78
1 The Handbook of Research on Science Education Chapter … “Learning Earth Sciences” Authors: Nir Orion and Charles R. Ault, Jr. 1. Introduction Great news! I’ve just been accepted into graduate school in geology with the opportunity to work on a terrific research project. Among other opportunities and challenges, the professor I’ll work with would like someone to do a photographic survey of the Lower Colorado River along the same route as traversed by an expedition of 150 years ago and documented in journals and watercolor paintings. The aim is to compare habitats and channels today with those from the past within the context of reconstructing climate trends in western North America. The work would be very similar to what I did in Argentina on my fellowship last year, where I visited Charles Darwin’s fossil collecting locales and compared his journal entries as well as sketches of landscapes made by the Beagle’s artist with present day photographs. I am very exited about getting started. I can’t believe that there is a project in geology so similar to what I have dreamed about doing. --Message from the second author’s son The indirect quotation above is from a real situation. It captures the challenges and opportunities for graduate study in earth science that echo the themes and claims developed in this chapter. The message is about a research opportunity and the nature of authentic inquiry. The proposed research crosses several disciplines, though is housed in geology and geomorphology. Extrapolations from the study have importance to understanding climate change on different scales in time and space. The data include works of art found in historical literature. The reconstruction of past habitats and the extrapolation of future ones serve the public interest in terms of guiding human actions in response to environmental change. The research has intrinsic appeal to some, social value to many. It is an example of what earth scientists at the dawn of the 21 st Century are doing. This example of a message from an excited new graduate student presents tangible imagery consistent with an idealized set of characteristics tentatively proposed in this chapter as representative of the earth sciences. Section 1, “Distinctive characteristics,”

Revision of Learning Earth Sciences 7-30-04-Nir

Embed Size (px)

Citation preview

Page 1: Revision of Learning Earth Sciences 7-30-04-Nir

1

The Handbook of Research on Science Education

Chapter …

“Learning Earth Sciences” Authors: Nir Orion and Charles R. Ault, Jr.

1. Introduction Great news! I’ve just been accepted into graduate school in geology with the opportunity to work on a terrific research project. Among other opportunities and challenges, the professor I’ll work with would like someone to do a photographic survey of the Lower Colorado River along the same route as traversed by an expedition of 150 years ago and documented in journals and watercolor paintings. The aim is to compare habitats and channels today with those from the past within the context of reconstructing climate trends in western North America. The work would be very similar to what I did in Argentina on my fellowship last year, where I visited Charles Darwin’s fossil collecting locales and compared his journal entries as well as sketches of landscapes made by the Beagle’s artist with present day photographs. I am very exited about getting started.I can’t believe that there is a project in geology so similar to what I have dreamed about doing.

--Message from the second author’s son

The indirect quotation above is from a real situation. It captures the challenges and

opportunities for graduate study in earth science that echo the themes and claims

developed in this chapter. The message is about a research opportunity and the nature of

authentic inquiry. The proposed research crosses several disciplines, though is housed in

geology and geomorphology. Extrapolations from the study have importance to

understanding climate change on different scales in time and space. The data include

works of art found in historical literature. The reconstruction of past habitats and the

extrapolation of future ones serve the public interest in terms of guiding human actions in

response to environmental change. The research has intrinsic appeal to some, social value

to many. It is an example of what earth scientists at the dawn of the 21st Century are

doing.

This example of a message from an excited new graduate student presents tangible

imagery consistent with an idealized set of characteristics tentatively proposed in this

chapter as representative of the earth sciences. Section 1, “Distinctive characteristics,”

Page 2: Revision of Learning Earth Sciences 7-30-04-Nir

2

introduces these features, suggesting that they are features of earth sciences with

particular importance for teaching and learning. There follows a profile of earth science

education worldwide, including trends evident over the past twenty-five years. This

profile focuses on significant reforms in geosciences education undertaken at the very end

of the 20th century: the trend away from disciplinary-based science education towards an

integrative, environmentally-based, earth systems approach, in part a consequence of

profound expectations for the science K-12 curriculum stemming from the “Science for

All” movement.

“Learning earth sciences,” Section 2 of this chapter, continues with careful

attention to the empirical record of learning earth sciences in schools. Section 2 identifies

the main characteristics of earth science education in the schools, such as the integration

of subjects within earth sciences and between earth sciences and environmental education.

Section 2 then proceeds to examine the cognitive aspects of learning earth sciences:

misconceptions, spatial visualization, temporal thinking, and systems thinking. This

section ends by reporting on the integration of learning environments within the earth

sciences and the prospects for cultivating environmental attitudes and insights from

learning earth sciences. The learning environments reviewed are: the outdoor and indoor

classrooms, the earth science laboratory, and the virtual world of computer environments.

Today’s ambitious reform agenda guided in general by the principle of “science

for all” and in particular by a theme of “citizen science” within earth and environmental

education scaffolds Section 3. Here the concern becomes how well, or how poorly,

teachers have adapted to calls for changing their philosophies of teaching: their

instructional goals, content priorities, value contexts, and teaching practices. Section 3

deals with the difficulties of reforming earth science education for science teachers who

have limited content knowledge and who may lack motivation to deal with new priorities

among subjects, unfamiliar learning environments, and changes in teaching strategies.

The chapter concludes with by challenging researchers to study teaching and

learning in the earth sciences not only as historically practiced in the traditional sense of

disciplinary based curriculum, but also as increasingly practiced, according to emerging

idealizations, as integrated study. The conclusion acknowledges that, from a research

Page 3: Revision of Learning Earth Sciences 7-30-04-Nir

3

perspective, we know very little about teaching and learning earth sciences when they

have been thoroughly contextualized: for example, in the context of inquiry about changes

in the climate of western North America. Such contexts value knowledge for the sake of

making public policy, not only theory-building and model-testing within the earth

sciences. Such contexts find promising data not only in records of sediments, but also in

historical photography, journals, and art. The chapter ends, in effect, with the challenge to

the next generation of researchers of learning earth sciences to embrace ambitious

integration and social contextualization as essential features of the subject. At the same

time, we make the call to preserve distinctive characteristics of the earth sciences when

setting objectives for student learning. 1.1 Distinctive characteristics

Evaluating the stature, role, and distinctiveness of learning earth sciences faces

difficulties largely avoided in the physical and life science fields that dominate science

education and research about teaching and learning sciences. There are any number of

historical reasons for its perceived low status and an equal number of reasons to call for

elevating its stature within the context of science education for all. More immediately,

there is a need to characterize the crucial features of the earth sciences adequately and

appropriately for the purpose of setting limits on the scope of research about learning

earth sciences for this chapter.

Every subject has something important to offer science for all. Much of the

challenge to curriculum designers intent upon reaching the goal of science for all is one

establishing priorities. There are limits on time, resources, and cognitive development that

must be respected. To begin this process, we suggest focusing on those features of the

field deemed important to organizing teaching and learning. These features ought to

encompass (1) an “intellectually honest” (Bruner, 19xxx) portrait of what scientists do

(e.g., date rocks radiometrically) and know as well as (2) ideas with high “conceptual

worth” (Toulmin, 19xxx) that have advanced thinking and solving problems through time

(e.g., the law of superposition). The host of individual fields that comprise the earth

sciences and the need to integrate these subjects within schools makes characterizing their

distinctive features imperative. Furthermore, characterizing crucial features of a subject

Page 4: Revision of Learning Earth Sciences 7-30-04-Nir

4

begins the process of determining its distinctive potential to make contributions to science

learning for all. The question is, “What’s so important to learn from earth science

Let us emphasize this notion of “distinctiveness” on four levels: disciplinary,

psychological, pedagogical, and socio-historical. Characterization of the crucial features

of a subject begins with attention to phenomena of interest (history of the earth, for

example) that are distinctive to the discipline, then turns to cognitively distinctive

challenges for learning these phenomena (psychological misconceptions about geologic

time, for example). Approaches to pedagogy must demonstrate their responsiveness to

such distinctive cognitive challenges (making use of outdoor learning or field study, for

example). The endpoint for characterization of a subject’s distinctive potential is

consideration of its social and historical context: how knowing about climate change and

its scale may matter in the personal and social lives of citizens, for example. Derived most

explicitly from the geosciences, the use of the label “earth sciences” encompasses a host

of fields and subfields in geology, hydrology, oceanography, meteorology, climatology

and even astronomy. Clearly, a definitive characterization of the crucial features of the

earth sciences remains well beyond the scope of this (and perhaps any other) chapter.

Nevertheless, there are heuristically useful questions to pose in the search for distinctive

features of the earth sciences. These features, to repeat, are ones useful to curriculum

design, framing the scope of research about teaching and learning earth sciences, and

promoting science for all. For this chapter our questions are:

1. What are the earth sciences about?

2. What distinctive features of earth science education merit the attention of

researchers and curriculum authors?

1.1.1 What are the earth sciences about?

Simply everything beneath our feet and above our heads, with concern as well for

how our collective actions fit within these realms. They are about all of the phenomena

addressed by an extensive array of disciplines as well as about how these different

disciplines understand the same phenomena from different perspectives. To learn about

the earth sciences is to learn about complex systems on many scales in time and space,

about the interactions of these systems with each other and us with them. Such a

Comment: I suggest to ommit this paragragh from two reasons: 1) We say it later in several places so if we have to cut this won't cause any harm. 2) I am not sure that I can collapse ES to a one ward and if I had to do I am not sure that I would chose scales.

Comment: I think that we can omit this section since by pointing on the distinctive features we also say in what we are differ. Since we have to cut this is a place that I feel that cutting causes no harms.

Page 5: Revision of Learning Earth Sciences 7-30-04-Nir

5

characterization unifies, to an important degree, the common concerns of the several earth

science disciplines.

Daunting as this opening assertion may seem, the task of characterizing learning

earth sciences is not intractable. The earth sciences are both similar to and distinct from

other fields of science.

1.1.2 What distinctive features of earth science education merit the attention of

researchers and curriculum authors?

In part, the history and philosophy of science, when turned toward the

examination of geological explanations in general and the concept of geologic time in

particular, reveal characteristic features of thought in the earth sciences (xxx Kitts, Gould

from American Scientist, Ault from JRST, refs. From Dodick, and Drifting Continents,

Shifting Theories—with its bibliography to check.) In part, the psychology of learning

earth science concepts unveils what is cognitively distinctive about this field (or set of

fields) as well.

This sense of duality—similar to, yet distinct from other disciplines; resembling

these fields in some important ways, but differing in other, perhaps even more important

respects—permeates not only research about learning the earth sciences but also the

explanatory approach to many earth science problems themselves. This strategy of

compare and contrast has proven essential to forming understandings of earth’s complex

features and systems that have resulted from long and complex histories. Indeed, Gould

has characterized approaches to problem solving in geology, paleontology, and evolution,

from Lyell and Darwin forward, as a distinctive historical style of argument and

explanation in science (Gould, American Sci. 19xxx cite here). The objects of

explanation—e.g., mountain building, ice age onset, seafloor topography, storm

generation, magma distillation, planetary coalescence, earthquake frequency—have

unique histories. As a consequence of individual history, each example of a basic category

has, at some level of resolution, features distinct from other examples of the category.

This insight into the nature of categorization of rocks, volcanoes, river deltas, clouds,

moons, and other objects of interest to the earth (and space) sciences contrasts with the

situation easily noted in chemistry and physics, where fundamental entities come in

Page 6: Revision of Learning Earth Sciences 7-30-04-Nir

6

categories whose members are quite often utterly indistinguishable from each other (xxx

Ault paper may have original author to credit, eg. Hanson): protons, atoms of carbon,

electro-magnetic field. At an important level, both from the perspective of the history and

philosophy of science and from the perspective of the cognitive psychology of learning

science, subjects depart from each other ontologically—in how they frame what is most

salient about reality, with consequences for its representation and analysis (xxx Driver

paper on constructivism cited here).

Quite obviously, most subjects are hybrids of theorizing and categorization, and

the distinction between fundamental entities that have complex and distinguishing

histories (for example, solar bodies) and those basic aspects of reality that differ from

each other in well-determined, rule-governed ways (for example, solar energies) refers

more properly to endpoints of a continuum, rather than to incommensurable opposites.

This acknowledgment brings us squarely to the problem of how distinctive

features characteristic of the earth sciences have particular consequences for teaching and

learning. In turn, the study of teaching and learning the earth sciences may reveal aspects

of the field that distinguish it from others. At the same time, as alluded to above, learning

earth sciences no doubt presents challenges and opportunities that resemble those

common to other sciences.

Five features (best described as “working hypotheses”) appear distinctive of the

challenges and opportunities afforded by learning earth sciences. For each of the five

features below, there appears an example adding a measure of tangibility. Indeed, these

are abstractions intended (or hypothesized) to characterize large swaths of inquiry.

Tentatively, five features of inquiry in the earth sciences useful to examining learning

earth sciences are:

1. The historical approach, pioneered by Charles Lyell and Charles Darwin, to

scientific inquiry (e.g., Darwin’s account of the reefs around coral atolls of the

Pacific: the islands as a sampling distribution across space and through time of

what happens to a volcanic island as it rises and subsides over immense,

unwitnessed durations).

Page 7: Revision of Learning Earth Sciences 7-30-04-Nir

7

2. The concern for complex systems acting over the earth as whole (e.g., the several

“spheres”: hydro, geo, atmos, and their interaction with the biosphere) as well as

analysis of their subsystems on more regional and local scales.

3. The conceptualization of very large-scale phenomena through time and across

space (e.g., “deep time” and the construction of the geologic time scale).

4. The need for visual representation as well as high demand upon spatial reasoning

(e.g., the role of geologic maps, contour maps, and the modeling of structures and

dynamic processes, such as ocean currents and storms, in three dimensions).

5. The integration across scales of solutions to problems (e.g., the validation of

meteor impact hypotheses with evidence gathered across scales from mineral

crystal to regional topography).

Understood in concert, particularly numbers 2 (complex systems), 3 (large scale),

and 5 (integration across scales), these abstracted, yet still quite tangible, features

distinctive of earth science inquiry, suggest themes at a more general level. The most

important of these themes stems from the realization that human action impacts earth

systems on global scales. In brief, people acting collectively have become geologic agents

and their societies can change climates across local, regional, and global scales. Human

communities consume earth resources and depend upon earth systems for the disposal of

wastes. Too obviously, degradation, scarcity, and pollution reach levels that threaten

human communities or interfere with vital “ecosystem services” that undergird

agricultural productivity, maintain habitat and biological diversity, clean both air and

water, and ameliorate climatic variation. Hence, there would appear to be no clear or

useful demarcation between learning earth sciences and learning environmental sciences.

In conclusion, learning earth sciences has distinctive challenges and opportunities.

These begin with attention to the historical methods of inquiry pioneered at the dawn of

geology and continue through appreciation of visualization to data representation and

reasoning. The concept of scale permeates historical methods and visualization tasks, both

as an obstacle to cognitive insight (phenomena happening on vast scales, well beyond the

purview of human experience) and an arbiter of convincing explanation (solutions to

problems on different scales must cohere). When geologic scale and historical complexity

Page 8: Revision of Learning Earth Sciences 7-30-04-Nir

8

are combined with basic ideas from physical and life sciences, earth systems thinking

emerges, with attention to dynamism on global scales of interest. These distinctive

features of the subject (useful as hypotheses defining the range of what constitutes

learning earth sciences, starting from the unrestricted premise that their domain is

everything under our feet and over our heads) combine to produce a whole that is more

than the study of a branch of science. It is a field of learning inseparable from

environmental issues and sciences, with implications for preparing citizens of

democracies for lives of social responsibility.

Our claim is not that this profile of what makes learning earth sciences distinctive

belongs exclusively to learning earth sciences; all disciplines would hopefully make claim

to holding implications for the preparation of citizens for lives of social responsibility in

democratic societies to one degree or another. The point is simply that the general themes

of interdisciplinary study, multidisciplinary study, environmental issues, and relationship

to social responsibility, invariably lie close to the surface when undertaking to learn earth

sciences. Moreover, this immediate relevance to social responsibility, this holism, is,

along with less general and more tangible features such as scale (in two senses: large

scale and crossing scales), historical method, visualization, and systems thinking, an

additional distinctive feature of learning earth sciences. Taken together, the several

working hypotheses about distinctive features of earth sciences—from cognitive

challenges such as visualization, through thinking in terms of dynamic systems, to a

profile that stresses stewardship and citizenship—make for a daunting, yet not intractable,

challenge of characterization.

1.2 Shifting profiles The stature and role of learning earth sciences in keeping with the goal of science

for all has shifted in recent decades. Examples of this shift exist worldwide and these

examples answer questions such as: 1. What status does and should earth science occupy in school science?

2. How has the profile of earth science education changed in recent decades?

3. What does learning earth sciences, when linked to environmental education, offer

as part of science education for all?

Page 9: Revision of Learning Earth Sciences 7-30-04-Nir

9

1.2.1 What status does and should learning earth sciences occupy in school science?

At the level where distinctions between earth and environmental sciences melt

away, there arises another general theme of extraordinary importance: the conduct and

understanding of sciences in social (therefore value) contexts. Because the domains of

earth sciences so demonstrably coincide with environmental issues, this branch of science

education is preeminently and necessarily multidisciplinary, interdisciplinary, and

holistic. This holism encompasses social sciences and humanities as well: the concept of

value and the ethics of caring about the human condition being well addressed within

these fields. The sciences often strive to remain rather silent and noncommittal on the

nature of moral agency and ethical behavior. Yet without knowledge gained from the

sciences, social systems for setting policy and personal decisions about life style

inevitably must blunder. Citizens with knowledge of earth sciences clearly have some

capacity to choose (or hold leaders accountable for choosing) policies in light of their

consequences for earth systems, and therefore, of the potential for society to exist in

profitable harmony with earth resources.

