Sei sulla pagina 1di 30

Sci & Educ

DOI 10.1007/s11191-013-9592-7

Teaching Energy Concepts by Working on Themes


of Cultural and Environmental Value

Ugo Besson • Anna De Ambrosis

 Springer Science+Business Media Dordrecht 2013

Abstract Energy is a central topic in physics and a key concept for understanding the
physical, biological and technological worlds. It is a complex topic with multiple con-
nections with different areas of science and with social, environmental and philosophical
issues. In this paper we discuss some aspects of the teaching and learning of the energy
concept, and report results of research on this issue. To immerse science teaching into the
context of scientific culture and of the students’ cultural world, we propose to select specific
driving issues that promote motivation for the construction of science concepts and models.
We describe the design and evaluation of a teaching learning path developed around the
issue of greenhouse effect and global warming. The experimentation with high school
students has shown that the approach based on driving issues promotes students’ engage-
ment toward a deeper understanding of the topic and favours further insight. The evolution
of students’ answers indicates a progressively more correct and appropriate use of the
concepts of heat, radiation, temperature, internal energy, a distinction between thermal
equilibrium and stationary non equilibrium conditions, and a better understanding of
greenhouse effect. Based on the results of the experimentation and in collaboration with the
teachers involved, new materials for the students have been prepared and a new cycle of
implementation, evaluation and refinement has been activated with a larger group of
teachers and students. This type of systematic and long term collaboration with teachers can
help to fill the gap between the science education research and the actual school practice.

1 Introduction

A considerable amount of educational research has been devoted to the teaching and
learning of energy concepts and phenomena. Many studies have pointed out students’
common conceptions that can create learning difficulties and different approaches for

U. Besson (&)  A. De Ambrosis


Department of Physics, University of Pavia, Via A. Bassi 6, 27100 Pavia, Italy
e-mail: ugo.besson@unipv.it
A. De Ambrosis
e-mail: anna.deambrosisvigna@unipv.it

123
U. Besson, A. De Ambrosis

teaching energy have been designed and experimented. In this paper we briefly review the
literature on students’ conceptions on energy, summarize some approaches to teaching
energy proposed by science education research, and then we propose an approach aimed at
integrating the Science Technology Society Environment (STSE) approach with the con-
ceptual and procedural dimensions of science learning. In this perspective we have
developed a teaching learning path, devoted to high school students, around the problem of
understanding the physical basis of greenhouse effect and global warming. We wanted to
strictly connect the environmental aspects and the scientific content, and we paid particular
attention to the conceptual progression and connections with basic energy concepts: dif-
ferentiating the concepts of work, heat, internal energy, temperature; considering the role
of radiation in thermal phenomena; understanding energy conservation and energy bal-
ances in stationary situations of thermal non-equilibrium.
The sequence has been tested in six high school classes, four with 17–18 year old
students and two with 15–16 year old students for a total of 121 students. We investigated
two research questions: (1) how can the study of a complex issue such as greenhouse effect
and global warming improve understanding of energy concepts and (2) what type of
materials, experiments, models and schematic representations can favor students’ under-
standing of this topic. In the present paper we describe the main features of the learning
sequence, and the methods, data sources and results of its implementation by a group of
teachers.

2 Students’ Conceptions and Conceptual Difficulties on Energy

Many researches have been devoted, since the eighties, to the teaching and learning of
energy concepts and phenomena and to students’ common conceptions.1 Research shows
that anthropocentric and vitalistic conceptions are prevalent in youngest pupils: energy is
connected to life, to movement and to the capability of doing actions, while it is not related
to non-living and motionless objects, except for objects explicitly devoted to storing and
supplying energy, like batteries and fuels. The idea that energy can transform in different
forms, largely conveyed by textbooks, is often meant as a kind of metamorphosis, like in
fairy tales, rather than a conservation principle of a measurable physical quantity.
Some researchers (Watts 1983; Viennot 2001; Goldring and Osborne 1994) point out
that students merge in an undifferentiated notion the concepts of energy, force and
momentum, using them almost as synonymous. Actually, an undifferentiated idea of force-
energy has been used by scientists for a long time, while a clear distinction of the physical
meaning of the two terms stemmed only in the second half of the nineteenth century.
Helmholtz, in 1847, entitled On the Conservation of Force (Krakt in German) his famous
work where he established a general energy conservation principle and the expression vis
viva (living force), introduced the first time by Leibniz, was used in high school textbooks
as synonym of kinetic energy until few decades ago.
For high school and college students, research shows the tendency to consider energy as
something producing actions and effects and thus consuming itself, rather than to use
energy conservation and degradation to explain phenomena.2 Students show difficulties in

1
See for example (Shayer and Wylam 1981; Solomon 1983; Watts 1983; Erickson and Tiberghien 1985;
Duit 1986).
2
See for example (Goldring and Osborne 1994; Loverude et al. 2002; Meltzer 2004; Rozier and Viennot
1991).

123
Teaching Energy Concepts

distinguishing between extensive and intensive quantities, in particular between heat and
temperature: they often use a mixed temperature-heat notion, or consider the temperature
of an object as a measure of the level of heat. Students speak about ‘‘heat contained in a
body’’ as a substance that can pass from a body to another, in a way that remembers the
ancient conception of caloric, and confuse heat with internal energy. Sometimes both heat
and cold are considered as substances that can be transferred from a body to another one.
Work is generally not connected to temperature changes, while the prevalent idea is that
only heat exchanges can increase or decrease the temperature of an object: this idea
survives also in university students and can be found among science teachers too.
Some of the students’ difficulties with the idea of heat derive from the difference
between its meaning in common language and in the language of science. This concept is
often used as synonymous of internal energy or at least of thermal energy, i.e., of the part
of internal energy connected to the changes of temperature. Difficulties in differentiating
the meaning of heat, work and internal energy hinder the understanding of the first law of
thermodynamics and in general of the energy conservation principle. The language gen-
erally used in textbooks does not clarify the ambiguity of common language because
expressions like heat supplied convey the idea that a body possesses heat in order to give
it. These remainders of the old conception of heat as fluid, not explicitly addressed, are
critical from the educational point of view because they can prevent a correct under-
standing of these concepts.
On the other hand, any distinction between heat and work disappears at the microscopic
level: interactions are of the same type, the difference is in their coherence or incoherence
and appears only at macroscopic or mesoscopic level where a large number of molecules
are involved (Besson 2003). For that reason, expressions like mechanical work and thermal
work could be properly used.
To overcome linguistic ambiguity one should abandon the words ‘‘heat’’ and ‘‘work’’
and replace them with expressions conceptually clearer like mechanical transfer of energy
and thermal transfer of energy. As Romer (2001) wrote
If you want to think up a good noun for ‘energy transferred by virtue of a temperature difference’ that
would be fine with me. Call it Harry, call it Quincy, anything except heat’’. Try it out on me next
time you have occasion to submit a thermodynamic AJP paper.
Learning progression of energy concepts across middle school grades has been studied by
Hee-Sun Lee and Ou Lidya Liu (Lee and Liu 2010). Based on the results obtained with a
sample of 2,688 middle school students in different schools and in different states, the
authors conclude that ‘‘students’ overall knowledge integration levels with energy concepts
are mediocre, that advanced energy concepts such as conservation are more difficult than
identifying energy sources and that the origin of this difficulty is in part related to the
increased demand for integrating many scientifically relevant ideas’’.
Understanding the energy conservation principle requires teachers and students to
differentiate the concepts of work, heat, internal energy, temperature, but also to clarify the
role of radiation. Energy transfer by electromagnetic radiation must be considered as work,
heat or as a third specific modality? There is no universal agreement in the literature on this
point. Some authors consider three types of energy transfer: work, heat and radiation. For
example, Solbes et al. (2009), underline that ‘‘energy variations may take place, not only
through work or heat, but also by means of radiation exchange processes’’ (see also
Domenech et al. 2007). On the contrary, other authors (the majority) consider only two
types of energy transfer, i.e., work and heat, and classify the transfer of radiation energy as
work or as heat according to the situation. For example, Zemanski (1957) wrote:

123
U. Besson, A. De Ambrosis

The gain or loss of internal energy, equal to the difference between the energy of the thermal
radiation which is absorbed and that which is radiated, is called heat.
Baierlein (1999, p. 18) simply affirmed:
Energy that is being transferred by conduction or radiation may be called ‘heat’. That is a technically
correct use of the word and, indeed, a correct use as a noun.
In the history of physics the terms radiant heat have been used for a long time even if
Maxwell wrote:
The phrases radiation of heat and radiant heat are not quite scientifically correct, and must be used
with caution. Heat is certainly communicated from one body to another by a process which we call
radiation. We have no right, however, to speak of this process of radiation as heat… when we speak
of radiant heat we do not mean to imply the existence of a new kind of heat but to consider radiation
in its thermal aspect. (Maxwell 1871, pp. 15–16)
Often in textbooks radiation is presented as a way of heat transmission, and this implicitly
suggests the idea that radiation and heat have the same characteristics, thus contributing to
students’ difficulties. The historical development of the idea of radiant heat and the study
of its characteristics compared to the properties of light shows how the process of
differentiation was a long one and required both experimental and theoretical efforts
(Besson 2012). This suggests that it is worthwhile to explicitly address the problem with
the students to promote a good understanding of the energy conservation principle and of
the first law of thermodynamics. More specifically, we think that from an educational point
of view it is worthwhile to distinguish between thermal conduction and radiation as two
ways of transferring energy. According to the terminology suggested above, we could say
radiative transfer of energy to be added to mechanical and thermal transfer of energy.

