiley.
Halliday, D., Resnick, R., & Walker, J. (1997). Fundamentals of physics (5th ed.). New York: Wiley.
Harrison, A. G., & Treagust, D. F. (1998). Modelling in science lessons: Are there better ways to learn with
models? School Science and Mathematics, 98(8), 420429.
Harrison, A. G., & Treagust, D. F. (2000). A typology of school science models. International Journal of
Science Education, 22(9), 10111026.
Hewitt, P. (1987). Conceptual physics. Menlo Park, CA: Addison-Wesley.
Heywood, D. (2002). The place of analogies in science education. Cambridge Journal of Education, 32(2),
233246.
Kane, J., & Sternheim, M. (1988). Physics (3rd ed.). New York: Wiley.
Leach, J., & Scott, P. (2003). Individual and sociocultural views of learning in science education. Science &
Education, 12(1), 91113.
Lemke, J. L. (1990). Talking science: Language, learning, and values. Norwood, NJ: Ablex.
Lofts, G., & Evergreen, M. (2007). Science quest 3 (3rd ed.). Milton, Queensland: Jacaranda Wiley.
Lofts, G., et al. (2004). Jacaranda physics 1. Milton, Queensland: Jacaranda Wiley.
McClintock Collective, & various authors. (1989). Classroom practice. Australian Science Teachers Journal
(Special Issue on Gender Inclusive Science Teaching), 35(3), 5074.
Mortimer, E., & Scott, P. (2003). Meaning making in secondary science classrooms. Maidenhead, Berkshire:
Open University Press.
Mulhall, P., McKittrick, B., & Gunstone, R. (2001). A perspective on the resolution of confusions in the
teaching of electricity. Research in Science Education, 31, 575587.
Nardelli, D. (2006). Science alive 2. Milton, Queensland: Jacaranda Wiley.
New South Wales Board of Studies. (2002). Physics stage 6 syllabus. Sydney, NSW: NSW Board of Studies.
New South Wales Board of Studies. (2003). Syllabus for science years 710. Sydney, NSW: NSW Board of
Studies.
Nuffield Chelsea Curriculum Trust. (1993). Nuffield primary science: Electricity and magnetism key stage 2
teachers guide. London: Collins Educational for Nuffield Chelsea Curriculum Trust.
Ohanian, H. (1994). Principles of physics. New York: Norton.
Serway, R., & Faughn, J. (2000). College physics (5th ed.). Orlando, FL: Saunders/Harcourt.

544

Res Sci Educ (2008) 38:529544

Spencer, R. (1999, 13/4/1999). A ridiculously brief history of electricity and magnetism, mostly from E. T.
Whittakers A History of the Theories of Aether and Electricity. Available: http://maxwell.byu.edu/
~spencerr/phys442/node4.html.
Stannard, P., & Williamson, K. (2006). Science world 9 (3rd ed.). South Yarra, VIC: MacMillan.
Stocklmayer, S. M., & Treagust, D. F. (1994). A historical analysis of electric currents in textbooks: A
century of influence on physics education. Science and Education, 3, 131154.
Stocklmayer, S. M., & Treagust, D. F. (1996). Images of electricity: How do novices and experts model
electric current? International Journal of Science Education, 18(2), 163178.
Storen, R., & Martine, R. (2004). Nelson physics units 1 and 2 (3rd ed.). Southbank, VIC: Thomson/Nelson.
Strike, K. A., & Posner, G. J. (1985). A conceptual change view of learning and understanding. In L. West &
L. Pines (Eds.), Cognitive structure and conceptual change (pp. 211231). Orlando, FL: Academic.
Treagust, D. F., Duit, R., Joslin, P., & Lindauer, I. (1992). Science teachers use of analogies: Observations
from classroom practice. International Journal of Science Education, 14(4), 413422.
van Driel, J., Verloop, N., & de Vos, W. (1998). Developing science teachers pedagogical content
knowledge. Journal of Research in Science Teaching, 35(6), 673695.
Victorian Curriculum and Assessment Authority. (2003). Physics Victorian certificate of education study
design. East Melbourne, Victoria: Victorian Curriculum and Assessment Authority.
Victorian Curriculum and Assessment Authority. (2005). Victorian essential learning standards: Science.
East Melbourne, Victoria: Victorian Curriculum and Assessment Authority.
Weicek, C., Zealey, B., et al. (2005). Physics in context The forces of life preliminary (2nd ed.). South
Melbourne, Victoria: Oxford University Press.
Wells, G. (1999). Dialogic inquiry: Toward a sociocultural practice and theory of education. Cambridge:
Cambridge University Press.
Wertheim, M. (1995). Pythagoras trousers: God, physics, and the gender wars. New York: Times Books/
Random House.

You might also like