The time has come for science education to situate itself squarely within the

educational conversation about social justice, poverty and wealth, sustainability, and the

human condition (xxx cite NSES content strand, Science in Personal and Social Lives).

This conversation is at the same time one about the nature of democratic institutions for

governing the use of earth resources and impacting earth systems. Because the features

distinctive of the earth sciences so clearly align with and contribute to these aims, it is a

domain of science learning at the cusp of citizenship, a context for learning attitudes

about the role of science in society.

1.2.2 How has the profile of learning earth sciences changed in recent decades?

In many respects education in earth sciences has, in fact, converged upon

environmental education in nations around the globe, as idealized in the preceding

paragraph. In addition, changes in curriculum have often treated the subject more from the

perspective of integration and systems (holism) rather than from the perspectives of

separate disciplines (reductionism). However, the infusion of earth science topics within

the educational system is still a long and complicated task. Comment: it seems to me that reading is more fluent without this section.

Page 10: Revision of Learning Earth Sciences 7-30-04-Nir

10

Reductionist philosophy has historically constrained the introduction of earth

sciences within school science curricula because the importance accorded by reductionist

philosophy to the disciplines of physics, chemistry, and biology. Reduction of science

literacy to competence within these three fields has allowed relatively limited time for

learning earth sciences. The reductionist or disciplinary paradigm works reasonably well

in keeping with the goal of science education as a preparation of a nation’s new

generation of scientists. From the perspective of science for all, it has serious limitations.

The shift towards a science for all paradigm has placed the earth sciences in a

better position within the science curricula of several countries (for example, Israel,

Taiwan, the United Kingdom, and the United States). In 1990, the American Association

for the Advancement of Science published the document Science for All Americans

(AAAS, 1990). This document, a part of the AAAS Project 2061, calls for major reforms

in relation to the goals and teaching and learning strategies of science in schools. The

new “science for all” paradigm perceives the main goal of science education in schools as

a preparation for the nation’s new citizens. Science for All Americans, in essence, defined

minimal levels of scientific literacy for all sciences by outlining objectives for all K-12

students. The Benchmarks for Science Literacy (AAAS, 1993), which followed the

Science for All Americans, document advocates a balance between scientific knowledge,

the processes of science, and the development of personal-social goals (Bybee & Deboer,

1994). The United Kingdom has adopted a similar approach in the new National

Curriculum for England and Wales (Department of Education and Science [DES], 1989).

This new paradigm, which has rapidly influenced other nations all over the world,

gives the earth sciences topics a more central status among the other topics of the science

curricula (Tomorrow 98, 1993). Nations, in part influenced by the socio-political “Green”

movement, have accepted that one of the tasks of science education in schools is to

develop environmental awareness and insight among future citizens. From this point of

view, there is confidence that by studying the earth sciences children might develop such

an understanding.

Another movement influencing the profile of earth sciences education in schools

is the shift from direct instruction towards constructivist pedagogy (Novak teach for

Page 11: Revision of Learning Earth Sciences 7-30-04-Nir

11

understanding xxx, Driver xxx constructivist paper, Driver, Guesne & Tiberghien, 1985;

Osborne & Wittrock, 1985; Bezzi, 1995). The constructivist approach acknowledges that

individuals must construct new understandings in light of personal experience and private

meanings. Constructivists, such as Driver (19xxx), recognize, in addition, the importance

representations of reality (models, diagrams, equations, category systems) play within the

epistemology of a subject. Learners must assume personal responsibility to construct

these representations and compare their thinking with that of others while in pursuit of

the goal of “achieving shared meaning” (Gowin, 19xxx). A constructivist might ask, “Is

the concept adequate to the purpose it serves?” rather than “Is the idea true?” From a

constructivist standpoint, pedagogy ought to engage students in learning meaning through

the use of concepts rather than expecting them to learn ideas simply from listening to

lectures and studying texts.

Conceptual change theory (Posner, xxx 19, Smith xxx) has also exerted a strong

influence over science teaching. Conceptual change theory recognizes that beliefs about

knowledge and conceptions of explanation shape student interests and efforts as they

attempt to learn science. Conceptual change theory, using historical examples of major

shifts in scientific conceptualizations, focuses on the importance of examining the

adequacy of ideas in the context of constructing an explanation or making predictions.

Ideas that are adequate resolve anomalies in plausible ways. In addition, they are

intelligible in terms of current understanding and fruitful in the creation of new

knowledge. Smith has elaborated upon conceptual change theory by describing the

understanding it fosters as “usefulness in a social context” (Smith, 19xxx). This

conception of understanding supports movement towards holistic curriculum while

respecting the disciplinary origin of an idea. The concept of sea floor spreading, for

instance, resolved anomalies in the pattern of magnetic fields recorded on ocean bottom

rocks (a pattern detected incidentally and puzzlingly during attempt to detect enemy

submarines during World War II; cite shifting theory book here 19xxx). Sea floor

spreading made plausible the notion of drifting continents; the concept has proven

enormously fruitful as a component (and precursor) of plate tectonic theory. Now,

understandings of geologic hazards due to seismic and volcanic activity depend upon

Page 12: Revision of Learning Earth Sciences 7-30-04-Nir

12

knowledge of plate tectonic theory. Public policy, from building codes to tsunami alerts,

has turned this knowledge into usefulness in a social context.

Problems, projects, and issues often provide a proper context for promoting

meaningful, even independent, learning. No doubt many students are exposed in their

daily lives or through the mass media to environmental issues such as earthquakes,

volcanoes, global atmospheric changes, journeying to Mars, pollution of the ocean,

sources of fresh water, energy resources, floods, hurricanes, landslides and avalanches,

and much more. These topics are contextual goldmines from a constructivist standpoint,

opportunities to engage students in the construction of meaning through the use of

concepts in personally relevant contexts.

Constructivism and holism have influenced the profile of learning earth sciences in

another and very fundamental way: the growing interest in earth systems education. Many

countries have undertaken to reform science teaching by placing greater emphasis on the

dynamic systems of the earth. These efforts are well presented through two books written

and edited by Vic Mayer: Global science literacy (Mayer, 2002) and Implementing

“Global science literacy” (Mayer, 2002). These books include works from nearly 30

authors from 15 different countries who describe the active implementation of new ideas

about geosciences education often based upon the earth systems approach. Mayer’s work

describes earth systems science as a framework for student learning very robustly.

1.2.3 What does learning earth sciences, when linked to environmental education, offer as

part of science education for all?

As previously noted, earth sciences education has shifted towards an

environmental and interdisciplinary based approach in many parts of the world. As we

enter the 21st century, the environmental perspective has gained great prominence in

western society. This development has been accelerating in view of the understanding

that present human behavior could bring destruction to many of the earth’s ecological

systems. Bybee (1993) uses the expression “the orange light has turned red” to

demonstrate how serious the problem has become. As a person entering the 21st century,

he sees the need to internalize and understand his contact with nature. He wishes to heed

the call to preserve the natural environment and to limit human damage to it.

Page 13: Revision of Learning Earth Sciences 7-30-04-Nir

13

Increasingly, scientific research yields understandings of how the natural systems of the

earth function. New knowledge informs us about the reciprocal relations natural systems

have with human activity. Such knowledge offers guidance to persons wishing to use

resources in an ecologically responsible manner as they improve upon their

understanding of humanity’s relationship with nature as Bybee advocates.

The trend of increased multidisciplinary and interdisciplinary research within the

sciences and across science and other fields has had a conspicuous impact on the earth

sciences. For decades a new field has grown rapidly: “Environmental Geology.” This

field embraces most of the topics traditionally addressed in the earth sciences topics, but

from the perspective of the reciprocal relations natural systems have with human activity

(Tank, 1983; Pickering and Owen, 1994).

In pursuit of a very ambitious agenda intended to reorganize our perspectives of

the earth, Lovelock (1991) points out that the planet earth is composed of several

dynamic, inter-related systems. Feedback loops linking these systems suggest for

Lovelock that the earth functions holistically as a super-organism, at least metaphorically

if not empirically. He states that only by developing a multi-dimensional perspective can

one understand the global picture. In this light, he proposes that environmental research

should be carried out with a multi-disciplinary, holistic approach, as opposed to the

reductionist approach in which scientists specialize in a specific narrow field and fail to

interpret their research within a wider, more holistic context.

Apart from the nearly self-evident proposition that any phenomena can (and

should) be understood from multiple perspectives, there remains the issue of holism

versus reductionism to resolve. Multiple specializations applied to solving a particular

problem may yield to multiple ideas about its solution. Multi-disciplinary, however,

might still mean reductionism—reductionism repeated many times in many ways. Calls

for holism in explanation and for understandings adequate to the task of interpreting

complex systems go beyond calling upon multiple disciplines. Something categorically

different is called for: hierarchical thinking.

There is no need to see holistic and reductionistic approaches in opposition to

each other, though they certainly may be in competition. They are best conceived of as

Page 14: Revision of Learning Earth Sciences 7-30-04-Nir

14

being related through a hierarchy consisting of different levels of organized ideas

responsive to different levels of observation (Ahl and Allen, 19xxx). Explanations for a

phenomenon are typically cast in a language quite different from the one used for an

initial description and hence at a different level of organized ideas. Lower levels offer

explanation in terms of entities more fundamental and universal than the phenomenon in

need of an explanation. Higher levels place the phenomenon of interest in a context that

aids interpretation and indicates it significance.

“Lower” in this sense means translation of the initial description into a new set of

terms proven successful in achieving an explanation; lower implies analysis in terms of

interacting components and their properties. Lower, in effect, means a reduction of the

description of the complex problem to the simplified terms of explanation. These

explanatory terms often have wide generality and usefulness across a range of

phenomena (for example, the application of classical, Newtonian mechanics to problems

of motion).

Interpreting the significance of a phenomenon means placing it in context at a

higher level of description. In addition, thinking hierarchically means giving credence to

the idea that there can be increasing complexity to phenomena. Patterns of interaction

may emerge at higher levels of observation that are not reducible to existing explanatory

terms from a lower level. Thinking responsive to the need to account for emergent

patterns (complexity) may add levels to the hierarchy. This simplified summary of

Hierarchy theory simply underscores the argument that not all that is interesting can be

satisfactorily accounted for with ideas from chemistry and physics; there is more to earth

science (and life science) theorizing than applying and extending basic physical science.

For example, consider the eruption of a Cascade volcano in the Pacific Northwest

of the United States. Magma forms at depth as rock melts; later, a gaseous, ash-laden

cloud bursts forth from the surface of the earth. The explanation of a particular eruption

at a reductionist level depends upon an analysis in terms of geophysics and geochemistry.

Reductionism casts the event in a set of terms (chemical composition, temperature,

pressure, phase change) completely different from those that describe the event itself

(melted rock and ash cloud). The eruptive events, in part, result from phase changes

Page 15: Revision of Learning Earth Sciences 7-30-04-Nir

15

caused by changes in chemistry, pressure and temperature. Measurements of these

fundamental properties replace descriptions of appearance as the analysis proceeds

towards a satisfactory explanation.

Perhaps a massive landslide ensues. Slope failure translates into the conversion of

potential energy to kinetic energy. The reductionist terms (temperature, pressure, phase

change, potential and kinetic energy) do not narrow the analysis. They abstract the

phenomena into formal categories with extraordinarily wide utility—a framework that

works in many contexts where explosive forces operate, volcanic or otherwise. However,

the reductionist path does not bear completely satisfactory fruit.

At the same time, understanding the eruption of a Cascade volcano demands

placing the eruption in a more general, geological context. Cascade volcanoes form an arc

running parallel to a subduction zone at the convergent boundary of the North American

continental and Juan de Fuca oceanic plates. Volcanic arcs themselves call for

explanation and the terms of the explanation are found in plate tectonic theory. This

theory, in turn, is the context that provides significance and interpretation for the eruption

of a Cascade volcano. In briefest terms, a physical imbalance of pressure resulting from

phase change and friction produces magma and causes an eruption (reductionist

explanation); at the same time, plate boundaries are recognized as likely zones of

volcanic activity (contextual interpretation) and this interpretation extends the idea of

cause. In terms of Hierarchy theory (Ahl & Allen, 19xxx; Allen & Hoekstra, 19xxx),

good understanding requires analysis in terms of both higher and lower levels of

observation, with different properties subject to measurement, categorization, and

representation at each level.

Ahl and Allen define a “complex system” as one in which fine details are linked

to large outcomes” (19xxx pp. 29-30). Any system is complex or simple depending upon

whether it is understood in terms of fine details being linked to large outcomes or in

terms of disaggregated parts whose interactions are unambiguous and not subject to the

fine detail/large outcome criterion. Complexity resides both in the nature of the

phenomena and in the explanatory commitment of the observer, for any system may be

observed on a number of levels:

Page 16: Revision of Learning Earth Sciences 7-30-04-Nir

16

In order to describe adequately a complex system, several levels need to

be addressed simultaneously. Levels may be ordered according to the scale

at which each operates, and scale of observation is fixed by the

measurement protocol. Complexity therefore involves relating structures

and processes that are observed at different scales. Reductionism deals

with complexity [sic] by narrowing [the] focus on systems parts so closely

that they are forced to appear simple, at least when disaggregated. By

focusing on issues of scale, levels of organization, levels of observation,

levels of explanation, and relationships between these levels, hierarchy

theory offers an alternative to mechanical, reductionist approaches to

complex systems. (Ahl & Allen, p. 30).

We suspect that as the ideas of Hierarchy theory make inroads into the sciences,

there will be profound implications for what science for all becomes. Earth systems and

ecosystems are clearly domains where ideas of scale, levels of organization, and levels of

explanation hold salience. Hierarchy theory underscores the importance of high level

characterization of phenomena (holism) together with the value of low level ones

(reductionism) to meaningful understanding. Combined with constructivist paradigm, the

key question for good explanation becomes, “Are the levels of the hierarchy adequate to

the need for understanding?” The implication for pedagogy is that context matters just as

much as mechanistic explanation.

Systems thinking, hierarchy theory, holistic explanation, and attention to scale and

complexity bind learning earth sciences and environmental sciences at an abstract level.

This is an ambitious agenda for science for all. Indeed, the environmental imperative has

achieved a central position in the field of earth science education. Several publications

that appeared during the 90’s (e.g., Mayer and Armstrong, 1990; Brody, 1994; Mayer,

1995 and Orion, 1996) stated that one of the advantages of studying the earth sciences is

the development of environmental awareness and insight. Earth sciences offer the

student—the future citizen—the knowledge and the ability to draw conclusions regarding

issues such as: conservation of energy and water as well as proper utilization of global

resources. In addition, the teaching of earth sciences may raise students’ awareness of

Page 17: Revision of Learning Earth Sciences 7-30-04-Nir

17

what is happening around them, in their local environment, in their country, and in the

world. Likewise, students who understand the environment in which they live and the

processes taking place in it might better know how to preserve it and how to behave

within it. They could develop better tools to judge and evaluate the changes taking place

in their environment.

1.2.4 Earth system science

In the first international conference on earth science education, which convened in

1993 in England, the proposal to reinforce the environmental aspect of earth science drew

widespread support from the participants (Carpenter, 1996; Orion, 1996; Mayer, 1996).

The title of the second international conference on earth science education, which

convened in 1997 in Hawaii, was: “Learning about the earth as a system” (IGEO, 1997).

Orion and Fortner (2003) suggested the “Earth systems approach” as a holistic

framework for science curricula that integrates earth science education together with

environmental education. It was suggested that the starting point for this integrated model

is the natural world, which is understood by studying the four earth systems: geosphere,

hydrosphere, atmosphere and biosphere. The study of each subsystem is organized

around geochemical and bio-geochemical cycles including the rock cycle; the water

cycle; the food chain; the carbon cycle; and energy cycles (which are included in all of

these cycles). The study of these cycles also emphasizes the relationships between the

different subsystems via transitions of matter and energy from one subsystem to another

(based on laws of conservation). Such natural cycles should be discussed within the

context of their influence on people's daily life, rather than being isolated to their specific

scientific domains. People are introduced in this model as unique, but integral parts of the

biosphere. There are many differences, of course, between people and other organisms;

this model emphasizes two of these: People's ability to produce tools, which is termed

technology, and People's natural curiosity and ability to investigate his environment,

which is called science. There is a close relationship between these two characteristics—

science and technology—as the progress of one contributes to the development of the

other. This model also connects the natural world and technology together, since all raw

materials originate from the earth systems (predominantly from the geosphere and

Page 18: Revision of Learning Earth Sciences 7-30-04-Nir

18

biosphere). The connection between man-made materials and the natural systems of the

earth closes this cyclic model. This connection symbolizes environmental quality, which

is profoundly effected by technology.

The cyclic aspect of the model emphasizes the understanding that society is a part

of the Earth's natural system and thus any manipulation in one part of this complex

system might adversely affect people. It is important to note that, in opposition to the

current models for teaching science, this model does not utilize physics or chemistry as

the basis of the curricular sequence. Instead, this model suggests that science studies

should start from the concrete world and utilize physics and chemistry as tools for

understanding science at a deeper and more abstract level.

The development of such new environmental-based science curricula includes the

definition of educational goals and objectives. The main educational goal is the

development of environmental insight, which includes the following two principles:

We live in a cycling world that is built upon a series of sub-systems

(geosphere, hydrosphere, biosphere, and atmosphere) which interact through

an exchange of energy and materials;

Understanding that people are a part of nature, and thus must act in harmony

with its laws of cycling. In order to develop environmental literacy, we

develop most of our Earth science programs within a systems framework; e.g.,

The Rock Cycle; The Water Cycle; The Carbon Cycle.

Thus, it is possible to break down the broad subject of environmental insight into

a set of more focused research problems whose purposes are:

1. To determine the levels of knowledge of school students on the subject of

environment.