3 What is Energy?

One of the problem in teaching energy is that students spontaneously ask the question
‘‘what is energy?’’, while for physicists this question has no simple answer or does not have
any answer at all, as Feynman wrote:
It is important to realize that in physics today, we have no knowledge of what energy is. We do not
have a picture that energy comes in little blobs of a definite amount. It is not that way. However, there
are formulas for calculating some numerical quantity, and when we add it all together it gives the
same number. It is an abstract thing in that it does not tell us the mechanism or the reasons for the
various formulas. (R. Feynman Lectures on Physics, Vol. I, p. 4-2).
The question can have a double meaning: one can ask about the essence of energy, or about
an operational and mathematic definition of energy as physical quantity.
As far as the first aspect is concerned, one could say that the question has nothing to do
with physics, as other similar questions about ‘‘what electricity is, or what time is?’’.
Nevertheless, the ontological character of energy is questionable. In the past both the
conception of energy as a kind of substance and as an abstract quantity, invented by
scientists to describe phenomena, were supported by different scientists. Attributing a
substance to physical quantities is a spontaneous tendency of common sense that can be
found sometimes also in scientists’ thinking (this is suggested for example by expressions
like ‘‘an electric charge q is moving in an electric field’’ or ‘‘let’s take a mass m’’ often
used in exercises and questionnaires). Energy, like a substance, is an extensive quantity
and satisfies a conservation principle. Nevertheless, it depends on the frame of reference
in the kinetic energy contribution, a property that does not fit well with the idea of

123
Teaching Energy Concepts

substance. Moreover it is additive only if the interaction fields are considered as consti-
tutive parts of the system. Unlike other extensive and conservative variables, such as
momentum, energy evokes images, suggestions and meaning far beyond the framework of
physics contents.3
Recently the New Age literature has adopted the term energy in its representations and
descriptions introducing the ideas of fluxes of positive or negative energy, of energetic
auras, energetic vibrations and so on. A physicist could distance her/himself from all these
polysemic implications as not pertinent to her/his work and not necessary for describing
physical phenomena or to solve problems. On the contrary, a teacher cannot ignore this
kind of images and suggestions that some students spontaneously associate to the concept
or to the word energy.
Even the second aspect of the question ‘‘what is energy?’’ poses problems not easy to
solve. Kinetic energy of a body or potential energy of a conservative force can be simply
defined, but it is difficult or even impossible do define energy in general, without any
adjective. Many textbooks in the past (also recent) defined energy as the ability to do work,
according to Maxwell’s definition:
The energy of a body may be defined as the capacity which it has of doing work, and is measured by
the quantity of work which it can do. (Maxwell 1871, p. 90)
But this contradicts the second law of thermodynamics and the theory of heat engines
(Lehrman 1973): two bodies at different temperatures can produce work if used
appropriately in a thermal engine, but they lose this capacity reaching thermal equilibrium
without losing energy. The ability to do work can be expressed by the free energy or by the
quantity exergy,4 (Ogborn 1986; Viglietta 1990). Hicks (1983) holds that the definition of
energy as capacity to do work ‘‘should not be used even as an initial definition, even with
remarks to its inadequacy’’.
The transformation of the definition in ability to produce changes, solicits the same
criticism as the previous one. As Ogborn (1986, p. 30) wrote:
Energy is not the ‘go’ of things… the possession of energy is not what drives, explains or account for
change… entropy or free energy is what decides if the change can happen.
Many researchers have underlined that the capacity of producing changes is related to the
existence of differences, of disequilibrium and that the variable measuring the distance
from equilibrium, and then the possibility of changes, is entropy (in mechanics the
tendency of potential energy toward a minimum must be considered too).
One can also say that the transfer of energy from a body to another or the transformation
from one form to another is a measurement of the change occurred [a similar definition is
given by Hecht (2007)]. Following a famous analogy proposed by Feynman in his Physics
Lectures, many scholars prefer a general definition of energy as a scalar quantity that is
conserved in any physical process and transformation. This is a definition in term of
properties, rather than a constructive and operational one, and implies that the conservation

3
As an example: ‘‘In this process… two electrons become pure energy because they annihilate with
positrons (anti-electrons)… As a spirit animating what exists around us, energy is in everything, sometimes
tangible as the Sun light, sometimes hidden in remote corner of reality… All bodies producing a field are
subjected to its action, thus all bodies, being no more than energy lumps, interact gravitationally the one
with the other’’ (translated by the authors from De Felice F. (2000), pp. 13, 23, 24).
4
The term was coined by Z. Rant in 1956, resuming some ideas introduced by Gibbs.

123
U. Besson, A. De Ambrosis

principle become the postulate of existence of energy itself: a type of ontological statement
very common in mathematics, but unusual in experimental sciences.
One can conclude that in teaching any simple and short definition of energy should be
avoided; in fact, like other physical quantities, such as temperature or mass, energy is a
quantity that cannot be defined by a unique sentence, it needs a progressive construction of
meaning and it can be really understood only after using it in different contexts and
problems.

4 Approaches to Teaching Energy Proposed in Science Education Research

Different teaching approaches to energy and related phenomena can be found in science
education research and in textbooks. We briefly present in the following some examples of
the themes and solutions proposed.
Some researchers propose a historical approach, developing a didactical reconstruction
of the historical development of the energy concept and conservation principle by showing
the process leading to the definition of the concepts as they are now, through debates,
changes, errors and progressive conceptual alignment. The historical reconstruction can be
mainly focused either on the development of the ideas and of the scientific and philo-
sophical debate (all changes but something remains constant, nothing is created nor
destroyed, the unity of nature…) or on the technological, economical and social issues
(typically, the birth of the industry, the measure of human work, the steam engines and the
industrial revolution).
For example, Coelho (2009) designed a path following some stages of the development
of the energy concept from Mayer, Joule and Thomson to Maxwell and Planck. The author
argues that interpreting, in the light of Mayer’s and Joule’s studies, the principle of energy
conservation as an equivalence principle could prevent the reference to energy as some-
thing that cannot be created neither destroyed, but only transformed and that, in this sense,
must be a real thing, a substance. The author remarks that many physics textbooks in the
last decades of the nineteenth century and at the beginning of the 20th used a formulation
of this type, and defined the first law of the thermodynamic as the ‘‘equivalence principle’’.
Baracca and Besson (1990) proposed a historical introduction of the energy concept
more centered on technological and social issues. The underlying idea is that the meaning
of the word ‘work’ in common language and in real life has to be taken into account rather
than disregarded. Therefore it is useful and necessary to develop a discussion and a
reflection on the historical origin and the motivations of the use of such word to indicate a
particular physical quantity. Starting from the economic need, in the newborn industry, of
measuring human work and of comparing it with the work made by animals and by
machines, the path presents the evolution of the definitions of work and mechanical energy
together with the problem of the optimization of the movement transmission (Smeaton,
Lazare Carnot). Subsequently, the problems related to the introduction of steam engines in
the industrial revolution are discussed and the theories of heat, the relationships between
heat and work and the idea of thermodynamic efficiency are presented.
A proposal of informal education based on a thorough historical analysis was realized
by Bevilacqua and Falomo (2010) in a historical interactive Laboratory on energy. The
laboratory presented the historical development of energy concept through the recon-
struction, the analysis and the interpretation of historical experiments, available to visitors
in an interactive exhibition open to students and teachers and to general public.

123
Teaching Energy Concepts

Other authors propose a STSE (Science Technology Society Environment) approach


focusing on problems related to the use of technology, the availability of energy resources,
the preservation of the environment, the social and economic issues, with a multidisci-
plinary perspective, often including activities involving the outside school context. For
example, Domenech et al. (2007) stressed the importance of taking into account the
relationships between science, technology, society and environment, and consider that
science education research on energy has given little attention to these aspects and to the
problem of students’ motivation. According to them, in students’ difficulties not only
conceptual, but also procedural and values aspects are involved and it is necessary to face
the problem in a global way, avoiding approaches dealing with single aspects of the energy
issue. They consider it necessary to show how the energy concept is used in different
scientific areas, and to encourage the students consider new problematic situations.
Concerning the conceptual sequence in introducing energy concepts, different solutions
can be found in handbooks and in research works.
Most high school physics textbooks, propose a gradual and progressive conceptual
sequence, following the usual separation of physics subjects in the curricula: at first, within
the topic of mechanics, are given the definitions of work, kinetic and potential energy and
subsequently the energy conservation is progressively extended to other physics phe-
nomena, so that the general concept of energy is constructed step by step. A variation in the
same frame is to start the construction of the energy concept from thermal phenomena, and
then to show the equivalence between heat and other quantities connected to energy.
On the other hand, a different path can be followed which presents from the outset a
general concept of energy as a scalar quantity obeying a law of conservation. In this case,
different situations are studied and compared concerning a variety of phenomena
(mechanical, electromagnetic, thermal, chemical, nuclear…), with examples of energy
balances in heat engines, electrical motors, power stations or other practical devices. Two
examples of this type of conceptual sequence, which can be defined a holistic approach to
energy, are given by Solbes et al. (2009) and Papadouris and Constantinou (2011).
Solbes et al. (2009) present the concepts of energy conservation and degradation as
cross-sectional ideas for all physics topics, and design and experiment a teaching sequence
with this objective for upper secondary school students.
Papadouris and Constantinou (2011) propose an approach to energy whose main
objective is developing the understanding of the nature of science (NOS) as an important
aspect of scientific learning. The energy concept is introduced as a ‘‘theoretical frame-
work’’ invented by scientists in order to describe the changes that happen in physical
systems. Within this framework the students are guided to gradually recognize the fun-
damental features of energy (transfer, conservation, degradation) and to use them in the
study of various situations. The didactic approach is based on a close correlation between
the evolution of the knowledge on the nature of science and the conceptual understanding,
with the hypothesis that these two aspects can support each other.