2. To test student understanding of the concepts of cycles and systems, as well to

identify misconceptions concerning the Earth's geo-bio-chemical cycles.

3. To explore the “deep time” concept in relation to the development of

environmental insight.

About a decade after introducing of the earth systems approach into the earth

sciences education community, Vic Mayer extended this idea further with the introduction

Page 19: Revision of Learning Earth Sciences 7-30-04-Nir

19

of Global Science Literacy (GSL; Mayer, 1997, 2002; 2003). His approach expands on the

argument for a new type of science curriculum for secondary schools. Instead of being

based on each of the major disciplines, as are almost all current science curricula, Mayer

argues that curricula should be conceptually organized around the “Earth System,”

including the science methodology of the system sciences, and capitalize on the cross-

cultural characteristics of science to establish greater understanding of the contributions of

all cultures. The Earth System approach indeed embodies holism in curriculum design.

Holism not only embraces the Earth System concept, but also extends learning

earth sciences into environmental domains and the context of social and political debate.

However, do scientists practice holistic science? Yes; holism exists as a basic goal of

research within the earth and space sciences community as the following example most

notably illustrates

The year 2003 witnessed in the United States the inauguration of an

unprecedented multi-disciplinary, earth and space science program of research:

EarthScope. The National Science Foundation (NSF), the United States Geological

Society (USGS), and the National Aeronautics and Space Administration (NASA)

together with a number of prestigious research universities have combined resources to

advance knowledge about North America’s “three-dimensional structure, and changes in

that structure, through time. By integrating scientific information derived from geology,

seismology, geodesy, and remote sensing, EarthScope will yield a comprehensive, time-

dependent picture of the continent beyond that which any single discipline can achieve.

Cutting-edge land- and space-based technologies will make it possible for the first time to

resolve Earth structure and measure deformation in real-time at continental scales. These

measurements will permit us to relate processes in Earth’s interior to their surface

expressions, including faults and volcanoes.” (EarthScope project plan 200xxx pp. 1-2)

EarthScope organizers fully expect to impact school and museum science in

substantial ways, both as an example of integrated science and a resource for real world

data. EarthScope is the preeminent example of “holistic” work in earth and space science.

Its education and outreach components are as essential as its primary investigations

Page 20: Revision of Learning Earth Sciences 7-30-04-Nir

20

because among its fundamental goals is achieving understandings of volcanoes and

earthquakes needed to promote public safety, commerce, and engineering.

The profile of learning earth sciences continues to shift as does the practice of

earth and space science: from isolated, disciplinary agendas, to integrated research with

outcomes of interest to the public; from separate concern for earth history and systems, to

convergence upon themes essential to environmental science and education; from less

reductionism to more holism; from direct instruction based upon text materials to

constructivist pedagogy with access to real world data.

2. Learning earth sciences

2.1 The main characteristics of the earth science education in schools Earth science education worldwide has undergone a process of revival during the

past decade. Since 1993 four international conferences on geoscience education have

been conducted in Europe, the USA, Australia, and Canada (Stow and McCall, 1996;

IGEO, 1997; IGEO, 2000; IGEO, 2003). Now, at the beginning of the 21st century, it is

well accepted among earth science educators that the overall purpose of earth science

education for ages 5-19 is to educate for citizenship rather than, or as well as, to prepare

students to become professional geoscientists. The aim is to maximize personal

development and to increase the cognitive and ethical understanding of all citizens with

respect to the workings of the global environment and the exploration for and the

exploitation of resources.

This goal will be achieved only if the very high educational potential of the earth

sciences becomes realized in schools, beginning with their potential to illustrate scientific

thinking, and continuing through several other distinctive features of these fields:

responsiveness to the importance of historical explanation, concern for complex systems,

dedication to solving problems at large and across many scales in time and space,

dependence upon spatial visualization, and convergence upon environmental study

2.1.1 Scientific thinking

The earth sciences encourage learners to be scientific detectives through the

development of intellectual and practical skills, attitudes, and methodologies that relate to

Page 21: Revision of Learning Earth Sciences 7-30-04-Nir

21

problem solving. Among these attributes of scientific thinking are: the role of conjecture;

the creation of multiple working hypotheses; the design of investigations; the making of

relevant observations; the development of a variety of recording skills: predicting, testing

of ideas, inferring, interpreting and the communicating of findings; induction; deduction;

and falsification.

To unravel processes that took place millions of years ago, geologists have

developed a distinctive way of thinking that involves retrospection. Geological inquiry

applies knowledge of present day processes in order to draw conclusions about the rocks

of the Earth's past. These conclusions clarify the picture of the materials, processes and

environments of past times. This particular contribution of retrospection to the students'

cognitive abilities together with the development of spatial visualization skills to an

extant that is rare in other science disciplines is almost unique to learning earth sciences

(Kali and Orion, 1996).

2.1.2 The dimensions of time and space

The earth sciences use and develop concepts common to the traditional sciences,

and some which are uniquely their own, in a conceptual framework that ranges from local

to global and involves the depths of time and the vastness of space. This large time range,

measured in millions and billions of years, and the huge spatial domains above and under

the Earth's surface are the objects of study: inner space, near space and even outer space

are involved (when comparisons with other solar and planetary systems are being

considered).

2.1.3 Environmental orientation

Today, more then ever, there is a worldwide recognition that living in peace with

our environment is more than just a slogan; it is an existential need. It is also agreed that

the understanding of each of the earth’s sub-systems and the environment as a whole is

indispensable in order to live in peace with the environment. This understanding is

actually what science all about.

Learning earth sciences has a major part to play in the environmental education of

society. It gives the students—our future citizens—the knowledge and the ability to think

about the importance and interrelationships of the lithosphere, atmosphere and

Page 22: Revision of Learning Earth Sciences 7-30-04-Nir

22

hydrosphere as well as subjects such as the utilization and conservation of energy, water,

and material resources.

There are two main schools of environmental studies. Both approaches examine

the interrelationships between people and the physical environment, however they differ

in their perspectives. One school is more concerned with the understanding of the

physical environment and studies primarily the five interacting Earth's subsystems or

spheres: atmosphere, biosphere, cryosphere (ice), hydrosphere and lithosphere.

The other school is more concerned with the environmental hazards from the

perspective of their impacts on humanity. This approach prioritizes the interrelation

between energy and environment Human society, in this approach, is an integral part of

the systems of the earth. Technology has a dual role in the interaction between society

and environment. On the one hand, technological revolution and energy exploitation have

dramatically degraded the capacities of many ecosystems. On the other hand, new

technologies can help in limiting environmental hazards and in providing alternative

energy resources.

Most importantly, learning earth sciences may deepen students’ awareness of the

physical surroundings of their homeland and enable them to participate in an informed

way in contentious matters such as exploitation versus conservation.

2.1.4 The interdisciplinary nature of the earth sciences

The earth sciences, by their very nature, form an interdisciplinary approach to

problems (recall the example of EarthScope above). Physical and chemical processes and

principles and biological processes and environmental understanding are needed to

explain geological phenomena, both present and the past. Therefore, the earth sciences

demonstrate the practical uses of physics and chemistry in our daily life. They also have

close relationships with biological topics such as evolution, ecology and the links

between rocks, soil, flora, and fauna.

The concept of the “Earth System” provides a conceptual model for curriculum

developers to use in promoting integrated science programs (Mayer, 1995). The majority

of research efforts of all the science disciplines relate to Planet Earth. Therefore the Earth

Page 23: Revision of Learning Earth Sciences 7-30-04-Nir

23

System approach is perhaps the best and most appropriate focus for such integrated

science courses.

2.1.5 From the kindergarten to the high-school

By classifying the learning concepts from the “concrete to the abstract,” topics

from the earth sciences can be presented appropriately to students of all levels of ability,

achievement, and age from the kindergarten to the high school. On one hand the earth

sciences deal with very concrete phenomena that one can learn through direct interaction

in the lab and/or the field.

However, at the other end of the concrete-abstract continuum there are very

abstract phenomena and high order skills that are involved in the understanding of many

earth sciences concepts. Gobert, (2000) in her study of students’ models of plate tectonics

and the interior of the earth, demonstrated that the plate tectonics model placed high

demands upon abstract thinking. She pointed out four challenging demands of this

abstract thinking:

1. The interior of the earth and the processes that take place there are outside of direct

experience.

2. The size scale of plate tectonic processes are difficult for the human mind to grasp.

3. The time scale of geological processes is beyond the perception of the human mind.

4. Comprehension requires the integration of several different types of information,

namely spatial, causal, and dynamic information.

2.1.6 Variety of learning environments in the teaching-learning process

Teaching earth sciences allows the integration of formal teaching into several

learning environments: the classroom, the laboratory, the field, the museum and the

industrial site. Former teaching strategies, which were essentially “lectures on the

subject,” have given way to student investigative work of a wide and varied kind. These

latter strategies make facts and processes part of tangible experience and introduce

students to field research methodology (Orion, 1993; Orion & Hofstein, 1994).

2.1.7 The relevance of earth sciences to our daily life

Generally, there is widespread agreement among educators that teaching science

in relation to an individual’s daily life, and to the local environment, is a very powerful

Page 24: Revision of Learning Earth Sciences 7-30-04-Nir

24

teaching strategy. Understandings learned from the earth sciences directly apply to many

important aspects of life: for example, the study of natural disasters (earthquakes,

volcanic eruptions, landslides, hurricanes, sea defenses, etc.), the mining of raw materials

(potable water, coal, oil, gas, construction materials etc.), the control of energy resources

(fossil fuels, nuclear, tidal, solar etc.) and the care of the local, regional, and global

environment (e.g., the preservation of wetlands or tropical forests, the greenhouse effect,

the ozone layer, rising sea levels, etc.). Xxx add NSES standard citation here as well

2.2. The cognitive aspects of learning earth sciences. The following section describes several traditions of research about learning earth

sciences. Collectively, these studies inform those whose aims are to fulfill the educational

potential of learning earth sciences as part of science for all. We have grouped studies of

cognitive learning in earth sciences as examples of alternative frameworks research,

studies of spatial visualization, examination of temporal thinking, and investigations of

systems thinking. A synthesis of these traditions ends the section on cognitive aspects of

learning earth sciences.

2.2.1 Alternative frameworks of learners concerning earth sciences concepts

The constructivist paradigm has dominated the field of science education during

the last two decades of the 20th century and the beginning of the 21st century. The most

predominant research produced during the constructivist era no doubt has been the

extensive studies of misconceptions, preconceptions, naive ideas, and, in more general

terms, alternative frameworks of students in relation to all aspects of the science curricula.

Although there are relatively few published studies of students’ alternative

frameworks in earth sciences education, there have emerged some generalized findings

and patterns (see Ault, 1994xxx, for an earlier review of this literature and related studies

of “expert and novice” styles of solving earth science problems). At least four independent

studies have explored each of the following four areas of learning earth sciences:

1. Studies of students’ perceptions of processes and mechanisms of geospheric

change. This area includes subjects such as plate tectonics, the rock cycle,

earthquakes, erosion, etc. (Ault, 1984; Happs, 1985; Ross and Shuell, 1993; Bezzi

and Happs, 1994; Lillo, 1994; Marques and Thompson, 1997; Schoon, 1989;

Page 25: Revision of Learning Earth Sciences 7-30-04-Nir

25

Gobert and Clement, 1998; Stofflett, 1994; Dove, 1997; Dove, 1999; Gobert,

2000; Kali, Orion and Elon, 2003; Libarkin, Anderson, Boone, Dahl and Kurdziel,

2004).

2. Studies of students’ understanding and conceptions of the Earth’s interior

(DeLaughter, Stein, Stein, and Bain, 1998; Gobert and Clement, 1998; Marques

and Thompson, 1997; Lilio, 1994; Nottis and Ketter, 1999; King, 2000; Beilfuss,

Dickerson, Libarkin and Boone, 2004). These studies cover K-12 students’,

undergraduates, and practicing teachers.

3. Studies of students’ and teachers’ perceptions of geological deep time (Happs,

1982; Marques, 1988; Oversby, 1996; Schoon, 1989; Marques and Thompson,

1997; Noonan-Pulling and Good, 1999; Trend, 1997, 1998, 2000; Dodick and

Orion, 2003a, 2003b)

4. Studies of students’ and teachers’ perceptions of hydrospheric processes and the

water cycle (Meyer, 1987; Fetherstonhaugh and Bezzi, 1992; Brody, 1994; Barker

1998; Taiwo, Ray, Motswiri and Masene, 1999; Agelidou, Balafoutas and

Gialamas, 2001; Dickerson, 2003; Ben-zvi-Assaraf and Orion, 2004; Beilfuss,

Libarkin and Boone, 2004). A detailed analysis of the literature concerning the

hydrosphere appears later in section 2.4.

Review of the above studies indicates that children, adolescents, and adults hold

alternative frameworks in relation to almost every topic in the earth sciences. These

alternative frameworks are seen across nations, cultures, and ages. Some of these

frameworks are no doubt preconceptions that emerge as students encounter difficult

abstractions about the earth in conflict with the scale of their everyday perceptions. For

example, static views of the geosphere and groundwater are common—dynamic insights

are less common. Students overestimate the effect of external forces of the earth observed

directly at its surface and fail to appreciate the importance of the internal forces shaping

structures. They struggle with their perceptions of geological time and spatial phenomena.

Finally, they often misconceive the interior of the earth and the state of matter within the

interior of the earth.

Page 26: Revision of Learning Earth Sciences 7-30-04-Nir

26

As previously explained, such preconceptions are predictable; however, review of

these studies holds another striking conclusion: the same preconceptions appear across

grade levels, from kindergarten to college. These studies indicate that schooling all over

the world has influenced only in a limited way the ability of students to construct

scientifically sound conceptions of the earth, congruent rather than in conflict with

knowledge from the earth sciences.

Sadly, the literature suggests that many teachers hold the same alternative

frameworks as their students and that even text materials foster misconceptions. Thus, it

seems that earth science education in many countries is trapped in a cycle of ineffective

instruction and inadequate learning—with preconceptions and misconceptions dominating

learning earth sciences. Research studies about earth science education have the potential

to break this non-productive cycle, but only if they are integrated with curriculum design

and implementation and in keeping with the changing profile of the earth sciences (e.g.,

Earth Systems approach, integration of subjects, convergence on environmental studies).

2.2.2 Spatial visualization

Teaching and learning earth sciences at all levels relies upon spatial reasoning of

many kinds. The phenomena of interest have spatial extent on many scales. Sometimes

their geometries are simple, but exist on grand scales: spiral structures of galaxies; gyres

in ocean circulation; axes of synclinal folds. Sometimes the geometries are confusing: the

intersection of complex topography with complicated stratigraphy, for example.

Sometimes the surfaces of interest are mapped indirectly: gravitational anomalies and

magnetic fields. And most confusingly, the geometries change with time.

In addition to the phenomena of interest and their challenging geometries, there

are the representations used by earth scientists to record and study them. These

representations place demand upon spatial reasoning as well. There are contour maps of a

host of phenomena to master, from topography to pressure gradients, from glacial

thickness to stress fields. Geologic maps contour time. Maps are two dimensional

representations yet often include data about three dimensional structures. Seeing “through

the surface” to visualize three dimensional structure is indeed challenging. Sometimes,

visualization requires skill at projecting structures from three dimensions onto two.

Page 27: Revision of Learning Earth Sciences 7-30-04-Nir

27

Consider also that visual patterns among sedimentary rocks record in three dimensions

events through time. In geology, visual pattern is the key to unlocking temporal puzzles.

While the basic dependence of geoscientists on spatial abilities has long been

recognized (Chadwick, 1978), the geoscience education community has only begun to

explore the vast array of spatial abilities which students must bring to bear in order to

understand essential geoscience concepts at all educational levels (McAuliffe, Hall-

Wallace, Piburn, Reynolds and Leedy, 2000). These spatial reasoning abilities may, in

fact, be quite distinct from the spatial abilities commonly associated with tasks in

learning chemistry (Dori and Barak, 2001; Pribyl and Bodner, 1987), in learning physics

(Pallrand and Seeber, 1984), and in learning engineering (Hsi, Linn and Bell, 1997).

The spatial objects that are studied in the geological sciences are usually large

enough to walk in physically (the field learning environment). They can also be readily

represented by block models and more sophisticated renderings in a virtual setting. In the

earth sciences, these blocks are not only visualized, but rotated, inspected, and modified

to reflect temporal changes.

The development of an understanding of deep geologic time by students has also

been shown to be related to aspects of spatial cognition (Dodick and Orion, 2003a,

2003b). There is also evidence showing that the outdoor field learning environment

specifically enhances the ability to connect static objects in the field (e.g. layers of

sedimentary rocks) into a coherent narrative which contains an understanding of change

through time at that given location (Orion, Ben-Chaim and Kali, (1997); Riggs and

Tretinjak, 2003). Other evidence presented below suggests a strong link between key

spatial abilities as students learn to investigate field problems at more advanced levels in

geology.

Kali and Orion (1996) characterized the specific spatial abilities required for the

study of basic structural geology. To do this they developed a geologic spatial ability test

(GeoSAT), in which students were required to draw two-dimensional cross-sections of

geological structures that were represented as block-diagrams. Their outcomes indicate

that the problem-solving involved in GeoSAT require a special type of spatial

visualization which they named VPA (Visual Penetration Ability). Spatial visualization is

Page 28: Revision of Learning Earth Sciences 7-30-04-Nir

28

defined as the ability to create a mental image from a “pictorially presented object” and to

operate different mental manipulations on those images. The manipulations usually

referred to are mental rotation and mental translation. In contrast, the manipulations

involved in VPA are to visually penetrate into a three dimensional mental image in order

to envision two dimensional cross-sections.