4.1 Our Approach

Observation of school practice, and the cooperation with in-service teachers, together with
specific research results, led us to the conviction that it is necessary to overcome a too de-
contextualized and technical approach to physics teaching. In particular, as far as energy
issues are concerned, it is necessary to immerse physics contents into the context of
scientific culture by discussing different interpretations and conceptions which caused
historical debates, sometimes not completely resolved, even today. This means to match

123
U. Besson, A. De Ambrosis

ideas and references of a STSE (Science Technology Society Environment) approach with
a special attention to conceptual, procedural and ethical dimensions of the scientific
learning.
For this aim we propose to select specific driving issues which can promote motivation
towards a progressive construction of physics concepts and models. This approach also
allows teachers and students to better clarify the scope and value of physics concepts and to
connect them to other disciplines and to the students’ cultural context. There is a wide
literature documenting the advantages offered by using driving questions and project-based
learning in promoting students’ motivation and understanding of fundamental aspect of
science (see for example Blumenfeld et al. 1991; Krajcik et al. 1998; and Edelson et al.
1999). Studies have been developed to design and test methodological and technological
supports favoring students’ engagement in investigations rich enough to promote moti-
vation and interest (Reiser et al. 2001; Reiser 2004; Sandoval and Reiser 2004).
In this line we have developed a teaching–learning path, devoted to high school stu-
dents, addressing the specific problem of the greenhouse effect and global warming and
aimed to develop in this context a better understanding of energy and energy conservation
concepts (Besson et al. 2010a).
Anomalous greenhouse effect and global warming are central issues in public debate
since the eighties (see IPCC 2007). Nevertheless, in addressing the global warming as a
socio-cultural issue the details of the physical aspects are very often neglected and the
physical basis of the phenomenon is treated in a hurried way to stress mainly the envi-
ronmental and economical aspects. So, the physical content and the general social issue
remain quite separate from each other and a true understanding of the phenomenon is not
favoured. Our aim is to tightly connect and match together the two aforementioned aspects,
thus providing the needed conceptual basis. The two aspects should interweave and support
each other so that the consciousness of the social relevance of the topic can provide
motivation to deepen the conceptual aspects and the conceptual understanding can supply
the necessary tools to follow the public debate critically, assume aware points of view,
develop knowledge autonomously and assume decisions.
The design of the teaching path is based on a ‘three-dimensional approach’ (Besson
et al. 2010b) which involves a synergic integration of three aspects: a critical analysis of
the scientific content in view of its importance for teaching, an overview of current
treatments (textbooks, common teaching) and an analysis of didactic research on the topic
(common conceptions of the topic and teaching learning sequences).

5 Students’ Conceptions and Common Learning Difficulties on Thermal Radiation


and Greenhouse Effect

Common conceptions on heat and temperature, phase changes, and thermal conduction
have been studied together with the development of these ideas after teaching.5 A number
of researches explored students’ conception on light, vision and colours, and learning
sequences on these topics have been produced.6
On the other hand, few researches have been carried out on students’ ideas about
thermal effects of radiation and the existing ones focus only on particular aspects. Redfors

5
See Stavy and Berkovitz (1980), Shayer and Wylam (1981), Erickson and Tiberghien (1985), Arnold and
Millar (1996), Cotignola et al. (2002), De Berg (2008).
6
See Guesne (1985), Andersson and Kärrqvist (1983), Kaminski (1989), and Chauvet (1996).

123
Teaching Energy Concepts

(2001) studied university students’ reasoning on the interaction between radiation and
metals involving atomic models and quantum theory. Some researchers investigated stu-
dents’ understanding of phenomena involving X-rays and radioactivity by focusing in
particular on interaction with living organisms and on risks for man and environment
(Lijnse et al. 1990; Millar 1994; Rego and Peralta 2006).
Other studies dealt with the understanding of the greenhouse effect in connection with
the problem of global warming of our planet7 and some researchers also propose ideas and
approaches for dealing with common students’ difficulties (Rye et al. 1997; Meadows and
Wiesenmayer 1999; Koulaidis and Christidou 1999).
It was found, for example that:
– the ozone layer depletion and radioactivity are often indicated as causes of global
warming and the skin cancer is considered as an effect of global warming;
– the idea of ‘trapping’ of Sun rays by atmosphere is used as explanation of the
greenhouse effect;
– the higher temperature inside the greenhouses is explained as the consequence of a
non-steady state in which more energy enters than exits.
The relationship between students’ science content knowledge on greenhouse effect and
global warming and their awareness of social activism has been studied at different age
levels and activities have been designed to favor students’ understanding of the connection
between personal energy use and global warming.8 These studies confirm the necessity of
carefully designed science curriculum to develop a coherent understanding of global
warming and help students make informed decisions about their personal lifestyle choices.
In order to focus on students’ conceptions on thermal effects of radiation and green-
house effect we carried out a preliminary investigation with groups of student teachers and
high school students. We gave a questionnaire to 51 student teachers and 121 high school
students. Data collected showed:
– a tendency to give absolute meaning to optical properties (transparency, absorptivity
and emissivity) as intrinsic characteristics of bodies;
– a lacking or incorrect consideration of infrared emission of bodies as a mechanism of
losing energy;
– a confusion between transitory phases, in which temperature changes, and steady state
situations;
– a difficulty in taking into account the systemic interdependence of all factors and
phenomena implied in energy budgets.
Concerning the greenhouse effect, only 6 % of student teachers and 9 % of high school
students gave an enough correct explanation in terms of solar radiation absorbed and of
thermal radiation emitted by the ground, 35 % proposed the idea of ‘trapping’ of sun rays
inside the greenhouse and 31–39 % reasoned only in terms of thermal isolation due to
atmosphere. The problem of Earth global warming was often confused with that one of
ozone layer depletion or ‘‘ozone hole’’ and its origin was generically attributed to pollu-
tion, injurious gases, deforestation, human heating.

7
See Boyes and Stanisstreet (1993), Rye et al. (1997), Groves and Pugh (1999), Koulaidis and Christidou
(1999), Meadows and Wiesenmayer (1999), Andersson and Wallin (2000), Papadimitriou (2004), Osterlind
(2005), Lester et al. (2006), Kilinc et al. (2008), Svihla and Linn (2012).
8
See for instance Lester et al. (2006), Cordero et al. (2008), Svihla and Linn (2012).

123
U. Besson, A. De Ambrosis

Fig. 1 Images from the web used by students. Left Some solar radiations are reflected by the Earth and by
the atmosphere. Right The increase of greenhouse gases in the atmosphere favours trapping of the reflected
energy of Sun radiation causing an increase of the temperature on the earth surface

Some of these misconceptions are favored by textbooks or web sources, which give
confuse or erroneous explanations, sometimes proposing the idea of trapping of sun
radiation or of thermal radiation, and describing energy fluxes without balance between
incoming and outgoing energy, thus confusing transitory and stationary conditions (see for
example Fig. 1).

6 Greenhouse Effect, Climate Change and Energy Concepts

Concepts and ideas related to energy are critical elements for understanding mechanisms of
greenhouse effect and global warming. Conversely, the problem of global climate change
can provide motivation stimuli, rich context and productive field of interest for pointing out
and clarifying concepts related to energy. Central points are the ideas of transmission and
transformation of energy, of energy conservation and balance, and the basic phenomena of
interaction between radiation and ordinary matter (reflection, absorption, transmission and
emission). At this purpose it is necessary to point out that when radiation meets ordinary
matter the energy is split up in energy of reflected, transmitted and absorbed radiation, and
a balance must exist between these energies and the energy of incoming radiation. All the
discourse and all explanations on this topic at school level can be handled in a phenom-
enological way by means of these concepts only.
A widespread shortcoming, which can be found not only among students but also in
textbooks and in didactic research articles, is the lack of consideration of radiation
emission and absorption by atmosphere and greenhouse gases and confusion between
absorption and reflection. The process of absorption of radiation (and the consequent
increase of internal thermal energy) is separated and independent of the emission of
thermal infrared radiation, which occurs always and for all bodies, independently of the

123
Teaching Energy Concepts

presence of incoming energy. They are two distinct phenomena and this is shown by the
fact that the emitted radiation is uncorrelated and different from the absorbed radiation. On
the contrary, reflected radiation is immediately produced when the object is touched by the
incoming radiation; it is of the same type (equal frequency) and correlated to the incoming
radiation. This distinction must be clearly discussed with students and attention must be
paid to the language used. It is misleading to speak of transformation of solar radiation into
Earth infrared radiation and of Earth infrared radiation transformed into atmosphere
infrared radiation or of re-emission of infrared radiation by the atmosphere. It is essential
to introduce the idea that the three processes of reflection, absorption and transmission
involve all objects in different proportion, and depend on the materials and on the spectrum
of the considered radiation (Besson 2009). It is necessary to stress that emission of radi-
ation occurs always for all objects in an amount and with a spectrum which depend on the
temperature and on the surface features.
For all these reasons the analogies between atmosphere (and greenhouse gases) and a
blanket and/or a shield are misleading (Svihla and Linn 2012). We think that it is more
useful to consider simple cases, for example the behaviour of one or more glass sheets, in
order to show by observations and experiments the presence of reflected, transmitted,
absorbed and emitted radiation, to well distinguish them, and to represent the related
energy balances by using opportune graphs. Subsequently, the glass example can be used
as model for more problematic cases, such as greenhouse gases, atmosphere, Earth surface,
clouds, ice, glaciers, oceans…
Moreover, a clear understanding is necessary of some key concepts and relationships
concerning thermal phenomena, a clear distinction among the quantities heat, temperature,
work, and internal energy, between process and state quantities, transient and stationary
conditions and thermal equilibrium. The study of global warming and climate change
issues offers a chance for introducing or for clarifying the above mentioned physics
concepts and reasoning patterns. In particular it is important to clarify that:
– The temperature of an object (system) is connected in a complex way to its internal
energy. Only a part of this last is strictly related (is proportional) to the temperature and
this part of internal energy is often called thermal energy.
– It is possible to increase the temperature of a system by increasing its internal energy.
This can be brought about both by doing work on it (heating without heat) or by giving
heat (heating by contact with a system at higher temperature).
– It is also possible to increase the internal energy of the system without increasing its
temperature, for instance by a phase change.
– In some processes a system may increase its temperature without receiving energy
from the outside (such as in chemical reactions or in dissipative movements).

7 Thermal Phenomena, Radiation Energy, and Greenhouse Effect: A Teaching


Learning Path

In this section we describe the activity sequence designed with the teachers. In the next one
we discuss the implementation by the teachers (Sect. 8.1), we summarize the results obtained
(Sect. 8.2), and present the new materials designed to improve the sequence (Sect. 8.3).
Greenhouse effect and climate change are complex topics involving many physical
properties and phenomena (Onorato et al. 2011). To obtain an effective learning, it is
necessary to proceed gradually, by focusing on each phenomenon involved, and to treat the

123
U. Besson, A. De Ambrosis

greenhouse effect as a synthesis of the multifaceted aspects previously studied. To promote


a stable and satisfactory understanding, it is not sufficient to simply explain at once what the
greenhouse effect is. It is necessary to realize a series of activities, which can provide a well
established background on which students can develop their own conceptual construction.
To this aim, it is important to activate a synergic integration between qualitative experi-
ments, theoretical systematizations, quantitative experiments and explicative models.
By taking into account students’ conceptions and difficulties mentioned in Sect. 2 and 5,
we aimed especially to reach the following cognitive objectives:
– to distinguish between concepts and phenomena initially confused in a global
undifferentiated notion (heat, visible radiation, infrared radiation);
– to separate properties of the objects and characteristics depending on interactions
(transparency, absorptivity, colour);
– to develop the concept of stationary condition in non-equilibrium state, this is crucial
for explaining many physical phenomena.
Preliminary work with small groups of students led us to specify a sequence of nec-
essary cognitive steps toward the construction of organic explanations of the greenhouse
effect and for a thorough understanding of the energy concept and the conservation
principle. These cognitive steps structure the teaching learning sequence in six phases with
the following objectives:
(a) to distinguish the quantities temperature, energy, heat and work, and realize that it is
possible to heat a body without giving heat and to give heat without heating the body;
(b) to recognize and explain a stationary condition of temperature for objects exposed to
sun or lamp radiation;
(c) to differentiate heat and radiation and recognize that objects emit thermal radiation;
(d) to understand that the behaviour of a material in its interaction with radiation depends
on the considered region of spectrum (glass is transparent to visible but absorbs
infrared thermal radiation);
(e) to put together and coordinate the new knowledge acquired in order to understand the
radiative greenhouse effect in a box-model;
(f) to extend the model for understanding the greenhouse effect on Earth and the problem
of climate change and global warming.
The complexity of the parameters involved in the explanation of the greenhouse effect
requires the construction of simplified didactic models of the phenomena according to the
particular student level. We have constructed a simplified model in analogy with solar
greenhouses, which allows explain the whole phenomenon in a conceptually correct way
even if it is approximated and incomplete (see below phase f).
Our teaching path includes lectures, laboratory and outdoor activities. Multiple expe-
riences are suggested to favour a progressive construction of knowledge, from the simplest
cases to the more complex ones, in a dialectic relation between hypothesis, experiments,
refined observations and more complex explanatory frameworks. This approach can con-
vey an idea of science as a human activity in continuous evolution and a dynamical
conception of models and theories, thus helping pupils to distinguish between models and
reality. Students are expected to understand that models offer only partial and limited
explanations and that further and more precise analysis is required. This means that the
construction of such a model fosters questions and allows new perspectives of inquiry.
We will describe in the following the main aspects and activities of each of the six
phases.

123
Teaching Energy Concepts

(a) To distinguish the quantities temperature, energy, heat and work, and realize that it is
possible to heat a body without giving heat and to give heat without heating the body
This implies an innovative treatment of calorimetry that includes also experiments in
which a body is heated by doing work on it (for example, by means of friction forces, by
compressing a gas, by means of an electric current in a lamp or in an electric wire), and the
analysis of situations involving the attainment of stationary temperatures in absence of
thermal equilibrium, for example the case of a room where heating is on.
We stress that it is possible to heat bodies without giving heat and to give heat without
heating the body. The interpretation of experiments of this type, together with other more
usual, helps to differentiate the quantities heat, work and internal energy and leads to
introduce and discuss the following relationship as consequence of the energy conservation
principle:
Energy ðentering into the systemÞ ¼ mcDT þ energy given by the system to the environment:

It is stressed that the quantity mcDT is an expression of the variation of internal energy
U connected with a temperature change, in the cases in which no phase change or other
effects than temperature increase are considered; this energy can be supplied to the system
by means of work, heat or radiation, then interpretation of mcDT simply as a formula for
heat exchange is misleading.
(b) To recognize and explain a stationary condition of temperature for objects exposed to
sun or lamp radiation
A brainstorming session is organised about the effects of solar radiation on living and non-
living objects (chemical reactions on photographic films, skin tanning, chlorophyll
photosynthesis, increase of temperature, etc.). Thermal effects occurring to an object
exposed to solar radiation are studied by means of specific experiences. Small aluminium
cylinders of equal masses and dimensions having white, black, and polished surfaces and a
transparent cylinder are exposed to solar or lamp light. A temperature sensor is inserted in
a little hole of each cylinder, while another sensor measures the ambient temperature
(Fig. 2). Graphs of temperature versus time for each cylinder are obtained with hand held
data loggers9 showing the process toward the stationary condition. Data are then analyzed
by the students by uploading them on the lab computers (Fig. 3).
The graphs show that each sample has a different stationary temperature which is also
different from the ambient temperature, and the subsequent cooling process. Discussion of
the results focuses on the fact that radiation produces thermal effects when it is ‘‘captured’’
by the material. An object can reflect, absorb or transmit the incoming radiation. An object
is transparent if part of the incoming energy passes through it without producing effects.
Interpretation of the graphs is based on the idea that a stationary condition is attained if
there is equality between the energy absorbed by the objects and the energy given up. This
issue is also useful to modify some incorrect ideas that many students have shown to
possess, for example:
The cylinder will increase its temperature indefinitely
It will raise the maximum possible temperature according to the material
The object will increase its temperature until it will be full of energy, so that it cannot receive any
more

9
We used the GLX data loggers by PASCO. We also used the metal cylinders, the metal bottles and the
plastic box provided by PASCO for the experiments described in the following.

123
U. Besson, A. De Ambrosis

Fig. 2 Measure of temperature of cylinders exposed to solar or lamp light

Fig. 3 Graphs of temperature versus time obtained by the students for cylinders exposed to solar or lamp
radiation

Moreover the differences in the graphs for cylinders of different colour or transparency
highlight the essential differences between thermal effects due to electromagnetic radiation
and due to heat conduction. This leads to the problem of differentiating heat and radiation.

123
Teaching Energy Concepts

In this phase the discussion should focus on the fact that radiation produces thermal effects
when it is absorbed by the material, while it passes through without producing thermal
effects when the material is transparent. Some students notice the peculiarity of this
behaviour:
The bottle does not heat up because it is transparent to the light and it lets the light go through; the
glass cylinder allows the light to go through, therefore it reaches a lower temperature
and recognize the difference from what happens with heat: heat heats up the material when
it passes through it, on the contrary, when radiation passes through the material it does not
heat up.
However, it is not possible to explain the difference between the stationary temperatures
of the differently coloured cylinders without analysing the problem of the emission of
radiation by the objects and the differences between visible and infrared radiation.
(c) To differentiate heat and radiation and recognize that objects emit thermal radiation;
By means of an infrared (IR) radiometer students can verify that bodies, at all
temperatures, emit radiation (Fig. 4). The changes of radiation energy emitted by an object
at different temperatures can be measured (values of temperatures under the ambient
temperature are also considered). A rapid increase of the emitted radiation power with the
temperature is verified. By inserting a layer of ‘‘clear’’ glass or plastic between the objects
and the radiation sensor, one can observe that the measured intensity drastically decreases.
This shows that most of the infrared radiation emitted does not pass through the glass or
plastic. On the contrary, visible light passes through the layer, in fact we can see the objects
behind the layer. The word clear, used in everyday language for glass and plastic, refers
only to visible radiation.
Experiments with the radiometer, although very simple, are meaningful because they
help students recognize that the objects, at any temperature, emit radiation and allow
understanding that in the energy budgets it is necessary to consider also the energy
exchanged as radiation.
(d) To understand that the behaviour of a material in its interaction with radiation
depends on the considered region of spectrum;