Based on their findings about VPA, Kali and Orion developed Geo3D, a software

package designed to assist high-school students in developing their VPA and in acquiring

the skills needed for understanding basic structural geology (Kali & Orion, 1997). Using

four case-studies, they showed that even with a short-term interaction with the software,

students significantly improved their ability to solve the problems involved in GeoSAT.

The advantage of technology in improving learners’ capability to solve problems that

require spatial skills has also been shown by Hsi, Linn & Bell (1997) in the area of

engineering. Interestingly, these authors also indicated a relatively short time-span in

which students acquired their spatial skills using the computer-based tools involved.

The NSF-funded Hidden Earth Project (Reynolds, Piburn, & Tewksbury)

successfully investigated the role of spatial visualization in an introductory geology

course. This project developed web-based versions of three standard visualization tests

(Cube Rotation, Spatial Visualization, and Hidden Figures) and a geospatial test,

containing items of the more visual aspects of geology, such as visualization of

topography from contour maps. Reynolds and others developed innovative instructional

modules for (1) Visualizing Topography, and (2) Interactive 3D Geologic Blocks. An

experimental group used these modules, and the control group did not. Although all

subjects profited from both the control and the experimental conditions, the effectiveness

of the treatment experienced by the experimental group was confirmed using Analysis of

Variance and a comparison of normalized gain scores. Very powerful gender effects have

also been demonstrated, with the experiment equalizing the performance of males and

females in a case where the performance of males was initially superior to that of

females. The experiment also was very effective at improving scores and lowering times

to completion on the spatial visualization test.

Page 29: Revision of Learning Earth Sciences 7-30-04-Nir

29

As part of the Hidden Earth Curriculum Project, Reynolds, Piburn, and Clark

(2004) conducted a detailed investigation of college student’s pre-instructional

knowledge, skills, and misconceptions about visualizing topography from contour maps.

Students completed pre-tests and post-tests, and selected students were interviewed to

assess what their initial skills and strategies were. These interviews exposed several

previously unrecognized misconceptions about topographic maps, and a Topographic

Visualization Instrument was developed to see how prevalent these misconceptions were

in a broader sample of students.

When one considers spatial cognition in a geoscientific problem solving context,

one can no longer only consider the static, two dimensional representation of three-

dimensional objects so frequently used as test items in traditional spatial abilities tests,

but must also begin to examine the temporal evolution of student understanding as they

explore a real 3D object and extract a temporal history from these spatially extended

geologic features.

Field investigations and simulated field work all involve problem solving when

properly constructed using an inquiry-based structure, and the incorporation of new

information continually shapes the investigations as work on the problem progresses. The

same can be said of well-constructed classroom computer-based interventions such as the

materials in Geo3D (Kali & Orion, 1997) and Hidden Earth (Reynolds, Piburn, Leedy,

McAuliffe, Birk, & Johnson, 2002).

Studies in geoscience education for Native American students show that students

from certain cultural backgrounds more readily learn geoscience in a field setting than do

others (Riggs, 2003; Riggs & Semken, 2001). It stands to reason that there is probably

some robust connection between place-based Indigenous cultures and field-based

learning via spatial abilities. Clearly, experience plays an essential role in developing

spatial reasoning ability.

2.2.3 Temporal thinking

In the history of geology there have been two discoveries, plate tectonics and

geological time, which have literally defined the way geologists view the earth.

Geological time means the understanding (aptly termed by John McPhee in 1980 as

Page 30: Revision of Learning Earth Sciences 7-30-04-Nir

30

“deep time”) that the universe has existed for countless millennia, and that humanity’s

earthly dominion is confined to the last milliseconds of the metaphorical geological

clock.

The influence of geological time is felt in a variety of scientific disciplines

including geology, cosmology, and evolutionary biology (Dodick and Orion, 2003c).

Thus, any scientist or student that wants to master any of these subjects must have a good

understanding of geological time. However, after reviewing the science education

research literature, Roseman (1992, p. 218) noted that there “was next to nothing

about…how kids’ understanding of notions of systems, scale or models develop over

time.” Since that time, there have been several large-scale studies of how students

understand this concept.

In general, the studies that have been completed on how students understand

geological time can be roughly divided into two groups: “event-based studies” and

“logic-based studies.”

Event-based studies include all research that surveys student understanding of the

vast duration of “deep time” (that is, time beginning with the formation of the earth or the

universe). In such studies, the general task is sequencing a series of events (for example

the first appearance of life on earth) absolutely, along a time-line, or relatively using

picture-sorting tasks. Often in such sequencing tasks, the subject is asked to justify their

reasons for their proposed temporal order. Such studies include: Noonan-Pulling and

Good’s (1999) research on the understanding of the origins of earth and life amongst

junior high students; a similar study by Marques and Thompson (1997) with Portuguese

students; and Trends’ studies on the conception of geological time amongst 10-11-year-

old children (Trend, 1997; 1998), 17-year-old students (Trend, 2001b), as well as amongst

primary teacher trainees (Trend 2000; 2001a).

Such studies largely reflect the subjects’ knowledge of particular events and most

of them involved qualitative research (structured interviews) with small sample groups. A

survey of the science education literature indicates that there has never been a large-scale

quantitative study of older student’s (junior high to senior high) understanding of

geological time.

Page 31: Revision of Learning Earth Sciences 7-30-04-Nir

31

In logic-based studies, the researcher is interested on the cognitive processes

undergone by students when confronted with problems of geologic time. It might be

added that such studies are more concerned with probing the subject’s logical processes

rather than their knowledge of earth science.

This approach is seen in the work of Ault (1981; 1982) and Dodick and Orion

(2003a; 2003b) Ault interviewed a group of forty students from grades kindergarten,

two, four, and six using a series of puzzles which tested how they understood (and could

reconstruct) a series of geological strata. Based on Zwart’s (1976) suggestion that the

development of people’s temporal understanding lies in the before and after relationship,

Ault (1981) theorized that children organize geological time relationally.

Based on his findings, Ault (1981; 1982) claimed that young (grade 2-6)

children’s concept of conventional time in a logical sense (reasoning about before and

after) was no impediment towards their understanding of geologic events. Indeed, many

of the children in his test group were successful at solving puzzles involving skills

necessary to an understanding of the logic, though not the extent, of geological time, such

as superposition and correlation. Nonetheless, in the field, these same children had

difficulties in solving similar types of problems, indicating that there was little transfer

from classroom problems to authentic geological settings. Children believed rock layers

in the field to be old based upon being dark or crumbly—not based upon their position in

a series of strata.

These difficulties can be traced to Ault’s (1981) research design, which was

influenced by Piaget’s (1969) work on time cognition. According to Piaget, a young

child’s understanding of time is tightly bound to his or her concept of motion, thus, the

research problems he used were taken from physics. However, the geological science

builds its knowledge of time through visual interpretation of static entities (formations,

fossils; Frodeman, 1995; 1996). Indeed there is no reason to suggest that an

understanding of the (logical) relationships amongst strata should necessarily allow one

to both conceptualize and internalize the entirety of geological time. In contrast, it is

possible that the two forms of understanding can be studied as separate entities.

Page 32: Revision of Learning Earth Sciences 7-30-04-Nir

32

Dodick and Orion (2003a, 2003b) conducted a large-scale study with junior high

and high school students, which included validated, reliable quantitative tools. In this

study, geological time was divided into two different concepts:

1. A (passive) temporal framework in which large scale geological events occur.

Such understanding depends upon building connections between events and

time. In the cognitive literature this is comparable to Friedman’s (1982)

associative networks, a system of temporal processing used for storing

information on points in time. By this reasoning, an understanding of

geological time should be mitigated by a person’s knowledge of such events.

2. An (active) logical understanding of geological time used to reconstruct past

environments and organisms based on a series of scientific principles. This is

similar to the work done in logic based studies, noted above. Based on this

definition, it might seem that students unfamiliar with geology might be

unable to reconstruct a depositional system; however, in structure, geo-logic is

comparable to Montagnero’s (1992; 1996) model of “diachronic thinking.” He

defines “diachronic thinking” as the capacity to represent transformations over

time; such thinking is activated, for example, when a child attempts to

reconstruct the growth (and decay) cycle of a tree.

Montagenro (1996) argues that there are four schemes, which are activated when

one attempts to reconstruct transformational sequences. In this study, three have been

translated to the logical skills needed to solve temporal problems involving geological

strata:

1. Transformation: This scheme defines a principle of change, whether

qualitative or quantitative. In geology it is understood through the principle of

actualism (i.e. “the present as key to the past”).

2. Temporal Organization: This scheme defines the sequential order of stages

in a transformational process. In geology, principles based on the three

dimensional relationship amongst strata (ex: superposition) are used in

determining temporal organization.

Page 33: Revision of Learning Earth Sciences 7-30-04-Nir

33

3. Interstage Linkage: The connections between the successive stages of

transformational phenomena. In geology such stages are reconstructed via the

combination of actualism and causal reasoning.

For the purposes of this research, Montagenro designed a specialized (validated)

instrument, the GeoTAT which consisted of a series of open puzzles which tested the

subject’s understanding of diachronic schemes as applied to geological settings.

In addition, two other questionnaires were distributed to sub-units of this

population to answer questions that arose through the use of the GeoTAT: (a) a Time-

Spatial Test (or TST), which tested the possibility that spatial thinking influences

temporal thinking. (b) a Stratigraphic Factors Test (SFT) which tested the influence of

(geological strata) dimensions on students’ temporal understanding. In addition,

qualitative research was pursued in the classroom and field by studying and interviewing

students who were studying geology and paleontology as part of their matriculation

studies.

As a result of this study it was possible to construct a model of temporal thinking

with the key features that influence a subject’s ability to reconstruct geological features in

time (or “reconstructive” thinking as we term it):

1. The transformation scheme which influences the other two diachronic

schemes.

2. Knowledge, most importantly empirical knowledge (such as the

relationship between environment and rock type) and organizational

knowledge (i.e. dimensional change).

3. Extra-cognitive factors such as spatial-visual ability which influence how

a subject temporally organizes 3-dimensional structures such as geological

strata.

Amongst students who were not taking geology as part of their school program it

was seen that there was a significant difference between samples composed of high

school and 9th grade students (on the one hand) and 7th grade students (on the other) in

their ability to understand geological phenomena using diachronic thinking. This suggests

that somewhere between grades 7-8 it should be possible to start teaching some of the

Page 34: Revision of Learning Earth Sciences 7-30-04-Nir

34

logical principles permitting one to reconstruct geological structures. These include

complex superposition (consisting of tilted strata), and correlation (two outcrop

problems) which rely on the use of isolated diachronic schemes, as well the integrated use

of all the diachronic schemes to solve complex problems of deposition.

Moreover, this research shows that the ability to think diachronically can be

improved if practiced in the context of learning earth sciences. A comparison of high

school (grade 11-12) geology and non-geology majors indicated that the former group

held a significant advantage over the later in solving problems involving “diachronic

thinking.” This relationship was especially strengthened by the second year of geological

study (grade 12), with the key factor in this improvement (probably) being exposure to

fieldwork. Fieldwork both improved students’ ability to understand the 3-dimensional

factors influencing temporal organization and provided them with experience in learning

about the types of evidence that are critical in reconstructing a transformational sequence.

The work of Riggs and Tretinjak (2003) supports this finding. Riggs and

Tretinjak studied a non-majors course in earth science for pre-service elementary school

teachers. They were able to shows that integrated field investigations enhance higher-

order content knowledge in geoscience, specifically the understanding of environmental

change through time as read from the sedimentary rock record. Prior to the field trip

students could identify past environments from sedimentary rock, but only after

completing the fieldwork unit were they able to understand these rocks as a dynamic

temporal/historical record. This is consistent with the findings of Dodick and Orion

(2003a; 2003b) who found a correlation between the understanding of geologic time and

spatial ability, which in turn implies that well-designed geologic field work will enhance

both, even for non-majors. There currently is no comparable data of this nature for

geoscience majors, nor do we fully understand the reasons for this correlation amongst

temporal/spatial/and field abilities.

In addition to the studies mentioned above, one might add the small body of

research which catalogues general misconceptions in geology, and includes within its

parameters problems related to geological time (Happs, 1982; Marques, 1988; Oversby,

1996; Schoon, 1989). Finally, one might note those works which have focused on the

Page 35: Revision of Learning Earth Sciences 7-30-04-Nir

35

practical elements of teaching the scale of time (Everitt, Good and Pankiewicz, 1996;

Hume, 1978; Metzger, 1992; Ritger and Cummins, 1991; Rowland, 1983; Spencer-

Cervato and Day, 2000). Unfortunately, these teaching models have never been critically

evaluated, so they are of untested value to the pedagogic literature.

2.2.4 Systems-thinking

Current earth science education is characterized by a shift towards a systems

approach to teaching and curriculum development (Mayer, 2002). Earth science educators

call for reexamining the teaching and learning of traditional earth science in the context of

the many environmental and social issues facing the planet (IGEO, 1997). Orion (1998,

2002) claimed that since the natural environment is a system of interacting natural

subsystems, students should understand that any manipulation in one part of this complex

system might cause effects in another part, sometimes in ways quite unexpected. The

understanding of physical systems such as the earth is also based on the ability to expand

the systems’ borders and expose hidden dimensions and interactions. Viewing the

expanded system of the earth reveals how groundwater and the atmosphere interact with

the geosphere, for example. Moreover, analyzing environmental problems such as

groundwater pollution involves questions such as: What was the cause of the groundwater

pollution? What will be the outcome of the pollutants in the groundwater system? How

could humans be affected? How long will those chemicals stay within the rocks?

The ability to deal with such questions requires backwards (retrospection) and

forwards (prediction) thinking skills. Mayer (2002) emphasizes that the development of

systems-thinking about the different earth systems, i.e., the geosphere, the hydrosphere,

the atmosphere and the biosphere (including humanity), is fundamental to environmental

literacy.

Systems-thinking is regarded as a type of higher order thinking required in

scientific, technological and everyday domains. Therefore, researchers in many fields

have studied systems-thinking extensively; for example: in the social sciences, (e.g.,

Senge, 1998), in medicine (e.g., Faughnan & Elson, 1998), in psychology (e.g., Emery,

1992), in decision making (e.g., Graczyk, 1993), in project management (e.g., Lewis,

Page 36: Revision of Learning Earth Sciences 7-30-04-Nir

36

1998), in engineering (e.g., Fordyce, 1988), and in mathematics (e.g. Ossimitz, 2000).

However, little is known about systems-thinking in the context of science education.

During the late 90's and the beginning of this decade three studies were conducted

in the Weizmann Institute of Science in relation to system thinking as part of the field of

learning earth sciences. Gudovitch and Orion (2001) studied systems thinking in high

school students and developed a system-oriented curriculum in the context of the carbon

cycle. Kali, Orion and Elon (2003) studied the effect of a knowledge integration activity

on junior high school students’ systems thinking, characterizing students’ conceptions of

the rock cycle as an example of systems-thinking. Ben-zvi-Assaraf and Orion (2004)

explored the development of system thinking skills at the junior high school level in the

context of the hydro (water) cycle.

Gudovitch (1997) examined students’ prior knowledge and perceptions

concerning global environmental problems in general and the role of people among

natural systems in particular. Importantly, the curriculum in this study provided a means

of stimulating students to explore the carbon cycle system. Gudovitch found that

students’ progress with systems-thinking consisted of four stages:

1. The first stage includes an acquaintance with the different Earth systems, and an

awareness of the material transformation between these systems.

2. The second stage includes an understanding of specific processes causing this

material transformation.

3. The third stage includes an understanding of the reciprocal relationships between

the systems.

4. The fourth stage includes a perception of the system as a whole.

Ault (1998) referred to drawing conclusions about past events as “retrodiction” (a

term drawn from Kitts, 19xxx) as opposed to prediction. Often retrodictions follow from

observations of phenomena in present time presumed to sample what has happened

through time. The challenge is “to hypothesize an arrangement by stages for what is

observed” (p. 196). Stages stand for periods of time; hence retrodiction and stage-

inference go hand in hand. Coral atolls, arc volcanoes, and river basins are often

explained as developing through stages over time. Examples of a volcano, coral atoll, or

Page 37: Revision of Learning Earth Sciences 7-30-04-Nir

37

river basin at any stage of development exist in the present. Hence, place substitutes for

time in order to make retrodictions; one example is another’s future.

Kali, Orion and Elon (2003) claimed that understanding the rock cycle is exactly

such a challenge and that such a challenge requires systems–thinking. They studied 7th

grade students who participated in learning a 40-hour unit. The main challenge was to

assist students in understanding the rock cycle as a system, rather than a set of facts about

the earth’s crust. The rock cycle is a system including the crust of the earth and is

characterized by a cyclic and dynamic nature. The rocks exposed on the surface of the

earth are only a small sample in time and space of constant material transformation

within the crust, driven by geological processes (e.g., weathering, sedimentation, burial,

metamorphism, melting, and crystallization of molten rocks, uplift and erosion). The rock

cycle can be viewed as a closed system, since hardly any material was added or removed

from this system in the time involved in students’ observations. Additionally, since the

size of the reservoirs of this system was almost constant over this time scale, it can also

be viewed as a system maintaining a dynamic equilibrium.

Kali, Orion and Eylon (2003), reported that while answering an open-ended

questionnaire, students expressed a systems-thinking continuum, ranging from a

completely static view of the system, to an understanding of the system’s cyclic nature.

They suggested placing dynamic thinking (which is a critical aspect of systems-thinking)

on a continuum, in which one side represents a static view, and the opposite side

represents a highly dynamic view of the system. On top of this continuum they

superimposed a dimension of interconnectedness. In the case of the rock cycle, they

based higher, more dynamic understanding upon making connections between parts of

the system. The degree of connectedness can therefore provide as means for determining

the degree of dynamics, and vice versa. This combined continuum served as a basis for

constructing a rock cycle systems-thinking continuum. At the low end of this continuum

they located students who presented a Product Isolation Model. Students thinking in

terms of this model express a lack of connectedness between parts of the system, indicate

poor dynamic thinking, and represent a completely static view of the rock cycle system.

At the opposite end of the continuum there were the students who thought dynamically

Page 38: Revision of Learning Earth Sciences 7-30-04-Nir

38

about material transformation within the rock cycle and therefore demonstrated a rich

understanding of the interconnectedness between parts of the system. With such a view

students were able to grasp the holistic idea that any material in a system can be a product

of any other material and apply this insight to novel situations. Such understanding was

considered as the highest level of systems-thinking in the context of the rock cycle. We

suggest that only at this level were students able to meaningfully understand the cyclic

nature of the system.

Between these two extremes were students that indicated different degrees of

systems-thinking. Students explanations classified under the category of Process-Product

Isolation Model reflect a view with dynamics limited to very small chunks of material

transformation within the rock cycle (between specific processes and their particular

products). Disconnected Internal-External explanations were placed higher on the

continuum, because they refer to the rock cycle as consisting of two different sub-systems

(i.e., the internal and the external systems). Though viewing parts of the rock cycle as

disconnected, this thinking did allow for larger chunks of material transformation within

these sub-systems. The most sophisticated alternative incorrect model concerning

material transformation within the rock cycle was the Model Lacking Burial & Melting

Processes, in which students viewed all the material transformation within the rock cycle,

except for one link (burial and melting of rocks). Therefore, they were placed higher than

Disconnected Internal-External explanations, just below the highest level of the systems-

thinking continuum.

It is important to note that students’ alternative incorrect models of the rock cycle

described above were not interpreted as misconceptions, or naive theories, about the

earth’s crust. Rather, placing these models on a continuum reflects the view that such

models can serve as basis for developing more sophisticated models, until the highest

level of understanding the cyclic nature of the system is reached. The progression within

this continuum is considered a result of adding connections between pieces of knowledge,

leading to higher levels of integrated knowledge.

This study also indicated that with appropriate teaching, students were able to

acquire systems-thinking in the context of the rock cycle. It was found that knowledge

Page 39: Revision of Learning Earth Sciences 7-30-04-Nir

39

integration activities led to a meaningful improvement in students’ views of the rock

cycle, towards the higher side of the systems-thinking continuum. Students became more

aware of the dynamic and cyclic nature of the rock cycle, and their ability to construct

sequences of processes representing material transformation in relatively large chunks

significantly improved. The success of the knowledge integration activity stresses the

importance of post-knowledge-acquisition activities, which engage students in a dual

process of differentiation of their knowledge and re-integration in a systems context.

Thus, the fact that, is encouraging. The findings also indicated that the systems-based

curricula design should include two stages:

1. A gradual knowledge building stage in which each of the system’s components is

studied in an inquiry process and gradually integrated into a holistic depiction of

the system.

2. A differentiation and re-integration concluding stage, which includes the dual

process discussed above.

Ben-zvi-Assaraf and Orion (2004) used a large battery of qualitative and

quantitative research tools in order to explore the development of system thinking skills

of junior high school students who studied the hydro cycle through learning with the

"Blue Planet" Program. The pre-test findings indicated that most of the students sampled

experienced substantial difficulties in all of aspects of systems-thinking. They even

struggled to identify basic system components. They entered the 8th grade holding an

incomplete and naive perception of the water cycle. At this stage they were only

acquainted with the atmospheric component of the cycle (i.e. evaporation, condensation,

and rainfall) and ignored the groundwater, biospheric, and environmental components.

Moreover, they lacked the dynamic and cyclic perceptions of the system and the ability to

create a meaningful relationship among the system components. The phenomenon of

disconnected “islands of knowledge,” which was reported by Kali, Orion and Eylon

(2003), regarding students’ abilities to connect a set of geological phenomena to a

coherent rock cycle, was found here as well. Most of the students were not able to link

the various components of the water cycle together into a coherent network. Some of

them demonstrated an ability to create a relationship between several components, but

Page 40: Revision of Learning Earth Sciences 7-30-04-Nir

40

even those students were not able at this stage to draw a complete network of

relationships.

In light of the initial knowledge and cognitive abilities of the students, the post-

test findings indicated that most of the students shifted from a fragmented perception of

the water cycle toward a more holistic view. About 70% of the students, who initially

presented only the atmospheric component of the hydro cycle, significantly increased

their acquaintance with the components and processes of the water cycle. For about half

of the students, this wide acquaintance with the systems’ components yielded an

improvement in their ability to identify relationships among components within the

system. Most of the students improved their dynamic perception of the system and about

one-third of them reached the higher level of cyclic perception. A meaningful

improvement was also noticed in relation to the students’ ability to identify hidden parts

of a system.

The triangulation of all the research tools indicates that only those students who

actively participated in the indoor and outdoor activities and submitted all the knowledge

integration assignments throughout the learning process reached the higher ability levels

of identifying a network of coherent relationships and hidden components of the system.

It is important to emphasize that not all the students who were actively involved within

the learning process reached those higher levels, but all students who presented such high

system thinking abilities did submit all the knowledge integration assignments. Some of

the students who were actively involved in the learning process, but did not develop the

dynamic cyclic perception of the system, could only identify relationships between one or

two components and could not perceive the overall cyclic nature of the system.

One factor that clearly influencing this ability is cognitive difference. Since

almost any population is cognitively heterogeneous one might expect a differential

cognitive development as was found by the current study. For some, the cognitive barrier

was the ability to perceive the dynamic relationship among the system’s components; for

others the ability to organize components within a network of relationships; and for

others the barrier was the ability to make generalizations. The finding that the common

factor for all those students who crossed all the cognitive barriers was their high

Page 41: Revision of Learning Earth Sciences 7-30-04-Nir

41

involvement within the learning process might indicate that system thinking is not only

influenced by the initial cognitive potential of a the students, but also by appropriate

learning strategies. In other words, system thinking is a cognitive ability that can be

developed through instructional learning. These findings together with the findings of

Kali, Orion and Eylon (2003), might also suggest that such learning should be based on

inquiry-based learning both indoors and outdoors and on knowledge integration activities.

The distribution of students’ achievements with regard to the different

components of systems-thinking were classified into four groups of skills. This

classification indicated that the development of system thinking in the context of the earth

systems consists of several sequential stages arranged in a hierarchical pyramid structure.

The cognitive skills developed in each stage serve as the basis for the development of the

next higher-order thinking skills.

The first group represented 70% of the students and includes “the ability to

identify the system’s components” and “the ability to identify the system’s processes.”

Both abilities can be classified together as “the system’s analysis skill.”

The second group includes two skills, both of which were presented by about 50%

of the sample: “the ability to identify relationships between separate components” and

“the ability to identify dynamic relationships between the system’s components.”

The third group includes three skills, which were presented by about 30-40% of

the sample population: “the ability to understand the cyclic nature of systems, the ability

to organize components and place them within a network of relationship,” and “the

ability to make generalizations.”

The fourth group was represented by a small number of the interviewed sample

population (10-30%) and includes the perception of the “hidden components of the

system” and the perception of the system within “the dimension of time,” namely the

ability to make a prediction (thinking forward) and the ability to look backwards at the

history of the system (retrospection).

The findings of a hierarchical notion and the interrelationships between dynamic

perception and cyclic perception which were found in the context of the hydro cycle are

in accordance with study of Kali, Orion, and Elon, (2003), which was conducted in the

Page 42: Revision of Learning Earth Sciences 7-30-04-Nir

42

context of the rock cycle, and with Gudovitch (1997) which was conducted in the context

of the carbon cycle. Thus, it suggested that these findings might be generalized to the

study of the earth systems.

In light of the findings and conclusions of the above studies, it is suggested that

the following aspects might contribute in improving students’ abilities to develop

systems-thinking.

1. Introducing the first steps of system thinking at the elementary school level,

namely skills such as the ability to identify the components of a system and

identifying relationships between two components. If the students enter junior

high school with adequate abilities of the lower levels of the system thinking

pyramid, more of them might be able to reach the higher levels of system thinking

during junior high school.

2. Focusing on inquiry-based learning.

3. Using the outdoor learning environment for the construction of a concrete model

of a natural system.

4. Using knowledge integration activities throughout all the stages of the learning

process.

2.2.5 Synthesis

Surprisingly (or not), there are connections among the several cognitive studies

mentioned above that were conducted separately. For example, Dodick and Orion

(1993a) have found an interrelationship between temporal thinking ability and spatial

thinking ability, while Orion, Ben-Chaim and Kali, (1997) as well as Riggs and Tretinjak

(2003) found that geological outdoor experiences might increase students’ spatial

thinking abilities. Ben-zvi-Assaraf and Orion (1994) found that systems-thinking in

relation to the earth is related to temporal thinking (retrospective thinking) and spatial

perception (the ability to perceive the hidden parts of a system). Here again, the outdoor

learning environment was found to be a very effective tool for developing a concrete,

realistic perception of nature that served as a cognitive bridge for the development of the

very abstract, high order thinking components such as temporal, spatial and system

thinking. Moreover, all of the above studies acknowledge the significance of alternative

Page 43: Revision of Learning Earth Sciences 7-30-04-Nir

43

frameworks that most students bring to earth science classes (no matter what age), and

therefore indicate the need to respond to preconceptions and misconceptions with

appropriate instruction, whether in the laboratory, the outdoors, the classroom, or when

working with computers.

Thus, although research in the area of learning earth sciences is quite limited, a

holistic framework has begun to emerge that links the cognitive elements that should be

stressed by teachers who work within an earth systems approach. Holism in this sense

refers to the interconnectedness of spatial reasoning and temporal thinking—not

surprisingly, exactly the slice of reality portrayed by a geologic map. Moreover, since this

research is still limited, it also suggests direction for research agendas for years to come.

2.3 The integration of learning environments within the earth sciences 2.3.1. A holistic approach to learning environments

An important characteristic of earth science education (and other sciences as well)

is the potential to conduct formal teaching in a variety of learning environments: the

classroom, the laboratory, the outdoors (field, museum, or industrial site), and the

computer. Orion (2001) suggests that in order to fulfill this potential and utilize these

environments to best effect, research related to earth science education should address the

following:

1. What are the educational advantages of each of these learning environments in

general and specifically in relation to earth science education?

2. What is the most appropriate context for utilizing each of these learning

environments, specifically in relation to earth science education as well as in

relation to science education in general?

3. What methods are needed to utilize these learning environments properly?

Responding to these questions should be done from a holistic perspective that

connects the several, different learning environments. The holistic perspective is not only

a way of viewing content, it also should inform the choice of a proper learning

environment and the appropriate learning tools.

2.3.2 The integration of the outdoor learning environment within the learning process

Page 44: Revision of Learning Earth Sciences 7-30-04-Nir

44

Review of the proceedings of the three International Geosciences Education

Organization international conferences on Geoscience education (IGEO, 1997; IGEO,

2000; IGEO, 2003) indicates a worldwide agreement on the central place of the outdoor

learning environment within earth science education.

Orion (1993a) suggested a holistic model that connects the outdoor and the indoor

learning environments (Figure 1). The guiding principle of this model is a gradual

progression from the concrete levels of the curriculum towards its more abstract

components. This model can be used for designing a whole curriculum, a course, or a

small set of learning activities (Orion, 1986, 1991). According to this model the outdoor

learning environment should be utilized early on in the concrete part of the learning

process, and should be mainly focused on concrete interaction between the students and

the environment. The outdoors offers tangible experiences relevant to learning abstract

concepts. The outdoor learning environment together with an indoor preparatory unit can

constitute an independent module, serving as a concrete bridge towards more abstract

learning levels. Thus, an outdoor learning activity should be planned as an integral part of

the curriculum rather than as an isolated activity.

Figure 1: The spiral model of integrating an outdoor learning activity within the indoors-learning process.

There is little doubt that initiating learning based upon a student’s own interest, or

at least upon a student’s understanding of why learning a specific topic matters

Field trip

Summary unit

Preparatory unitConcrete

Abstract

Page 45: Revision of Learning Earth Sciences 7-30-04-Nir

45

personally, might serve as a powerful tool for “meaningful learning” (in the sense defined

by Novak, 1977xxx, who has elaborated upon David Ausubel’s term, “meaningful verbal

learning” Ausubel 1966xxx). Meaningful learning is an ambiguous term used frequently

in this chapter. At this stage, we wish to constrain its ambiguity.

Meaningful learning may commence once the student grasps the meaning of a

new concept (i.e., both the sense of its use and its reference to a set of events or objects;

its signification of pattern or regularity in the occurrence of events; Gowin, 19xxx) and

hence can feel the significance of a question. The question may stem from personal

interest or arise from deliberate instruction. Meaningful learning happens when

connections are made through the use of concepts between interesting questions and their

appropriate answers (Gowin, 19xx) and when learners actively organize new meanings

into frameworks of ideas adequate to solving problems and interpreting events in novel as

well as familiar contexts (Novak, 1977xxx; Novak and Gown, 19xxx). Concepts, when

learned meaningfully, make reference to tangible events and derive meaning from clearly

articulated relationships to other concepts. To learn meaningfully means to organize

concepts into useful structures, to feel the significance of ideas, and to connect questions

with answers. The results of meaningful learning are hierarchical conceptual structures

with the power to subsume new knowledge either by “progressively differentiating”

existing categories of thought or by reconciling current ideas, hence integrating new

knowledge into a modified conceptual structure containing entirely new categories

thought, though not in isolation. Meaningful learning proceeds in terms of what is already

known and hence of personal relevance while fashioning “superordinate categories” to

integrate new knowledge in conflict with old (Novak, 1977xxx).

In Orion’s holistic model combining indoor and outdoor environments, the

learning process starts with a “meaning construction” session. In this session, students

converse, with guidance by the teacher, to discover what interests them about a particular

subject. Depending on the subject and the school’s location, this stage takes place in a

relevant outdoor environment or in a versatile indoor space. In the former environment

the function of the teacher is to mediate between the students and tangible phenomena. In

the indoor environment the teacher’s role is to motivate students’ interest by exposing

Page 46: Revision of Learning Earth Sciences 7-30-04-Nir

46

them to phenomena that are related to the subject through the using of pictures, video

films, computer software, Internet sites, and written texts.

It is impossible to go beyond the “meaning construction” stage, without briefly

describing the characteristics of the outdoors as a learning environment and to clearly

identify its advantages and disadvantages.

According to Orion (1993), the main role of an outdoor learning activity in the

learning process is to offer direct experience with concrete phenomena and materials.

Kempa and Orion (1996) suggested another aspect of the outdoor learning environment:

learning the methodology of field research, which plays a very important role in scientific

disciplines such as biology, ecology and geology. Thus, the goal of the outdoor learning

environment includes two main objectives: (a) learning basic concrete concepts through

direct interaction with the environment and (b) learning field investigation methodology.

The unique element of the outdoor experience is not in the concrete experiences

themselves (which could also be provided in the laboratory and classroom), but the type

of experiences: experiences that have emotional intensity. Students could view slides of a

dune and investigate quartz grains in the laboratory, but it is only when climbing the

steep front slope of a sand dune does a student experience of the structure a dune

perceptually and physically. Experience adds affect to the resources for constructing new

concepts. Experiential activities that facilitate the construction of abstract concepts

encourage long-term retention and meaningful learning in part, Orion speculates, because

affect enhances memory.

Historically, the interactionist view (attributed to Dewey, 1896) has stressed the

cognitive contribution of interaction with the physical environment. Dewey’s work, for

example, acknowledges the value of implicit associations acquired through extensive

experience with a phenomenon of interest. These implicit associations—familiarity with

properties and possibilities—are, for Dewey, necessary precursors to explicit

understandings. Explicit understandings exist in bodies of formal knowledge (subjects).

Subjects are repositories of resources for solving different kinds of problems. Children

and adolescents, however, typically have little experience in need of these resources.

Dewey argues that education should extend the experience of the learner into domains

Page 47: Revision of Learning Earth Sciences 7-30-04-Nir

47

where these resources have value. The goal is for the learner to see knowledge as

purposeful.

Dewey feared the sterility of content presented to students as specialized

knowledge bereft of the reasons for its construction:

Those things which are most significant to the scientific man and most valuable in

the logic of actual inquiry and classification drop out. The really thought-

provoking character is obscured, and the organizing function disappears. (Dewey

[1902] 1990, 205xxx)

Dewey valued the logic of action in the context of actual inquiry (e.g., learning

field investigation methodology in the outdoors classroom). He thought insight into this

logic would inform teaching in a positive manner. In addition, much of his work

emphasized the need to respect the nature of the child and the importance of childhood

experience. He was Rousseau’s obvious successor in this regard. His goal was to

reconcile two traditions: subject-centered and child-centered instruction. He noted, quite

rightly, that schooling expected children to learn ideas without understanding the

purposes they served. The outdoor environment appears essential to revealing these

purposes when learning earth sciences as well as to unveiling the logic of inquiry in field

settings.

In addition to Dewey’s interactionist viewpoint, Gibson’s (1966) theory of direct

perception supports the validity of Orion’s model of outdoor-indoor activity. Gibson

argues that perception should be understood as a process of obtaining information from

activity, rather than as a (passive) process of constructing representations of the situation

and operating on those representations (Greeno, Collins and Resnick, 1996).

One point is most crucial to understand: the outdoor learning environment addresses

phenomena and processes that cannot be cultivated indoors. The outdoors is a very

complicated learning environment and includes a large number of stimuli that can easily

distract students from meaningful learning. Thus, the first task of teachers and curriculum

developers is to identify and classify phenomena, processes, skills, and concepts which

can only be learned in a concrete fashion outdoors, and those that can be learned in a

concrete fashion indoors. In addition, it is important to identify those abstract concepts to

Page 48: Revision of Learning Earth Sciences 7-30-04-Nir

48

which the outdoors contributes little to student understanding. In such cases, more

sophisticated indoor tools (such as pictures, films, slides and computer software) must be

substituted.