Fig. 4 Measuring the radiation


energy emitted by different
bodies

123
U. Besson, A. De Ambrosis

Laboratory activities are performed and discussed with the students to point out that:
– the emission spectrum of an object contains also invisible radiation beyond the red;
– the emission spectrum of a source depends on the temperature (at temperature near to
room temperature the emitted radiation is in the far infrared region);
– optical properties of materials depend on the considered region of the spectrum (on the
frequency or wavelength of the radiation);
– some materials are transparent to visible light but absorb far infrared radiation.
In order to further develop these aspects, graphs like the one of Fig. 5, representing the
spectrum of the sun, are discussed.
By using a glass prism the emission spectrum of a lamp, for example of an overhead
projector (Fig. 6), is observed on a screen. By intercepting the visible part of the light
exiting from the prism, students can observe through a digital camera the presence of
radiation beyond the red colour that they cannot see at naked eyes, i.e., the infrared
radiation. Aiming at seeing the invisible, simple observations are performed using a digital
camera (or a cell phone) having near infrared sensibility. For example the IR radiation
emitted by a remote control can be photographed as shown in Fig. 7.
These photos show that the remote control emits radiation and that the black plastic that
covers the remote control is opaque to visible light but it is transparent to the near infrared
radiation emitted by the video-camera. These experiences have been effective and
involving for students as this excerpt of a student report shows:
We created a rainbow with the help of an overhead projector as a source of light and an optical prism.
Using an infrared camera, we saw that there is ‘‘light’’ beyond the red colour of the visible spectre.
We discussed about visible light, its wavelength range and about ultraviolet and infrared light. We
established that we naturally cannot see infrared light, but we can observe it, using a camera. We
‘‘saw’’ infrared radiations emitted by a remote control through the camera, and using the same
camera, we looked at a lamp with different current intensities and voltages.
It is now possible to resume the problem of the different maximum temperature reached by
the two objects of different colour when exposed to the lamp radiation (higher for the black
one than for the white, see phase b and Fig. 3).
Students can observe, from Fig. 3, that in the transitory phase the slope of the heating
curve is steeper for the black cylinder than for the white, whilst the cooling time (for an

Fig. 5 Spectrum of solar radiation. On the horizontal axis the emission range of a body at 300 K is
indicated. The emission spectrum of this body would be very small and entirely below the curve of the solar
spectrum. The range of transparency of glass is represented and shows that glass is not transparent to
radiation emitted by bodies at a temperature around 300 K, i.e., at room temperature

123
Teaching Energy Concepts

Fig. 6 How to obtain a little rainbow by means of a glass prism and an overhead projector

Fig. 7 Two images of the same remote control obtained with a digital video-camera. The left photo shows
the infrared radiation emitted by the remote control. In the right photo the remote control is lit up by a beam
of infrared radiation emitted by the video-camera. Students can observe that the cover of the remote control
is opaque to the visible light but transparent to the near infrared radiation

equal temperature interval) is almost the same. This shows well that ‘‘blackness’’ and
‘‘whiteness’’ are properties which depend on the considered region of the spectrum: the two
objects, which we see as black and white under the normal sun or lamp light, because they
have very different absorptivity for solar (and lamp) radiation, have almost the same
emissivity and absorptivity for the far infrared thermal radiation emitted at the considered
temperatures. We can say that they are almost ‘‘black’’ for this radiation spectrum. They
therefore absorb visible radiation at different rate and emit infrared radiation with about
equal rate (at the same temperature), and therefore they reach different stationary tem-
peratures. To further confirm this conclusion the emitted radiation can be directly

123
U. Besson, A. De Ambrosis

measured by means of the radiometer. For example, we used an aluminium disk divided in
black, white and polished sectors and warmed by electric heather.
(e) To put together and coordinate the new knowledge acquired in order to understand
the radiative greenhouse effect in a box-model;
Students are asked to organize and to use the acquired knowledge to study and interpret the
behaviour of a small greenhouse box under solar or lamp radiation.
A black aluminium plate is placed on the bottom of a plastic box and it is exposed to
solar or lamp radiation (Fig. 8a). The temperature of the plate is measured by means of a
temperature sensor until the temperature becomes stationary. The measurements continue
with the clear glass (or plastic) lid on the top of the box. In this way a small greenhouse is
created (see Fig. 8b). Before carrying out the experiments, the students are asked to predict
and to draw graphs of the temperature of the plate versus time in the two cases. Two
examples of experimental graphs are shown in Fig. 9a, b.
To interpret the experimental results students need to use what they have just learned on
the thermal effects of radiation and to reorganize their knowledge in order to construct a
coherent representation of the energy fluxes in the process of attainment of the stationary
condition. The students are asked to represent the energy fluxes (in and out) for the plate
and the cover in the transitory and stationary conditions. A possible scheme is presented to
students and is shown in Fig. 10.
The figure refers to an ‘‘ideal’’ lid, which is considered transparent to all solar radiation
and absorbs all thermal far infrared radiation (of course no actual lid satisfies exactly these
conditions). The size of the arrows is qualitatively related to the amount of the energy
fluxes. The different fluxes are:
RS solar radiation entering into the box;
RO radiation from the environment;
RLu % RLd radiation emitted by the lid up and down;
RP radiation emitted by the plate;
QP and Q0 P heat flux from the plate to the lid through the air inside the box and to the air
outside through the walls;
QL heat flux from the lid to the air outside.

Fig. 8 Measuring the temperature of the black plate exposed to solar radiation without and with the plastic
lid on

123
Teaching Energy Concepts

Fig. 9 a Example of temperature versus time graph for the metal plate exposed to lamp radiation first
without then with the lid on. b Example of temperature versus time graph for the metal plate exposed to
solar radiation first without then with the lid on

In order to avoid the common idea of ‘trapping’ of sun rays, used as explanation of the
greenhouse effect, we stress that the emitted radiations RP and RL are not solar radiation
reflected by the plate and the lid but IR radiation emitted by the plate and the lid according
to their temperature.
(f) To extend the model for understanding the greenhouse effect on Earth and the
problem of climate change and global warming.

123
U. Besson, A. De Ambrosis

Fig. 10 Radiation and heat fluxes in the small greenhouse box in stationary condition

Fig. 11 The global annual mean energy budget of the Earth–Atmosphere system (for the March 2000 to
May 2004 period). Numerical values are in W/m2. (from Trenberth et al. 2009, p. 314)

The description of the energy fluxes involved in the energy budget of Earth–Atmosphere
system is quite complex, also in the schematic representations that are provided in
literature, as Fig. 11 shows. A more simplified version has been offered to students, based
on the model previously used (Fig. 10) to describe the stationary situation in the small
greenhouse box.
In the simplified model an analogy is established between elements of the Earth–
Atmosphere system and those of the small greenhouse:
– The Earth surface is the analogue of the black plate: it absorbs the incident radiation
and emits infrared radiation depending on its temperature.

123
Teaching Energy Concepts

– The atmosphere plays a role similar to the lid: it is transparent to most solar radiation,
but it absorbs most far infrared radiation emitted by the Earth, and emits infrared
radiation depending on its temperature.
Absorption of the infrared radiation emitted by the Earth is due primarily to water
vapour, clouds and CO2, with a smaller contribution from O3, N2O, CH4 and a little
contribution of other anthropogenic gases, such as the chlorofluorocarbons like CFCl3.
These gases, called greenhouse gases, emit infrared radiation toward the Earth and
toward the outer space. The energy fluxes in this simplified model are qualitatively rep-
resented in Fig. 12.
In comparing the atmosphere to the lid of a greenhouse, it is necessary to remark that no
heat flux can exit from the upper surface towards the space and that the thermal radiation
coming from the outside space RO is absolutely negligible, so that QL, Q0 P and RO must be
disregarded. Moreover, the atmosphere temperature is not uniform along the vertical
direction, so that RAd [ RAu and this fact is represented qualitatively by a different width
of the arrows. By writing the equations of the energy budgets of the Earth surface and of
the atmosphere, students are led to the surprising result that radiation energy emitted by the
atmosphere toward the Earth surface is a bit larger than the energy coming to the Earth
surface from the Sun.
Result of the in and out energy fluxes is a mean stationary temperature on the Earth
surface compatible with life. It is worth pointing out that the greenhouse effect has a
fundamental role in keeping the temperature on the Earth’s surface warm enough for life.
What it is potentially dangerous is an increase in this effect, the anomalous greenhouse
effect, not the effect in itself. An increase of CO2 and of the other greenhouse gases in the
atmosphere produces a variation of the energy budget of the Earth–atmosphere system with
a consequent change of the mean global temperature so as to re-establish the equilibrium
between the entering and exiting energy: to a decrease of R0 E, corresponds an increase of
R00 E, of RAd, and RAu. As a consequence, an increase of the temperature at the Earth
surface takes place.

Fig. 12 Simplified model of radiation and heat fluxes in the Earth–atmosphere system. RS indicates the
solar radiation; R0 S the solar radiation reflected by the atmosphere and by the Earth surface; R00 S the solar
radiation absorbed by atmosphere, in particular by clouds; RE is the radiation emitted by the Earth, most of
which (R00 E) is absorbed by greenhouse gases, water vapour and clouds and the remaining (R0 E) passes
through the atmosphere and goes into the outer space; RAd represents, the radiation emitted by the
atmosphere towards the Earth; RAu the radiation emitted towards the space. QE represents the heat passing
from the Earth’s surface to the atmosphere

123
U. Besson, A. De Ambrosis

Based on the study of the simplified model of Fig. 12, the students can then analyze a
representation like the one in Fig. 11, and verify the energy balances by using the
numerical values given in this figure.