Consider a location where students find that an outcrop reveals an anticline and

begin to infer geological processes that might have produced this structure. Are they

ready to approach this task? Or is the challenge too novel? Many of the concepts useful to

drawing conclusions about the anticlinal structure sedimentation, superposition, and

initial horizontality can be better explained through lab observations and simulations.

Following the understanding of these concepts students who arrive to this specific

outcrop can conclude that the layers are not located in their original setting. Then,

through a field observation they might decipher the anticline structure. From this point, a

better understanding of the three dimensional nature of a folded structure as well as the

folding mechanism can be effectively achieved through the use of computer software and

hand held models (Kali and Orion, 1997).

Following the “meaning construction” stage, conducted either outdoors or

indoors, the first phase of a specific learning spiral starts in the indoor learning

environment. The length of time of this phase is varied; it is entirely dependent on the

specific learning sequence. The main aim of this phase is to prepare the students for their

outdoor learning activities. This preparation deals with reducing what is termed by Orion

and Hofstein (1994) as the “novelty space” of an outdoor setting (Fig. 2). The novelty

space consists of three factors: cognitive, geographical and psychological. The cognitive

novelty depends on the concepts and skills that students are asked to deal with throughout

the outdoor learning experience. The geographical novelty reflects the acquaintance of

the students with the outdoor physical area. The psychological novelty is the gap between

the students’ expectations and the reality that they face during the outdoor learning event.

The novelty space concept has a very clear implication for planning and

conducting outdoor learning experiences. It defines the scope of preparation required for

an educational field trip. Preparation that considers the three novelty factors reduces the

novelty space to a minimum, thus facilitating meaningful learning during the field trip.

Working with the materials that the students will meet in the field and conducting

Page 49: Revision of Learning Earth Sciences 7-30-04-Nir

49

simulations of geological processes through laboratory experiments directly reduces

cognitive novelty. To reduce the geographic and psychological novelty of the outdoor

learning experience teachers may turn first to slides, films, maps, and secondly to detailed

information about the event. Students should know the purpose of outdoor learning, the

learning method, the number of learning stations, the length of time, the expected weather

conditions, the expected difficulties along the route, etc. Safety briefing is a must as well.

Figure 2: The three dimensions, which identify the novelty space of an outdoor learning activity

The next phase in this cycle is the outdoor learning activity. The curriculum

materials for the outdoor learning experience should lead students to interact directly with

the phenomenon and only secondarily, if at all, with the teacher.. The teacher’s role is to

act as a mediator between the students and the concrete phenomena. Some of the

students’ questions can be answered on the spot, but only those, which might be

answered according to the evidence uncovered in the specific outdoor site. Otherwise

time and resources, including the students’ attention, is wasted on activities that might be

done elsewhere. Lectures, discussions and long summaries should be postponed until the

next phase, which is better conducted in an indoor environment.

Marques, Paria and Kempa (2003) explored Orion’s model within the Portuguese

earth sciences curriculum. Their study supported the importance of preparation for the

outdoor learning experience. They also found a positive influence of this learning

environment on students’ learning. However, their study also highlighted the difficulties

Page 50: Revision of Learning Earth Sciences 7-30-04-Nir

50

teachers faced in adapting to the novel, outdoor learning environment. (The last section of

this chapter returns to this issue in detail.)

Geo3D software (Kali & Orion, 1996) nicely illustrates an example of the

indoor—outdoor cycle. The design of this software fosters the development of spatial

visualization skill. Most geological outcrops hide elements of the three dimensional

configuration of geological structure. Even having observed a structure such as an

anticline in the field, most students have difficulty perceiving its three dimensional form.

Thus, the outdoors is not as suitable a learning environment as is a computer simulation

for the development of spatial visualization (Kali and Orion, 1996). At the same time,

many of the software’s tasks are based on the same concrete phenomena that students

have observed and identified during their geological field trips.

Gudovich and Orion (2003) researched how to integrate the computer and the

laboratory learning environments within the framework of a distance learning course. The

study included the development of a website with detailed visual and textual instructions

for conducting hands-on lab activities and an identical kit of all the equipment and

materials needed for conducting a specific experiment. The web-site also included

questions that students have to answer following their activity and send by email to the

teacher. This model was tested in the following two settings:

1. In the school's lab where students sat in small groups (2-3) in front of a computer

with the activity kit on the table and conducted the lab activity independently

following the visual and the textual instructions of the distant learning website.

The teacher remained in the lab in order to address any needs of the students.

2. In an out of school setting where a student sat alone without a teacher and

conducted the lab activity by himself/herself.

Following their initial study of distance learning, Gudovich and Orion

administered a battery of qualitative and quantitative instruments in order to find out

more about the role of the teacher during lab activities. Findings indicated that it is

possible to decrease the need for teachers during lab procedures and to decrease the

chance of failure. Analyzing students’ difficulties during the activities indicates that those

difficulties are not unique to distant learning setting.

Page 51: Revision of Learning Earth Sciences 7-30-04-Nir

51

2.3.3 Integrating the lab learning environment within the learning sequence

Although there are many laboratory-based earth science units for various age

levels all over the world, little has been published concerning the role of the laboratory

learning environment within the earth science education.

A review of many such lab-based units indicates that the main role of the

laboratory is to demonstrate or simulate the earth's processes. However, little has been

published concerning the influence of simulations on the development of misconceptions

among school students. Another unexploited area of the earth science laboratory

environment is its great potential for contributing to the development of inquiry scientific

skills and thinking.

Inquiry in the geosciences has a unique characteristic: its “experiments” in the

grandest sense have already been conducted by nature. They are unfundable and

unreplicable. No one can send glacial ice across a continent or carve a Grand Canyon.

Consequently, many geological inquiries are of a retrospective type—trying to unravel

what happened in the past, using “fingerprints” left on the earth. Frodeman, (1995)

describes geology as an interpretive and historical science, which “embodies distinctive

methodology within the sciences.” He further argues that “the geologist picks up on the

clues of past events and processes in a way analogous to how the physician interprets the

signs of illness or the detective builds a circumstantial case against a defendant” (p. 963).

Edelson, Gordin and Pea (1999) describe the geosciences as being “observational

sciences” that emphasize comparisons and contrasts among features of the earth in time

and place. Inference based upon comparison and contrast, especially when considered

across different scales in time and place, differs from inference based upon the results of

experimentation (Ault, 1998). Both approaches are empirical, quantitative, and subject to

scrutiny using rules of logic. They offer different milieus for illustrating the meaning of

some of the most basic constructs of scientific thinking: for example, observations,

conclusions, and hypotheses.

A traditional method for categorizing inquiry curricula is to analyze the degree of

structure or openness of the activities they include (Schwab, 1962; Herron, 1971; German

et al, 1996). Using such methods, inquiry-based curricula can be placed anywhere on a

Page 52: Revision of Learning Earth Sciences 7-30-04-Nir

52

continuum extending from completely structured curricula on one side to completely

open curricula on the other.

Those who advocate inquiry in the science curricula for all accept that the

educational system ought to enable students bring students to a stage where they will be

able to design, conduct, and analyze their own investigations, then communicate their

findings. However, the appropriate stages for engaging students in open inquiry is not

clear, nor are the means for bringing students to a stage in which they will be able

autonomously to design and conduct their own experiments. While some researchers

suggest designing a variety of activities to suit a diversity of cognitive developmental

stages in a classroom (e.g. Germann, 1989), others suggest preparing students for open

inquiry by engaging them with well structured investigations (e.g. Edelson et al., 1999).

One of the rarely asked questions regarding inquiry learning concerns the

cognitive prerequisites necessary for using open inquiry methods. Elshout and Veenman,

(1992) claim that “In unguided-discovery learning, one expects high metacognitive skill

and intellectual ability to be essential requisites to keep the learning process going”

(p.135). It is therefore reasonable to claim that students should understand the meaning of

some of the most basic concepts used in scientific methodologies before they can begin

an independent inquiry process. Such understanding provides the means for making

hypotheses, designing experiments, collecting and analyzing data, and reporting their

findings. Unfortunately, evidence exists indicating that students in junior and senior high

schools have severe difficulties in understanding the essence of the scientific method.

They have, in effect, failed to learn scientific method as a content with its own concepts

and principles. Zohar (1998) reported that junior high school students had difficulties in

understanding the difference between their experimental results and their conclusions.

Solomon, Duveen & Hall (1994) reported that high school students had difficulties in

distinguishing between descriptions and causal explanations. Tamir (1989) claims that

“students do not understand the concepts that underlie the processes of scientific

investigations. These concepts (e.g., hypothesis, control) are not easy to understand…”

(p. 61).

Page 53: Revision of Learning Earth Sciences 7-30-04-Nir

53

Learning earth sciences has a role to play in remedying this situation. Kali and

Orion (2004) suggest that the earth sciences education has the potential to provide

students, at beginning stages of their science education, with basic inquiry skills that are

required for further open-ended inquiry endeavors. They developed a 34-hour lab-based

curriculum unit for junior high school students, focusing on geological processes that

transform the materials within the crust of the earth—“The Rock Cycle”—and organized

this curriculum into nine structured inquiry modules. To foster students' awareness of the

different inquiry routes embedded in the inquiry modules, each of the modules was

followed by a MIR (Metacognitive Inquiry Reconstruction) assignment. In these

activities linguistic terms were used as organizing schemes. Students examined their

investigation with “scientific inquiry spectacles,” and categorized different stages of the

inquiry with terms such as observations, hypotheses, and conclusions.

MIR presents assignments in three alternative variations, from which the students

choose. Each variation is a different combination of solutions to students’ needs and

difficulties. The first variation starts with a short verbal summary of the inquiry path,

with missing words for the student to complete. It continues using the linguistic schemes

(terms like “observation”, “hypothesis” and “conclusion”), for organizing the inquiry

process. This variation was designed for students who feel comfortable with verbal and

structured environments, and who have difficulties in reconstructing the inquiry path

without assistance.

The second variation includes a list of verbal expressions that constitute the entire

inquiry path in the related module. Students are asked to name the appropriate inquiry

construct for each expression. This variation is very similar, in the degree of structure, to

the first variation. It was designed to support students who are uncomfortable with filling

in missing words in a text, as required in the first variation.

The second MIR variation enables students to focus on the inquiry constructs,

without representing their understanding of the inquiry path. In the third variation,

students list the stages of the inquiry path in the module, using the inquiry constructs, and

represent this path in any way they choose, using verbal or graphic means.

Page 54: Revision of Learning Earth Sciences 7-30-04-Nir

54

Kali and Orion (2004) tested the influence of learning an inquiry-based “Rock

Cycle” curriculum and its accompanying MIR activities on student ability to distinguish

between observations, hypotheses, and conclusions on a sample of 582 students in 7th and

8th grade from 21 classes sharing 14 teachers at 8 junior high schools in Israel. The

schools represented urban, suburban, and rural societies. The study used a large battery of

qualitative and quantitative research tools in a pre-test/post-test structure.

The pre-test outcomes indicated that the 7th and 8th grade students included in this

study had considerable difficulties in understanding concepts underlying the scientific

method. The large and significant pre-post differences found in many of the classes

indicated the high potential for an inquiry-based “Rock Cycle” program to develop and

distinguish among three basic elements of scientific thinking (observations, hypotheses,

conclusions).

The large improvement in students’ scientific thinking skills, found in many of the

classes, might have been a result of students’ engagement with the unique inquiry

methods of geoscience. Students focused their tangible observations on materials of the

earth. They drew conclusions from “experiments” that were conducted by nature in the

past and did not design their own investigations.

However, Kali and Orion also found no improvement among classes taught by

teachers who did not properly adopt the inquiry-based teaching strategy. These teachers

taught the “Rock Cycle” unit in the traditional manner. Thus, despite the great importance

of appropriate curriculum materials, they are not sufficient in themselves for inducing

cognitive development amongst students. Sometimes teachers are the limiting factor in

students’ ability to exploit the potential of “The Rock Cycle” in developing scientific

thinking skills.

2.3.4 The computer learning environment.

The role of the computer learning environment for learning earth sciences is

growing as the availability of computers as learning tools increases in many countries.

The computer is mainly used within the earth science education for the following learning

purposes:

Page 55: Revision of Learning Earth Sciences 7-30-04-Nir

55

1. Demonstration through animations and simulations of earth processes and

phenomena that cannot be demonstrated in the lab or in the field (Kali and Orion,

1996; Reynolds et al., 2002; Kali, 2003).

2. Data collection and knowledge integration (Kali, Orion and Elon, 2003).

3. Knowledge presentation through multimedia software (Chang, 2004; Orion,

Dubowski and Dodick, 2000).

4. Web-based learning that provides a variety of learning resources such as on-line

information about phenomena and processes around the planet earth and other

planets as well; on-line data bases; communication with peers and experts; and the

possibility of distant learning (Slattery, Mayer and Klemm, 2003; Gudovich and

Orion 2004).

Although there has been some research concerning the computer learning

environment in the context of the earth sciences (as mentioned above), it is still in its

embryonic stage and there is a great need to explore the unique contribution of this

environment to learning earth sciences.

2.4. Research and the development of curriculum materials The main goal of earth science education is to improve the way students learn

about and understand our planet. To achieve this goal, the end product of science

education research should be:

1. Development of learning materials and learning strategies for a wide range of

students and teachers.

2. Development of appropriate teaching materials and strategies, as well as preparing

teachers to fully implement such strategies.

Following the holistic earth systems approach for teaching earth science and the

holistic approach of integrating the earth sciences’ learning environments together with

emphasis on a concrete-abstract continuum, it is only natural to present a holistic model

for the curriculum materials development for learning earth sciences. In this section, we

report in detail about a curriculum for teaching the water cycle from earth systems as well

as environmental issues perspectives. The curriculum, “The Blue Planet,” emerged from

a “design research” effort.

Page 56: Revision of Learning Earth Sciences 7-30-04-Nir

56

Edelson, Gordin and Pea (2004) advocate for “design research” as a powerful

model for the development of effective learning tools. They used this model to develop

inquiry-based software for the study of climatology through visualization. In design

research, the study of learning takes place in the context of designing and revising

curriculum materials based upon careful study of student response to these materials.

Orion’s (2002) helical model of research, curriculum development, and

implementation is similar. In this model, each curriculum development effort starts with a

pre-development study to identify misconceptions, preconceptions and learning

difficulties associated with the specific subject. The findings from this stage serve as a

basis for the first curriculum development phase. An implementation phase follows

curriculum development. The implementation phase involves in-service training for a

small number of teachers who will teach the curriculum to their classes.

An evaluation study follows the implementation stage. The results of the

evaluation inform the second iteration of curriculum development. In turn, this phase is

followed by a wider implementation cycle. Evaluation happens again, this time on the

wider scale, leading to a third curriculum development stage, and so on. This model helps

to adapt curriculum materials for specific ages and varieties of students found in different

classrooms. It responds to the difficulties both students and teachers are having with the

curriculum. This model of research, curriculum development, implementation, and

evaluation should continue so long as the curriculum is in place.

2.4.1 Pre-development of “The Blue Planet” curriculum

Based upon Orion’s helical model, research preceded and followed the

development for 8th grade students of an earth systems unit on the hydrosphere, “The

Blue Planet.” In order to examine students’ prior knowledge and understanding in

relation to the water cycle, a “zoom-in” analysis was conducted. Quantitative research

tools were used with a large sample in order to obtain a general picture of students’

knowledge and perceptions. Later, qualitative research tools were used with a smaller,

randomly selected sample in order to gain insight into “misconceptions” or “alternative

concepts” and to validate the quantitative tools.

The pre-development study included the following two objectives:

Page 57: Revision of Learning Earth Sciences 7-30-04-Nir

57

1. Identify prior knowledge of the water cycle held by students entering junior high

school.

2. Explore their perceptions of the cyclic and the systemic nature of the water cycle.

Review of the literature concerning the predevelopment phase revealed that in

spite of the crucial importance of water from the environmental perspective, most of the

studies that have been conducted in this area have concentrated on students’ perceptions

of the physical aspects of the water cycle, namely, changes in the water state (Bar, 1989;

Bar & Travis, 1991). An ERIC search in 2002 revealed only a few published studies that

focused on children’s perceptions of the water cycle in the environmental context of the

earth. Agelidou, Balafoutas, and Gialamas (2001) reported that students do not perceive

how human activities are related to water problems and their consequences. Specifically,

they do not recognize the principal factors responsible for these problems.

Fetherstonhaugh and Bezzi (1992) reported that after 11 years of schooling, students

could only present simplistic and naïve conceptions of the water cycle. Moreover, the

students showed a poor and inadequate scientific understanding of groundwater as a part

of the water cycle.

Brody (1994) conducted a meta-analysis study of about 30 articles published

between 1983 and 1992 that dealt with difficulties of middle and high-school students in

understanding different subjects connected with water. Only a few of those articles dealt

with the environmental aspects of water, whereas at least 80% of them focused on the

following three areas of difficulty:

1. Understanding the chemical and physical processes such as condensation,

evaporation, and the molecular structure of water.

2. Understanding the significance of water for processes that take place in living

organisms.

3. Understanding interdisciplinary subjects such as water resources, and the

social and scientific linkages of these topics.

Taiwo, Ray, Motswiri, and Masene (1999) confirmed that students’ perceptions of

the water cycle were influenced by their cultural beliefs and to a large extent by their

pseudoscientific knowledge about cloud formation and rainfall. Barker (1998) reported

Page 58: Revision of Learning Earth Sciences 7-30-04-Nir

58

that in spite of the fact that about 90% of the water absorbed by the roots is lost by

evaporation, mainly through the leaves, 50% of the students in his study claimed that

plants retain all the water that they absorb.