8 Using the Teaching Learning Sequence in High School Classes

8.1 Organization of the Work and Implementation by the Teachers

The sequence was implemented in 2010 in six high school classes, four with 17–18 years
old students and two with 15–16 years old students, for a total of 121 students. We gave
the teachers a document describing the teaching learning sequence and we discussed it in
periodical meetings where the teaching plan of each teacher was progressively designed
with the necessary variations due to the different class situations.
Data collection was carried out by means of different tools:
– An initial and a final questionnaire. The questionnaires were designed by the
researchers and the teachers in a collective work. Aim of the initial questionnaire was
to obtain information on student’s ideas on thermal phenomena, thermal effects of
radiation and greenhouse effect as result of previous instruction and of information
from media and day life experience. The final questionnaire was on the same topics and
contained some questions identical as the initial one, with others more related to the
activity carried out by the students. They consisted of 6 and 7 open ended questions
respectively.
– Work sheets filled in by the students during the experimental activity. The worksheets
included students’ predictions, experimental results, comparison between predictions
and results, interpretations and conclusive remarks, relevant aspects of the experiments.
– Video and audio taping of some class activities and videos produced by the students to
document some of the experiments carried out.
– Teachers’ reports on the class work collected in log books.
All documents have been discussed in periodical meetings with the teachers and
uploaded in a web site devoted to the project.
Teachers prepared a final report where they described which elements of the proposed
sequence they considered essential, how they passed from the research plan to their
individual teaching plan and to the actual activity in classroom, and the results they
observed at the end of the sequence. After having taught the sequence, the teachers con-
tinued to work with us in 2011 to restructure the teaching of thermal phenomena in their
classes, to elaborate new materials for students and a teaching guide for the modified
sequence (materials are available at http://fisica.unipv.it/pls/Materiali.htm - L’Energia e la
sua conservazione).
Since the school situations were very different, teaching plans and methods presented
relevant differences but were in line with the core of the sequence project, the cognitive
progression, the aims, and the conceptual and methodological design.
We found it important to discuss with the teachers clearly and explicitly about the role
of different elements of the sequence, by distinguishing a core of contents, conceptual
correlations and methodological choices, which are essential, and a cloud of elements that
can be re-designed or skipped by teachers. This core-clouds structure (Besson et al. 2010b)
is useful both to permit teachers’ changes and to control them. In fact, physics education

123
Teaching Energy Concepts

research has shown that teachers tend inevitably to modify research based proposals to fit
them with the practical constraints of the class situation and with their personal prefer-
ences, attitudes and knowledge (Hirn and Viennot 2000; Pinto 2005; De Ambrosis and
Levrini 2010). Our experience confirms how it is important to analyse and discuss these
changes with the teachers and to compare them with the rationale and the general aims of
the original proposal. This work usually produces an improvement and an enhancement of
the original proposal thanks to the teachers’ contributions.

8.2 Results

Students’ answers to the initial and final questionnaires and their comments in the work-
sheets were analyzed to gain information on the improvement of their understanding of
energy concepts and of the greenhouse effect through the activity sequence. Answers were
at first analyzed by each teacher and grouped in categories. Then the categories were
discussed and revised by the whole group to obtain a shared categorization.
The answers show a progressively more correct and appropriate use of the concepts of
heat, thermal conduction, radiation, temperature, internal energy. In particular, the idea of
internal energy, almost non-existent at the beginning, becomes more and more a central
element in students’ descriptions. Most of the students realized both the possibility of
heating a body by using work or radiation and the difference between ‘‘heating’’ and
‘‘giving heat’’. In the following we summarize the results obtained.

8.2.1 Realizing that Bodies at Any Temperature Emit Radiation

Analysis of the pre- and post-tests shows a significant increase of the number of students
who consider thermal radiation emitted by bodies at ordinary temperature (77 vs. 14 %,
data reported in this section refer to 18–19 year old students):
A hot iron emits a lot of radiation that vanishes when the iron is cooling down, and in any case it
cannot be seen.
Ice emits very little, not visible radiation.
The black bottle heats up and cools down more than the others because black is a perfect absorber
and also a perfect emitter.
Glass is transparent to visible radiation, but it is opaque to infrared rays.
However, many students partly forget to consider thermal radiation in more complex
systems such as the greenhouse box, where thermal radiation emitted by the plastic lid is
often disregarded, even when the emission of the plate is correctly taken into account. This
reveals students’ difficulty in controlling the coherence of their explanations when many
aspects and parameters are involved in the physical situation. It appears the spontaneous
tendency to simplify the descriptions by considering only one (or few) factor and forgetting
the others, although already known and utilized in other simpler cases.

8.2.2 Explaining a Stationary Condition of Temperature as a Consequence of Energy


Balance

At the beginning, the idea that a stationary temperature is a consequence of a balance of


energy entering and exiting from the system was lacking in students’ reasoning. Com-
parison of pre-test and post-test results clearly shows an increase both of the awareness that
a stationary temperature is reached by a body exposed to sun radiation (100 vs. 80 %) and

123
U. Besson, A. De Ambrosis

of the explanations of the stationary temperature as due to a balance of the energy entering
and exiting from the system (59 vs. 13 %).
After a while, the temperature of the cylinders becomes constant because there is equilibrium
between absorbed and given up radiation.
The black one reaches a higher temperature and in about 20 min the temperature is constant because
there is equilibrium between absorbed and emitted radiation.
The temperature will be stable when the power emitted will equal the power absorbed.
Nevertheless, we found a non negligible persistence of the idea of ‘‘saturation’’ (32 vs.
67 %): according to this idea materials have their own maximum possible temperature that
once it is reached necessarily cannot increase yet. This fact can be interpreted as due to the
slowness of the process needed to pass from explanations in terms of object properties to
reasoning in terms of interactions and balances.
The cylinder absorbs heat from the sun, its temperature increases, then it reaches saturation and the
temperature keeps constant.
We observed how the reasoning used by students was strongly influenced by the first
presentation of the topic (examples, descriptions, approximations, explanations…);
actually they tended to reuse, for the new situation, the same reasoning pattern used
before, even if it was not appropriate. We call this tendency, observed also in other topics,
the ‘‘imprinting’’ phenomenon in science learning (borrowing the term introduced by K.
Lorenz in different contexts).
In particular, we noticed that in the classes where the teacher, following the usual
curriculum, had already dealt with thermal phenomena, with experiments and exercises
stressing thermal equilibrium situations and processes leading to equal temperature of the
interacting bodies and of the bodies with the environment, many students tried to apply the
same reasoning pattern to the case of bodies exposed to sun (or lamp) radiation, thus
interpreting the constant temperature as due to thermal equilibrium with the environment:
The bottles reach a thermal equilibrium between environment and system.
The condition of constant temperature is the condition of thermal equilibrium with the environment,
heated by the lamp.
Such explanations were not found in students of the classes where the analysis of thermal
phenomena started just with the experiments dealing with the heating of bodies exposed to
lamp or sun radiation. In this case, reasoning based on the idea of energy balance was more
frequent to explain the difference between the constant temperature of the object and of the
environment.
To avoid the ‘‘imprinting’’ effect it is useful to present since the onset a wide panorama
of different situations and factors, in a qualitative and simplified way, in order to convey
interpretations not limited to a particular aspect. For example, in introducing thermal
phenomena it is useful to consider both situations of thermal equilibrium and stationary
non equilibrium situations, in familiar simple cases. The idea that a stationary temperature
condition does not entail thermal equilibrium is developed in the materials developed in
2011 where experiments, examples, and theoretical arguments are proposed to stress this
aspect.

8.2.3 Understanding the Greenhouse Effect

Concerning the understanding of the greenhouse effect, the results confirm the importance
of passing through all the cognitive steps of the teaching sequence giving enough time to
each of them and spiraling back to the previous ones after a time, in a different context.

123
Teaching Energy Concepts

Presenting the entire explanation of the greenhouse effect in a unique step is not effective;
the phenomenon is complex and needs a progressive rapprochement.
The experiments with the greenhouse box and the representation of energy fluxes to
interpret the experimental results were useful and effective to model the greenhouse effect
on the Earth. Students easily transferred to the greenhouse effect on the Earth the expla-
nations (sometimes even incorrect) acquired by analysing the box model, having well
understood the similarities of the two situations. Results obtained in the items related to the
greenhouse effect show a strong decrease of explanations based only on thermal isolation
(4 vs. 37 %) and an increase of correct or almost correct reasoning, i.e., at least differ-
entiating phenomena of absorption and emission of visible and infrared radiation (57 vs.
17 %).
Because radiation emitted by the black plate (in the green house box) is blocked by the lid. This fact
produces an increase of the temperature inside the green house. Moreover, the lid prevents con-
vection from inside to outside and favours the increase of temperature inside.
Solar radiation passes through, but radiation emitted by the Earth does not pass.
However, many explanations still reveal some imprecision and confusion:
Solar radiations pass through the transparent roof of a green house, but after they have hit the ground
they change wavelength and cannot go out. It is the same mechanism of the Earth to keep a stable
temperature; instead of the glass roof we have the atmosphere.
The rays, hitting the greenhouse, heat the material of the plate that gives up heat inside, but this heat
cannot go out and then the temperature inside is higher than outside.
Moreover, the idea of ‘‘trapping’’ of sun rays is still present (30 %).
Inside a greenhouse it is much warmer than outside because the transparent material lets part of the
sun rays pass through, but they are trapped inside the greenhouse and warm it up.
Because it (the plastic cover) keeps inside a fraction of the radiation that produces an increase of the
internal energy.
The increase of CO2 and greenhouse gases due to human activity in the atmosphere produces the
greenhouse effect, which means trapping of part of sun radiation in the atmosphere which contributes
to global warming. It is the same phenomenon that, in small-scale, happens in a greenhouse.
It is worthwhile to notice that some students recognised and explicitly remarked a change
in their idea of ‘‘trapping’’, but this change did not mean that they have completely
abandoned such idea, thus confirming the persistency of this conception and the difficulty
of a radical reasoning change:
I was thinking that the radiation trapped in the greenhouse was the one emitted by the sun, on the
contrary it is the one emitted by the plate.
It is interesting to compare these data with the results we obtained with a group of 51
graduated student teachers: only 6 % gave a correct explanation, 35 % expressed the idea
of trapping of sun rays and 31 % sustained that sun radiation enters and heats, but heat
cannot get out or can get out partially by conduction and convection.

8.3 Design of New Materials

Results confirm the basic choices inspiring the design of the cognitive path and its
effectiveness for learning energy concepts. However, at the same time some critical ele-
ments in the teaching paths were pinpointed. The teachers recognized that the idea of
energy budget and its ‘‘structuring’’ role to interpret the considered phenomena were not
sufficiently stressed during the implementation of the sequence. Moreover, they
acknowledged that the students’ conceptual difficulties in passing from the understanding

123
U. Besson, A. De Ambrosis

of single parts of the sequence to a global understanding of the greenhouse effect were
underestimated.
The students’ responses and teachers’ final reports showed the complexity of the topic
and its manifold conceptual implications, and gave hints to respond to the research
question: what type of materials, experiments, models and schematic representations can
improve students’ understanding of this topic. Focusing on this question and on the
obtained results, we revised the teaching sequence and prepared new materials for the
students and a teacher guide. A source of difficulty underlined by the teachers was the lack
of booklets for students designed to help them in deepening and systematizing the different
topics also with a personal study, since textbooks appeared inadequate and sometimes
misleading and the students’ notes were insufficient for the task.
We prepared three types of material: sheets for experiments; booklets for students’
personal study and deepening; suggestions for teachers.
Worksheets for experiments describe the objectives of the experiment, its role in the
sequence, not the expected result, but the issues addressed. They include different phases:
prediction, measurement, data collection and processing, interpretation of the results and
final questions. Experiments are considered part of the scientific discourse on the topic, in a
dialogue between theory, hypothesis and experiments. The booklets are not designed as
handbook chapters or scientific popularizations, but to be adherent as much as possible to
the experimental activities and to resume their particular aspects, by taking into account
also what students have observed. Aims of these materials are to propose a discussion of
the experiments, to clarify the results obtained, the problems stemmed, the conclusions that
can be drawn by referring to the whole teaching path. Moreover, the material systematizes
the results obtained in a theoretical framework coherent as much as possible though
simplified and not definitive, also referring to other experimental results. Typical students’
difficulties and conceptions are taken into account and addressed with theoretical and
experimental arguments. We tried to clearly differentiate between conclusions which
derive directly from experiments and explanations involving microscopic models, based on
entities such as atoms, molecules, electrons, quanta. The school level suggests teachers be
cautious and to limit explanations to a descriptive phenomenological model. When
structural models are introduced, we try to define the appropriate level of elaboration and
detail and the theoretical reference frame, with drastic simplifications but without con-
ceptual mistakes or general inconsistencies.
The revised path is divided in three phases in order to facilitate its insertion in the usual
curriculum: T (thermal phenomena), R (radiation), and S (greenhouse effect). Moreover,
we developed an introductory phase O to supply the knowledge on optics and waves
necessary for the other phases and to introduce the wave model for light and electro-
magnetic radiation. This part can be used with students who did not study the topic before
or proposed to recall topics already studied.
Phase T is a crucial one, different from the usual approach to heating and cooling
phenomena proposed in textbooks. We suggest to consider with students, since the
beginning, different way to increase the body temperature (mechanical work, friction, heat,
chemical reactions, lamp and microwaves radiation), and to show that it is possible to heat
bodies without giving heat and to give heat without heating. Experiments are proposed and
discussed where thermal equilibrium is reached and others where stationary temperature
conditions are attained without thermal equilibrium, due to the energy fluxes balance, also
referring to everyday life examples. A precise theoretical differentiation between thermal
equilibrium conditions and stationary non equilibrium conditions is given. Descriptions
based on the energy concept are used and energy is introduced as a quantity useful to

123
Teaching Energy Concepts

quantitatively connect different phenomena and to describe processes and transformations


involving different empirical areas. The link between internal energy and temperature is
introduced in a qualitative or semi-quantitative way.
Phase R deals with the dependence of the properties of radiation-matter interaction
(absorption, reflection, transparency, and emissivity) on the radiation frequency. The aim is
to pass from the idea of radiation as a whole to the idea of spectrum with a distinction in
different types, and in particular between visible and infrared radiation. The experiments
presented in Sect. 7 are described and discussed and the temperature versus time graphs of
bodies exposed to sun or lamp radiation are accurately analysed by distinguishing the
almost linear initial part, its bending, and the stationary final part, which is a key point for
understanding the following phase S.
Phase S aims at integrating the knowledge acquired in the previous work in order to
give a coherent and correct interpretation of the greenhouse effect in a box-model and of
the greenhouse effect on the Earth so that the general problem of global warming can be
discussed.

9 Concluding Remarks

Energy is a central concept in physics and is a key concept for understanding physical,
biological and technological world. Many researches in science education have studied
how to present energy concepts and phenomena, different approaches have been proposed
and tested, while students’ difficulties and conceptions have been investigated. Energy is a
complex topic with multiple connections with different science areas and with social,
environmental and philosophical issues.
The choice of a particular teaching approach is not a direct consequence of technical
content knowledge, but involves implicit or explicit epistemological assumptions on
learning processes, on the role of science teaching in cultural development, on the nature of
science and on its connections with social and environmental issues.
However, in order to produce effective teaching/learning sequences, it is necessary to
design teaching paths based on research work and empirical observations. As our expe-
rience also showed, even activity sequences and teaching methods that appear properly
organized and structured to the designer may reveal ineffective and problematic for stu-
dents and unable to address their difficulties. For that reasons well structured experi-
mentations with students are necessary.
In this paper we have discussed some aspects of the teaching and learning of energy
concepts, and we have reported results of research in science education on this issue. Then
we have described a teaching learning path, devoted to high school students, that has been
developed around a specific driving issue, greenhouse effect and global warming, to
promote the progressive construction of physics concepts and models.
Using the sequence with high school students has shown that the approach based on
driving issues promotes students’ engagement toward a deeper understanding of the topic
and favours further insight.
Data collected allowed us: (1) to focus on the limitations in the traditional way of
dealing with thermal phenomena and optics as separate topics, not connected with the
general energy problem, (2) to identify with more detail particular students’ difficulties on
thermal effects of radiation-matter interaction and on the interpretation of stationary
conditions in terms of energy balances.

123
U. Besson, A. De Ambrosis

Analysis and discussion with the teachers of the class work results was followed by a
revision of the teaching sequence and by the production of new materials for students and
for teachers. These materials are the result of a didactical reconstruction of the content
matter realized by the group of teachers in collaboration with our research group starting
from the results obtained with the students. Thus, a new cycle of implementation, evalu-
ation and refinement has been activated this year with a larger group of teachers and
students in order to take into account effects and elements of the materials that only a
contextualised practice can reveal. We think that this type of systematic and long term
collaboration with teachers can help to fill the gap, widely recognized, between the results
of science education research and the actual school practice.

Acknowledgments We would like to thank all the students and teachers who participated in this work.
Research was partly supported by the National Project Piano Lauree Scientifiche (Degrees in Science
Project) funded by the Italian Ministry of Education.

References

Andersson, B., & Kärrqvist, C. (1983). How Swedish pupils, aged 12–15 years, understand light and its
properties. European Journal of Science Education, 5(4), 387–402.
Andersson, B., & Wallin, A. (2000). Students’ understanding of the greenhouse effect, the societal con-
sequences of reducing CO2 emissions and problem of ozone layer depletion. Journal of Research in
Science Teaching, 37(10), 1096–1111.
Arnold, M., & Millar, R. (1996). Learning the scientific story: A case study in the teaching and learning of
elementary thermodynamics. Science Education, 80(3), 249–281.
Baierlein, R. (1999). Thermal physics. Cambridge: Cambridge University press.
Baracca, A., & Besson, U. (1990). Introduzione storica al concetto di energia. Firenze: Le Monnier.
Besson, U. (2003). The distinction between heat and work: An approach based on a classical mechanical
model. European Journal of Physics, 24, 245–252.
Besson, U. (2009). Paradoxes of thermal radiation. European Journal of Physics, 30, 995–1007.
Besson, U. (2012). The history of cooling law: When the search for simplicity can be an obstacle. Science &
Education, 21(8), 1085–1110.
Besson, U., De Ambrosis, A., & Mascheretti, P. (2010a). Studying the physical basis of global warming:
Thermal effects of the interaction between radiation and matter and greenhouse effect. European
Journal of Physics, 31, 375–388.
Besson, U., Borghi, L., De Ambrosis, A., & Mascheretti, P. (2010b). A three-dimensional approach and
open source structure for the design and experimentation of teaching learning sequences: The case of
friction. International Journal of Science Education, 32, 1289–1313.
Bevilacqua, F., & Falomo, L. (2010). Energia questa trasformista—Laboratorio storico interattivo. Pavia:
La Goliardica Pavese.
Blumenfeld, P. C., Soloway, E., Marx, R. W., Krajcik, J. S., Guzdial, M., & Palincsar, A. (1991). Motivating
project-based learning: Sustaining the doing, supporting the learning. Educational Psychologist, 26(3),
369–398.
Boyes, E., & Stanisstreet, M. (1993). The greenhouse effect: Children’s perception of causes, consequences
and cures. International Journal of Science Education, 15, 531–552.
Chauvet, F. (1996). Teaching colour: Designing and evaluation of a sequence. European Journal of Teacher
Education, 19(2), 119–134.
Coelho, L. R. (2009). On the concept of energy: How understanding its history can improve physics
teaching. Science & Education, 18, 961–983.
Cordero, E., Todd, A. M., & Abellera, D. (2008). Climate change education and the ecological footprint.
Bulletin of the American Meteorological Society, 89(6), 865–872.
Cotignola, M. I., Bordogna, C., Punte, G., & Cappannini, O. M. (2002). Difficulties in learning thermo-
dynamic concepts: Are they linked to the historical development of this field? Science & Education,
11, 279–291.
De Ambrosis, A., & Levrini, O. (2010). An empirical study for reconstructing the appropriation path in the
case of special relativity. Physical Review Special Topics-Physics Education Research, 6(020107),
1–11.