Since the literature provided only a limited basis for students’ alternative

frameworks concerning the water cycle, a pre-development study was designed and

implemented among a population of about 800 junior high school students. Five types of

research tools were used in the study: Likert type questionnaires, open questions,

drawings, word associations and interviews. The following is a brief description of the

different tools.

Groundwater Dynamic Nature Questionnaire (GDN): This Likert-type

questionnaire identifies students’ prior knowledge and understanding of the dynamic

nature of the groundwater system and its environmental relationship with humans.

Cyclic Thinking Questionnaire (CT): This questionnaire contains two parts. The

first part includes a Likert-type questionnaire in which students mark their level of

agreement with seven statements concerning the cyclic nature of the hydrosphere and the

conservation of matter within earth systems. The second part includes two open

questions: In the first question students define the concept of the cyclic process in nature;

in the second question they give an example according to their definition.

Global Magnitude Questionnaire (GM): This questionnaire contains two parts.

The first part includes a Likert-type questionnaire in which students indicate their level of

agreement with five statements about the scale of each stage of the water cycle. In the

second part, students grade on a scale of 1 to 10 the relative global quantity of water that

exists within each of the following components of the water cycle: oceans, glaciers,

rocks, soil, groundwater, lakes, precipitation, tap water, sewage water, and humans. In

this scale, 1 was the major reservoir of water on earth and 10 was the smallest reservoir.

Assessing Students’ Knowledge (ASK): In this Likert-type questionnaire, students

indicate their level of agreement with three statements concerning the physical and

chemical processes within the water cycle.

Drawing Analyses (DA): There is evidence that young children can communicate

scientific ideas through their drawings by drawing what they know about a particular

Page 59: Revision of Learning Earth Sciences 7-30-04-Nir

59

object rather than what they see. In this task, the students draw the water cycle, using in

their drawings as many items as possible from a list of stages and processes of the water

cycle provided to them. The students receive assurances that no one expected a highly

artistic drawing.

In order to increase the reliability and consistency of the analysis of the drawings,

the authors each individually coded the drawings of 20 students. After comparing and

discussing the two separate analyses, a standardized coding system was developed.

Additionally, follow-up interviews were conducted with 50 students, in which students

were asked to elaborate on their drawings. The drawings were analyzed according to the

following criteria:

1. Presence of earth systems. 2. Depiction of processes. 3. Examples of human consumption or pollution. 4. Cyclic conception of the water cycle as a series of links among water cycle

components. Word Association (WA): Word Association directly probes for associations

among a set of concepts. Students write down all the water cycle concepts familiar to

them, then later classify them in relation to a unifying concept such as processes in the

water cycle, location, geosphere, hydrosphere, biosphere and atmosphere, human use of

water, and environmental and chemical aspects.

Interviews: Interviews with 40 students validate the analysis of their

questionnaires and provide insights into students’ perceptions of the water cycle. During

the interviews, students read their answers and indicate whether they still agree with their

drawings or responses to the questionnaire. They then elaborate on their answers.

Transcription and qualitative analysis of the questionnaires from the

predevelopment study indicated that most of the students demonstrated an incomplete

picture of the water cycle and held many misconceptions about it. Children that drew the

water cycle usually represented the upper part of the water cycle (evaporation,

condensation, and rainfall) and ignored the ground water system. More than 50% of the

students could not identify components of the ground water system even when they were

familiar with the associated terminology. In their mind, underground water was a static,

Page 60: Revision of Learning Earth Sciences 7-30-04-Nir

60

sub-surface lake. Furthermore, they imagined that water chemistry was constant

throughout the entire water cycle (no purification by evaporation). Presumably,

environmental insight regarding water pollution and water conservation requires

connecting the stages of the water cycle to the processes that modify water quality and

abundance. The water cycle alterative frameworks held by more than 50% of the students

do not bode well for learning environmental insights.

Cyclic thinking correlated significantly with drawing the water cycle to include

its groundwater component. A student who drew the underground water system held the

following concept about the cyclic nature of the water cycle: “I absolutely disagree.

There is no starting point and no end point in the water cycle. It is a continuous process.”

Analysis of the pre-development study findings might suggest that the ability of

students to perceive the hydrosphere as a coherent system depends on both scientific

knowledge and cognitive abilities.

The knowledge component has two elements:

1. Factual-based knowledge that includes acquaintance with the components of the

water cycle and awareness of its processes.

2. Process-based knowledge, namely a deep understanding of the various processes

that transform matter within the water cycle.

The cognitive component also has two elements:

1. Cyclic thinking, namely the understanding that the water cycle is a system which

has no starting or end points; just the same matter, but in different forms,

transformed over and over again within the system.

2. Systemic thinking, which is the ability to perceive the water cycle in the context

of its interrelationship with the other Earth systems.

2.4.2 Development and evaluation of the “The Blue Planet” curriculum

The findings of the pre-development study served as a basis for the development

of an interdisciplinary program named “The Blue Planet.” This program focused on the

water cycle as an example of the relationships seen amongst the various earth systems.

The evaluation effort examined the effect of “The Blue Planet” program on the earth

science learning of 700 junior high school students. The evaluation focused on:

Page 61: Revision of Learning Earth Sciences 7-30-04-Nir

61

1. Exploring students’ conceptions and attitudes about people’s relationships with

the earth system.

2. Identifying the alternative frameworks students possess with regard to the

components of the water cycle.

3. Identifying changes in knowledge and cognitive skill among students.

Two additional research tools were used in this stage:

Concept maps: The students were asked to create concept maps at the beginning

and end of the learning process. Comparison of the number and type of items between the

concept maps served as a measure of changes in students’ knowledge and understanding

of processes. The number of connections within the concept map served as an indication

of students’ understanding of the relationship between the components of the water cycle.

Observations: In order to track the learning event itself regular observations were

conducted in the classes. The observer used a structured observation report that directed

her to document the type of activities of both students and the teacher.

The following are the main findings of the evaluation study:

1. The observations indicated that for the most part, the teachers concentrated

primarily on scientific principles and only very little on the cognitive aspects of

the connections between the water cycle and the other earth systems, or between

the water cycle and environmental case studies. In addition, most teachers tended

to ignore the constructivist activities developed in light of the findings of the pre-

development study. These were activities intended to correct students’

misconceptions and to develop a broader, more coherent conception of the water

cycle within an earth systems context.

2. A significant improvement was found in student’s level of knowledge, namely

acquaintance with the components of the water cycle.

3. A significant improvement was found in relation to students’ understanding of the

evaporation process. However, in relation to all the other processes only a minor

improvement was found.

4. The analysis of the cyclic and systemic thinking questionnaires showed some

improvements in students’ understanding of the different types of

Page 62: Revision of Learning Earth Sciences 7-30-04-Nir

62

interrelationship among the earth systems. However, even after completing The

Blue Planet program, poor understanding of the systemic nature of the water

cycle dominated student thinking. Most of the students showed a fragmented

conception of the water cycle and made no connections between the atmospheric

stages of the water cycle and the geospheric (underground) stages of the water

cycle.

These findings indicate that improvement in knowledge is not enough for the

development of environmental insight. For this purpose students should develop their

cognitive abilities of cyclic and systemic thinking, through learning activities that were

directly developed for this purpose. Although such activities were provided, teachers

tended to ignore them. They need to understand that simply gaining knowledge about the

components of the water cycle does not contribute to progress in the development of

environmental insight.

3. The teaching aspect: the difference between professional developmentand professional change

We cannot complete the holistic view of the ability to engage students in

meaningful learning of the earth sciences without paying close attention to teaching. As

determined in the evaluation of “The Blue Planet” program, teachers may limit the

introduction of new content and new learning strategies within schools. Teaching

strategies that are needed in order to achieve meaningful learning about earth systems and

environmental insights are quite different from the traditional science teaching (Table 1).

Table 1: A comparison between traditional science teaching and proper earth systems teaching. Traditional science teaching Proper earth systems teaching The main purpose is to prepare the future scientists of a society

The main purpose is to prepare the future citizens of a society

Disciplinary-centered teaching Multidisciplinary teaching

Teacher-centered teaching Child-centered teaching

Content-based teaching Integration of skills within contents

The teacher as the source of knowledge/information

The teacher as a facilitator of learning and mediator of knowledge

“Chalk and talk” based teaching Inquiry based teaching

School-based learning Multiple learning environments: Classroom, lab, outdoors and computer.

Curriculum that is derived from the scientific world

Curriculum that is derived from the real world of student experience

Page 63: Revision of Learning Earth Sciences 7-30-04-Nir

63

Traditional assessment Alternative assessment In order for teachers to move from the left column of Table 1 to the right one—

from traditional to earth systems teaching—they must change their goals for student

learning, the contents of their curricula, and their approaches to instruction. Clearly, this

shift constitutes a major change in philosophy, from reductionism and disciplinary-driven

schooling towards holism and attention to educating students for lives of social

responsibility within democratic societies. The shift presented in table 1 is valid for any

genuine “Science for All” teaching, however teaching about earth systems demands

something more: that teachers actually teach earth science subjects, an area in which

many science teachers in many countries have little or no scientific background (King,

2003; Orion 2003b). Furthermore, students learn these subjects best—and often can only

learn field methodologies of investigation—when teachers make use of the outdoor

learning environment. Most traditional science teaching ignores this environment.

The use of the term “professional development” is misleading and contributes to

the difficulties of making a genuine change in teaching style and focus. Professional

development is far too restrictive a concept. The task to be accomplished exceeds what

we might expect of professional development. It requires participation and commitment

on many levels, from community and school to business and academia.

Orion (2003b) has reported on the outcome of a long-term (10 years) study within

the “storm’s eye” of the new Israeli “Science for All” curricula for junior high and high

school. This intensive work included participating in the committees that designed the

new “Science for All” curricula for junior high and high school; taking a central role in a

team that has developed learning materials for these two programs; and leading and

taking a practical role in hundreds of in-service training hours in each of the 10 years

both in in-service training centers and in the teachers’ schools and classes.

Study of this decadal process has produced four Ph.D. dissertations (Kali & Orion

2003; Dodick & Orion 2003b; Ben-zvi-Assraf & Orion 2004; Kapulnick, Orion & Gniel

2004) and one Master’s thesis (Midyan, 2003). This action research has included

qualitative and quantitative data collected with questionnaires, interviews and classroom

observations. All together, these different studies examined the practice of science

Page 64: Revision of Learning Earth Sciences 7-30-04-Nir

64

teaching and learning for about 1000 science teachers and their students. Most of the

studies were conducted at the junior high school level, but also included teachers and

students from the elementary and the high school levels. In addition to observing teachers

and evaluating student learning, these studies addressed systemic reform from the points

of view of principals, superintendents, curriculum developers, academic scientists, the

ministry of education, as well as in-service training and pre-service education programs

for teachers.

Each of the different and separate studies indicated that despite their participation

in long term in-service training programs, the vast majority of the teachers did not

undergo genuine professional development. Professional inertia was the rule. The studies

indicated a clear gap between the teachers’ positive declaration about their development

as were expressed through questionnaires and interviews and their actual practice in

classrooms. This gap was especially clear in regards to their implementations of new

teaching methods and new subject matter.

In addition to teachers’ reluctance to implement new teaching methods and

incorporate new scientific topics, the interviews with the teachers revealed four additional

factors which prevented them from genuinely implementing a reform:

1. A general apprehension about change.

2. The feeling that the training institutes outside of school regional centers did not

provide them with practical tools needed to overcome their apprehension about

unknown areas.

3. The lack of support from the school management, which does not provide them

the needed resources for adapting new teaching strategies such as laboratory

equipment, computers, a reasonable number of students for working in a

laboratory, resources for using the outdoor learning environment, and a

reasonable amount of teaching hours.

4. A double-standard from the Ministry of Education in general and more

specifically from their science education inspectors. The teachers claimed that

they were confused, since on one hand the Ministry of Education initiated the

reform and their inspectors encouraged them to participate in the regional centers’

Page 65: Revision of Learning Earth Sciences 7-30-04-Nir

65

INST (xxx what does INST stand for?) program, but at the same time the Ministry

of Education did not provide them with needed resources. Furthermore, the

Ministry called upon the inspectors to implement a national testing regime—in

many respects institutionalizing the antithesis of the new “Science for All”

approach.

The above findings suggest that the movement of teachers towards an earth

systems approach with the spirit of the “Science for All” paradigm consists of more than

professional development. For these teachers it is a paradigm shift and many (actually

most) cannot or do not really want to undergo such a huge change.

In light of these considerations, Orion (2003b) suggested that an effective model

for professional change should include the following components:

1. In the first stage the teachers have to experience the new methods and contents as

learners. Positive experiences as learners will help them to be both convinced of

the effectiveness of the new paradigm and later to deal with their students’

learning difficulties from the perspective of the difficulties that they experienced

as learners.

2. The school’s management should be an integral part of the INST and take on the

commitment of facilitating the implementation of the new reform.

3. The first teaching experiences with the new methods and contents should be done

with close support from the INST experts.

4. The INST leaders should be equipped with psychological knowledge and skills to

deal with reservations and oppositions which result from a fear of change.

The world is complicated and diverse and the Israeli example is that of just one

nation. Yet the conflict between reform efforts and testing priorities is worrisome and

certainly experienced elsewhere. Most importantly, the Israeli case illustrates the need for

research in science education to address many contexts, from integrating curriculum to

changing teaching paradigms.

4. Summary

Page 66: Revision of Learning Earth Sciences 7-30-04-Nir

66

The first decade of the 21st century finds earth science education in a more central

place in science curricula than a decade before. The progress of earth science in schools

all over the world is closely related to its central role in the development of

environmental insight among future citizens. However, the ability of earth science to

establish itself as a sustainable course of study in schools is highly dependent on the

ability of science teachers to overcome many barriers, including their own lack of

background and the persistent low stature of the field. This low stature is a function of the

failure to understand “what’s so special about learning earth sciences.”

Learning earth sciences offers the distinct potential of seeing through the

landscape and through time. Its many subjects unite to conceive of the world as dynamic,

interacting systems, themselves composed of stabilizing cycles. These systems operate on

many scales in time and place, some so vast as to challenge the limits of imagination. The

earth sciences represent phenomena of interest in visual forms: contour maps, block

diagrams, and computer models of virtual worlds, both of the interior earth and its

changing climate. These representations place distinctive demands on the cognitive

capacities of learners. Making sense of earth’s processes and patterns, structures and

changes, systems and cycles, depends upon visualization and spatial reasoning as well as

recognizing bias in the human-scale perception of events.

Understanding how the earth works requires retrospection and retrodiction—

making inferences about the past. By interpreting the present as the outcome of natural

experiments on vast scales and sleuthing out its causal history, earth sciences set the stage

for making extrapolations about possible futures. These extrapolations inform our actions

with information about risks, from seismic to atmospheric. On local, regional, and global

scales human interact with earth’s natural systems, becoming agents of geologic,

climatic, and evolutionary change. This power carries heavy responsibility; learning earth

sciences offers lessons students need in order to develop their capacity to exercise this

responsibility.

This chapter presents a holistic view of earth sciences education and a holistic

model to achieve a meaningful learning of the earth sciences. This model combines an

educational vision (development of environmental insight through adopting the earth

Page 67: Revision of Learning Earth Sciences 7-30-04-Nir

67

systems approach) together with a research agenda, curriculum development for all the

learning environments, preparing teachers for the implementation of the new curriculum

materials, and the teaching strategies and tactics that are appropriate for each learning

environment (classroom, laboratory, outdoor and computer). The vision encompasses

how learning earth sciences may contribute to gaining insight into the nature of scientific

investigation and scientific reasoning in several contexts. Nevertheless, the conclusion

remains that depending upon the earth science disciplines in isolation, either from each

other or from the humanities and social sciences, to set the agenda for learning earth

sciences will fail to serve the public good. We need to respect students, their families, and

their communities as sources of ideas, issues, and problems to solve through application

of knowledge about earth systems.

Research has a central role in this holistic plan. It should provide an

understanding of students’ difficulties with the learning process and identify the

appropriate learning and teaching strategies for overcoming cognitive barriers to spatial

and temporal thinking, to retrospection, to understanding phenomena across scales, to

integrating several subjects, and to developing the cognitive capacity for systems

thinking. In addition, the research agenda should provide the basis for the development of

curriculum materials, the sequencing of learning, and productive paths for teachers to

follow in overcoming internally and externally imposed barriers to reform. We know

much too little from a research perspective about thoroughly contextualized, fully

integrated, earth systems thinking linked to environmental studies and centered on

students personal and social lives. If we are to have curricula that do these things, then we

must understand better what obstacles are and how to overcome them.

The good news that emerges from this chapter is that there are sound studies that

have already been done that can show the way for progress, but the better news is that

these studies are still few and there is room for many young researchers to join the

bandwagon and make their mark in the earth science education field and in the future of

humankind on earth.

Page 68: Revision of Learning Earth Sciences 7-30-04-Nir

68

References

Agelidou, E., Balafoutas, G., & Gialamas.,V. (2001) Interpreting how third grade junior high school students represent water. International Journal of Education and Information, 20, 19-36. American Association for the Advancement of Science (AAAS) (1990). Science for All Americans. New York: Oxford University Press. American Association for the Advancement of Science (AAAS) (1993). Benchmarks for science literacy. New York: Oxford University Press. Ault, C.R., Jr. (1981). Children’s concepts about time no barrier to understanding thegeologic past. Unpublished doctoral dissertation, Cornell University, Ithaca. Ault, C.R. Jr. (1982). Time in geological explanations as perceived by elementary schoolstudents. Journal of Geological Education, 30, 304-309.