123
Teaching Energy Concepts

de Berg, K. C. (2008). The concepts of heat and temperature: The problem of determining the content for the
construction of an historical case study which is sensitive to nature of science issues and teaching-
learning issues. Science & Education, 17, 75–114.
De Felice, F. (2000). Gli incerti confini del cosmo. Milano: Bruno Mondadori Editore.
Domenech, J. L., Gil-Perez, D., Gras-Martı̀, A. G., Guisasola, J., Martinez-Torregrosa, J., Salinas, J., et al.
(2007). Teaching of energy issues: A debate proposal for a global reorientation. Science & Education,
16, 43–64.
Duit, R. (1986). In search of an energy concept. In R. Driver & R. Millar (Eds.), Energy matters (pp.
67–101). Leeds: University of Leeds.
Edelson, D., Gordin, D., & Pea, R. (1999). Addressing the challenges of inquiry-based learning through
technology and curriculum design. Journal of the Learning Sciences, 8(3), 391–450.
Erickson, G., & Tiberghien, A. (1985). Heat and temperature. In R. Driver, E. Guesne, & A. Tiberghien
(Eds.), Children’s ideas in science (pp. 52–84). Philadelphia: Open University Press.
Goldring, H., & Osborne, J. (1994). Students’ difficulties with energy and related concepts. Physics Edu-
cation, 29, 26–31.
Groves, F., & Pugh, A. (1999). Elementary pre-service teacher perception of the greenhouse effect. Journal
of Science Education and Technology, 8, 75–81.
Guesne, E. (1985). Light. In R. Driver, E. Guesne, & A. Tiberghien (Eds.), Children’s ideas in science (pp.
10–32). Philadelphia: Open University Press.
Hecht, E. (2007). Energy and change. Physics Teacher, 45, 88–92.
Helmholtz, H. (1847). Über die Erhaltung der Kraft (On the Conservation of Force). Berlin: G. Reimer.
Hicks, N. (1983). Energy is the capacity to do work—or is it? Physics Teacher, 21, 529–530.
Hirn, C., & Viennot, L. (2000). Transformation of didactic intention by teachers: The case of geometrical
optics in grade 8 in France. International Journal of Science Education, 22(4), 357–384.
IPCC Intergovernmental Panel on Climate Change. (2007). Fourth Assessment Report (AR4). http://www.
ipcc.ch/index.htm.
Kaminski, W. (1989). Conceptions des enfants et des autres sur la lumière. Le BUP, 716, 973–996.
Kilinc, A., Stanisstreet, M., & Boyes, E. (2008). Turkish students’ ideas about global warming. International
Journal of Environmental & Science Education, 3(2), 89–98.
Koulaidis, V., & Christidou, V. (1999). Models of students’ thinking concerning the greenhouse effect and
teaching implications. Science Education, 83, 559–576.
Krajcik, J., Blumenfeld, P., Marx, R., Bass, K., Fredricks, J., & Soloway, E. (1998). Inquiry in project-based
science classrooms: Initial attempts by middle school students. Journal of the Learning Sciences, 7(3),
313–350.
Lee, H. S., & Liu, O. L. (2010). Assessing learning progression of energy concepts across middle school
grades: The knowledge integration perspective. Science Education, 94, 665–688.
Lehrman, R. L. (1973). Energy is not the ability to do work. Physics Teacher, 11, 15–18.
Lester, B. T., Li, Ma., Lee, O., & Lambert, J. (2006). Social activism in elementary science education: A
science, technology and society approach to teach global warming. International Journal of Science
Education, 28, 315–339.
Lijnse, P. L., Eijkelhof, H., Klaassen, C., & Scholte, R. (1990). Pupils’ and mass-media ideas about
radioactivity. International Journal of Science Education, 12, 67–78.
Loverude, M. E., Kautz, C. H., & Heron, P. R. L. (2002). Student understanding of the first law of
thermodynamics: Relating work to the adiabatic compression of an ideal gas. American Journal of
Physics, 70, 137–148.
Maxwell J. C. (1871, 10th ed. 1891). Theory of heat. London: Longmans, Green & Co, reprint 1902.
Meadows, G., & Wiesenmayer, R. (1999). Identifying and addressing students’ alternative conceptions of
the global warming: the need for cognitive conflict. Journal of Science Education and Technology, 8,
235–239.
Meltzer, E. D. (2004). Investigation of students’ reasoning regarding heat, work, and the first law of
thermodynamics in an introductory calculus-based general physics course. American Journal of
Physics, 72, 1432–1446.
Millar, R. (1994). School students’ understanding of key ideas about radioactivity and ionizing radiation.
Public Understanding of Science, 3(1), 53–70.
Ogborn, J. (1986). Energy and fuel: The meaning of the ‘go of things’. School Science Review, 242, 30–35.
Onorato, P., Mascheretti, P., & De Ambrosis, A. (2011). ‘Home made’ model to study the greenhouse effect
and global warming. European Journal of Physics, 32, 363–376.
Osterlind, K. (2005). Concept formation in environmental education: 14-year olds’ work on the intensified
greenhouse effect and the depletion of the ozone layer. International Journal of Science Education, 27,
891–908.

123
U. Besson, A. De Ambrosis

Papadimitriou, V. (2004). Prospective primary teachers’ understanding of climate change, greenhouse


effect, and ozone layer depletion. Journal of Science Education and Technology, 13, 299–307.
Papadouris, N., & Constantinou, C. P. (2011). A philosophically informed teaching proposal on the topic of
energy for students aged 11–14. Science & Education, 20(10), 961–979.
Pinto, R. (2005). Introducing curriculum innovations in science: Identifying teachers’ transformations and
the design of related teacher education. Science Education, 89, 1–12.
Rant, Z. (1956). Exergie, ein neues Wort für ‘‘Technische Arbeitsfähigkeit’’ (Exergy, a new word for
‘‘technical available work’’). Forschung auf dem Gebiete des Ingenieurwesens, 22, 36–37.
Redfors, A. (2001). University physics students’ use of models in explanations of phenomena involving
interaction between metals and electromagnetic radiation. International Journal of Science Education,
23, 1283–1301.
Rego, F., & Peralta, L. (2006). Portuguese students’ knowledge of radiation physics. Physics Education, 41,
259–262.
Reiser, B. J. (2004). Scaffolding complex learning: The mechanisms of structuring and problematizing
student work. Journal of the Learning Sciences, 13(3), 273–304.
Reiser, B. J., Tabak, I., Sandoval, W. A., Smith, B. K., Steinmuller, F., & Leone, A. J. (2001). BGuILE:
Strategic and conceptual scaffolds for scientific inquiry in biology classrooms. In S. M. Carver & D.
Klahr (Eds.), Cognition and instruction: Twenty-five years of progress (pp. 263–305). Mahwah, NJ:
Lawrence Erlbaum Associates.
Romer, R. H. (2001). Heat is not a noun. American Journal of Physics, 69, 107–109.
Rozier, S., & Viennot, L. (1991). Students’ reasoning in thermodynamics. International Journal of Science
Education, 13, 159–170.
Rye, J. A., Rubba, P. A., & Wiesenmayer, R. L. (1997). An investigation of middle school students’
alternative conception of global warming. International Journal of Science Education, 19, 527–551.
Sandoval, W., & Reiser, B. (2004). Explanation-driven inquiry: Integrating conceptual and epistemic
scaffolds for scientific inquiry. Science Education, 88(3), 345–372.
Shayer, M., & Wylam, H. (1981). The development of the concepts of heat and temperature in 10–13 years-
olds. Journal of Research in Science Teaching, 5, 419–434.
Solbes, J., Guisasola, J., & Tarin, F. (2009). Teaching energy conservation as a unifying principle in physics.
Journal of Science Education and Technology, 18(3), 265–274.
Solomon, J. (1983). Learning about energy: How pupils think in two domains. European Journal of Science
Education, 5, 49–59.
Stavy, R., & Berkovitz, B. (1980). Cognitive conflict as a basis for teaching quantitative aspects of the
concept of temperature. Science Education, 64(5), 679–692.
Svihla, V., & Linn, M. C. (2012). A design-based approach to fostering understanding of global climate
change. International Journal of Science Education, 34(5), 651–676.
Trenberth, K. E., Fasullo, J. T., & Kiehl, J. (2009). Earth’s global energy budget. Bulletin of the American
Meteorological Society, 90, 311–323.
Viennot, L. (2001). Reasoning in physics, the part of common sense. Dordrecht: Kluwer.
Viglietta, L. (1990). Efficiency in the teaching of energy. Physics Education, 25, 317–321.
Watts, D. M. (1983). Some alternative views of energy. Physics Education, 18, 213–217.
Zemanski, M. W. (1957). Heat and thermodynamics. New York: McGraw-Hill Book Company.

123

Potrebbero piacerti anche