Ault, C.R., Jr. (1984). The everyday perspective and exceedingly unobvious meaning.Journal of Geological Education, 32, 89-91 Ault, C.R., Jr. (1998). Criteria of excellence for geological inquiry: The necessity of ambiguity. Journal of Research in Science Teaching, 35, 189-212. Bar, V. (1989). Children's views about the water cycle. Science Education, 73, 481-500.

Bar, V and Travis, A. S. (1991) Children's views concerning phase changes. Journal of Research in Science Teaching, 28, 363-382. Ben-zvi-Assaraf, O. and Orion, N. (2004). The development of system thinking skills in the context of earth systems education. The Journal of Research in Science Teaching. Accepted for publication. Bezzi, A. 1995. Personal construct psychology and the teaching of petrology atundergraduate level. Journal of Research in Science Teaching, 33, 179-204. Bezzi, A., and Happs, J. C. (1994). Belief systems as barriers to learning in geological education. Journal of Geological Education, 42, 134-140. Bloom, B.S. (1956). Taxonomy of Educational Objectives, Handbook I: Cognitive Domain, N.Y.: David McKay. P. 196. Brody, M.J. (1994). Student science knowledge related to ecological crises. International Journal of science teaching, 16, 421-435. Bybee, R.W. (1993). Reforming science Education - Social perspectives & personal reflections. Teachers College Press, Columbia University. P. 198

Page 69: Revision of Learning Earth Sciences 7-30-04-Nir

69

Bybee, R.W and Deboer, G.E. (1994). Research on goals for the science curriculum. In Gabel, D. (Ed.), Handbook of research in science teaching and learning. 357-388. MacMillian, New York. Carpenter, J.R. (1996). Models for effective instruction of Earth science teachers in the USA. In D. A. Stow & G.J. McCall (Eds.), Geosciences education and training in schools and universities, for industry and public awareness. AGID special publication series No 19. Rotterdam: Balkema.

Page 70: Revision of Learning Earth Sciences 7-30-04-Nir

70

Chadwick, P. (1978). Some aspects of the development of geological thinking.Journal of Geological Teaching, 3, 142-148. Chang, C.Y. (2004). Could a laptop plus the liquid crystal display projector amount to

improved multimedia geoscience instruction? Journal of computer Assisted Learning. 20, 4-10.

De Jong, T. and van Joolingen, W.R. (1998). Scientific discovery learning with computer simulations of conceptual domains. Review of Educational Research, 68, 179-201. DeLaughter, J., Stein, S., Stein, C. and Bain, K. (1998). Preconceptions about earth science among students in an introductory course. Eos, 79, 429. Department of Education and Science/Welsh Office. (1989). Science in the National Curriculum. Her Majesty’s Stationery Office. Dewey, J., (1896). The reflex arc concept in psychology. Psychological Review, III, 357-370. Dickerson, D.L., (2003). Naïve conceptions about groundwater among pre-service science teachers and secondary students. Paper presented in March 2003 in Holly Springs, NC at the Annual Conference of the North Carolina Association for Research in Education Dodick, J. T. and Orion, N. (2003a). Cognitive factors affecting student understanding of geological time. Journal of Research in Science Teaching, 40 (4), 415-442. Dodick, J. T., and Orion, N. (2003b). Measuring student understanding of “deep time”. Science Education, 87(5), 708-731. Dodick, J. T., and Orion, N. (2003c). Geology as an Historical Science: Its Perception within Science and the Education System. Science and Education, 12 (2), 197-211. Dori, Y. J., & Barak, M. (2001). Virtual and Physical Molecular Modeling: Fostering

Model Perception and Spatial Understanding. Educational Technology & Society, 4(1), 61-74.

Dove, J.E. (1999). Exploring a hydrological concept through children’s drawings. International Journal of Science Education. 21, (5) 485-497

Dove, J.E. (1997). Student ideas about weathering and erosion. International Journal of Science Education, 19:8, 971-980. Driver, R.,Guesne, E. and Tiberghien, A. (1985). Children's ideas in science. Open University Press, Milton Keynes. P. 208 Edelson, D.C., Gordin, D.N., & Pea R.D. (1999). Addressing the challengeinquirybased learning through technology and curriculum design. The Journal ofLearning Sciences, 8, 391-449.

Page 71: Revision of Learning Earth Sciences 7-30-04-Nir

71

Emery, R. E. (1992). Parenting in context: Systemic thinking about parental conflict aninfluence on children. Journal of Consulting and Clinical Psychology, 60, 909-9

Everitt, C.L., Good, S.C. and Pankiewicz, P.R. (1996). Conceptualizing the

inconceivable by depicting the magnitude of geological time with a yearlyplanning calendar. Journal of Geoscience Education, 44, 290-293.

Faughnan, J. G. & Elson, R (1998). Information technology and the clinical curriculum

Some predictions and their implications for the class of 2003. Academic Medici73, 766-769.

Fetherstonhaugh, A. and Bezzi, A. (1992). Public knowledge and private understanding: Do they match? An example with the water cycle. Paper presented at the 29th international Geological congress, Kyoto, Japan, (August - September 1992). Fordyce, D. (1988). The development of systems thinking in engineering education:An interdisciplinary model. European Journal of Engineering Education, 13, 283-292. Fortner, R.W and Mayer, V.J. (Eds), (1998) Learning about the Earth as a system. Proceedings of the second International Conference on Geoscience Education. Colombus, OH: earth systems education, The Ohio State University. ERIC DocumentED422163. Frodeman, R. L. (1995). Geological reasoning: Geology as an interpretive and historical science. Geological Society of America Bulletin, 107(8), 960-968. Frodeman, R. L. (1996). Envisioning the outcrop. Journal of Geoscience Education, 44, 417-427. Friedman, W. (1982). Conventional time concepts and children’s structuring of time.

In W. Friedman (Ed.), TheDevelopmental Psychology of Time. New York: Academic Press.

Germann, P.J., Haskins, S., & Auls, S. (1996). Analysis of nine high school biology

laboratory manuals: Promoting scientific inquiry. Journal of Research in Science Teaching, 33, 475-499.

Gibson, J.J. (1966). The senses considered as perceptual systems. Boston: Houghton

Mifflin. pp. 335. Gilbert, J., Osborne, R and Fensham, P. (1982). Children's science and its consequencefor teaching. Science Education, 66,623-633. Glasersfeld, E. von (1987). Learning as a constructive activity. In: Jannet, C. (ed.) Problems of Representation in Teaching and Learning of Mathematics. Gobert, J. and Clement, J. (1998). Effects of student-generated diagrams versus student-generated summaries on conceptual understanding of causal and dynamicknowledge in plate tectonics. Journal of Research in Science Teaching, 36, 39-53.

Page 72: Revision of Learning Earth Sciences 7-30-04-Nir

72

Gobert, D.J. (2000). A typology of casual models for plate tectonics: Inferentialpower and barriers to understanding. International Journal of science Education. 22, 937-977.

Graczyk, S. L. (1993). Get with the system: General systems theory for business offici

School Business Affairs. 59, 16-20. Greeno, J., Collins A., and Resnick, L. (1996). Cognition and learning. In Handbook

of educational psychology. Gudovitch, Y. and Orion, N. (2001). The carbon cycle and the Earth systems -

Studying the carbon cycle in an environmental multidisciplinary context. Proceedings of the 1st IOSTE Symposuim in Southern Europe, Paralimni, Cyprus.

Gudovitch, Y. and Orion, N. (2003). Happs, J. C. (1982). Some aspects of student understanding of two New

Zealand landforms. New Zealand Science Teacher, 32, 4-12. Happs, J.C., (1982). Some aspects of student understanding of rocks and minerals. Working paper of the Science Education Research Unit, University of Waikato, New Zealand. Happs, J.C., (1985). Some aspects of student understanding of glaciers and mountains. Working paper of the Science Education Research Unit, University of Waikato, New Zealand. Herron, M.D. (1971). A critical look at practical work in school science. School

Science Review, 71, 33-40. Hsi, S., Linn, M. C., & Bell, J. (1997). The role of spatial reasoning in engineering

and the design of spatial instruction. Journal of Engineering Education, 86(2), 151-158.

Hume, J. D. (1978). An understanding of geological time. Journal of Geological Education, 26, 141-143. IGEO, (1997). Fortner, R.W and V.J. Mayer, (Eds), Learning about the Earth as a system. Proceedings of the second International Conference on Geoscience Education (1998). Colombus, OH: earth systems education, The Ohio State University. ERIC Document ED422163. IGEO, (2000). Conference proceedings of the 2nd meeting of the International Geosciences Education Organization. Editor: Clark, I. Sydney, Australia. IGEO, (2003). Conference proceedings of the 3rd meeting of the International Geosciences Education Organization. Calgary, Canada.

Page 73: Revision of Learning Earth Sciences 7-30-04-Nir

73

Kali, Y. and Orion, N. (1996). Relationship between Earth science education and spatial visualization. Journal of Research in Science Teaching. 33, 369-391. Kali, Y. and Orion, N. (1997). Software for assisting high school students in thespatial perception of geological structures. Journal of Geoscience Education. 45, 10-21. Kali, Y. and Orion, N. (2004). The Effect of Knowledge Integration Activities on Students’ Perception of the Earth’s Crust as a Cyclic System. Submitted for publication. Kali, Y., Orion, N. and Elon, B. (2003). A Situative Approach for Assessing the Effect of an Earth-Science Learning Program on Students’ Scientific Thinking Skills. Journal of Research in Science Teaching, Kapulnick, E., Orion, N. and Ganiel, U. (2004). In-service “science & Technology” training programs: What changes do teachers undergo? Paper accepted for publicationin the NARST annual meeting, Vancuver, USA. King, C. (2000). The Earth’s mantle is solid: Teachers’ misconceptions about the Earth and plate tectonics. School Science Review, 82(298), 57-65. King, C. (2000). Krajcik, J. Blumenfeld, P.C., Marx, R.W., Kristin M.B., & Fredricks, J. (1998).

Inquiry in project-based science classrooms: Initial attempts by middle schoolstudents, 7, 313-350.

Leighton, J, and Bisanz, G. L. (2003). Children's and adults' knowledge and reasoning about the ozone layer and its depletion. International Journal of Science Education, 25, 117-139. LeGrand, H. (1991). Drifting Continents and Shifting Theories. Cambridge University

Press, Cambridge. Lewis, J. P. (1998). Mastering project management: Applying advanced concepts of

systems thinking, control and evaluation, resource allocation, N.Y.: McGraw-Hill. Lillo, J. (1994). An analysis of the annotated drawings of the internal structure of the Earth made by students aged 10-15 from primary and secondary schools in Spain. Teaching Earth Sciences, 19(3), 83-89. Libarkin, J.C., Anderson, S., Author, Boone, W., Dahl, J., and Kurdziel, J. (2004). College students’ ideas about geologic time, Earth’s interior, and Earth’s crust. Journal of Geoscience Education. (in press) Lovelock, J. (1991). Healing Gaia - Practical medicine for the planet. Harmony books, New York. P. 192

Page 74: Revision of Learning Earth Sciences 7-30-04-Nir

74

Marques, L. F., and Thompson, D. B. (1997). Portuguese students’ understanding atage 10/11 and 14/15 of the origin and nature of the Earth and the developmentof life. Research in Science and Technology Education, 15, 29-51

Marques, L. and Thompson D. (1997). Misconceptions and conceptual changesconcerning continental drift and plate tectonics among Portuguese students aged 16-17. Research in Science and Technological Education 15(2), 195. Marques, L. Paria, J. and Kempa R. (2003). A study of students' preconceptions of theorganisazion and effectiveness of fieldwork in earth sciences education. Research in Science and Technological Education 21, 265-278. Mayer, V.J., and Armstrong, R.E. (1990). What every 17 year old should know about

planet earth: The report of a conference of educators and geoscientists. Science Education, 74(2), 155-165.

Mayer, V.J. (1995). Using the Earth system for integrating the science curriculum. Science Education, 79, 375-391. Mayer, V.J. (ed.) 2002. Global science literacy. Dordrecht, The Netherlands: Kluwer.

Mayer, V.J. and Fortner, R.W. 1995. Science is a study of earth. Columbus: Earth systems Education Program, The Ohio State University.

Mayer, V.J. and Fortner, R.W. 2002. A case history of science and science education policies. IN: Mayer, V.J. (ed.) 2002. Global science literacy. Dordrecht, The Netherlands: Kluwer.

Mayer, V.J. (2003). (Ed.), Implementing global science literacy. Ohio State University. McAuliffe, C., Hall-Wallace, M., Piburn, M., Reynolds, S., and Leedy, D. E. (2000). Visualization and Earth Science Education. GSA Abstracts with Programs, 32, A-266. McPhee, J. (1980). Basin and Range. Farrar; Straus & Giroux: New York. Metzger, E. P. (1992). The strategy column for pre-college science teachers: Lessons

on time. Journal of Geological Education, 40, 261-264 Meyer, W., (1987). Venacular American theories of earth science. Journal of Geological Education, 35:193-196. Midyan, Y. (2003). The effect of in-service training on science teachers attitudes and practical use of the outdoor as a learning environment. Unpublished MSc thesis.Weizmann Institute of Science, Israel. Montagnero, J. (1992). The development of the diachronic perspective in children. In

F. Macar. Time Action and Cognition (pp. 55-65). Amsterdam: Kluwer Academic Publishers.

Montagnero, J. (1996). Understanding changes in time. London: Taylor and Francis.

Page 75: Revision of Learning Earth Sciences 7-30-04-Nir

75

Noonan-Pulling, L. C., and Good, R. G. (1999). Deep time: Middle school students’ ideas on the origins of earth and life on earth. Paper presented at the National Association for Research in Science Teaching Annual Meeting, Boston, MA. Nottis, K., & Ketter, K. (1999). Using analogies to teach plate tectonics concepts. Journal of Geoscience Education, 47(5), 449-54. Osborne, R. and Wittrock, M. (1985). The generative learning model and its implications for learning science. Studies in Science Education, 5, 1-14. Ossimitz, G. (2000). Development of Systems Thinking Skills web site: http:www-sci.uni-klu.ac.at/~gossimit. Orion, N. (1996). An holistic approach to introduce geoscience into schools: The Israeli model - from practice to theory. In D. A. Stow & G.J. McCall (Eds.), Geosciences education and training in schools and universities, for industry andpublic awareness. AGID special publication series No 19. Rotterdam: Balkema. Orion, N. (1998). "Earth science education + environmental education = Earth systems education". In Fortner, R.W and V.J. Mayer, (Eds), Learning about the Earth as a system. Proceedings of the second International Conference on Geoscience Education (pp. 134-137). Colombus, OH: earth systems education, The Ohio State University. ERIC Document ED422163. Orion, N. (1993). A practical model for the development and implementation of field trips, as an integral part of the science curriculum. School Science and Mathematics, 93 (6), 325-331. Orion, N. and Hofstein, A. (1994). Factors that influence learning during scientific field trips in a natural environment. Journal of Research in Science Teaching. 31 (10), 1097-1119. Orion, N., Ben-Chaim, D. and Kali, Y. (1997). Relationship between Earth scienceeducation and spatial visualization. Journal of Geoscience Education, 45, 129-132. Orion, N. (2002). “An earth systems curriculum development model”. In Mayer, V. (Ed.), Global science literacy (pp.159-168). Dordecht, The Netherlands: Kluwer. Orion, N., Dubowski, Y. and Dodick, J. (2000). “The educational potential ofmultimedia authoring as a part of earth science curriculum - A case study”. Journal of Research in Science Teaching. 37, 1121-1153. Orion, N. (2003a). “The outdoor as a central learning environment in the global science literacy framework: from theory to practice”. In Mayer, V. (Ed.), Implementing global science literacy (pp.33-66). Ohio State University. Orion, N. (2003b). “Teaching global science literacy: a professional development or a professional change”. In Mayer, V. (Ed.), Implementing global science literacy (pp.279-286). Ohio State University.

Page 76: Revision of Learning Earth Sciences 7-30-04-Nir

76

Orion. N., and Fortner W. R. (2003). Mediterranean models for integrating environmental education and earth sciences through earth systems education. Mediterranean Journal of Educational Studies. 8, (1) 97-111. Oversby, J. (1996) Knowledge of earth science and the potential for its development.School Science Review, 78, 91-97. Pallrand, G. J., & Seeber, F. (1984). Spatial ability and achievement in introductory

physics. Journal of Research in Science Teaching, 21, 507-516. Piaget, J. (1970). Structuralism. Basic Books: New York. Pickering, K. and Owen, L. (1994). An Introduction to Global Environmental Issues. Routlege, London and New York. P. 390 Pribyl, J. R., & Bodner, G. M. (1987). Spatial ability and its role in organic chemistry:

A study of four organic courses. Journal of Research in Science Teaching, 24, 229-240.

Reynolds, S. J., Piburn, M. D., Leedy, D. E., McAuliffe, C. M., Birk, J. P., & Johnson, J. K. (2002). The Hidden Earth: Visualization of Geologic Featuresand their Subsurface Geometry. Paper presented at the National Associationfor Research in Science Teaching annual meeting, New Orleans, LA.

Riggs, E. M. (2003). Field-Based Earth Science Education and Indigenous

Knowledge: Cognitive and Cultural Elements of Geoscience Education for Native American Communities. Science Education, accepted for publication.

Riggs, E. M., & Tretinjak, C. A. (2003). Evaluation of the Effectiveness of a

Classroom and Field-Based Curriculum in Sedimentation and Change Through Time for Pre-Service Elementary School Teachers. Paper presented at the Geological Society of American annual meeting, Seattle, WA.

Riggs, E. M., & Semken, S. C. (2001). Earth Science for Native Americans.

Geotimes, 49(9), 14-17. Ritger, S. D., and Cummins, R. H. (1991). Using student created metaphors tocomprehend geological time. Journal of Geological Education, 39, 9-11. Roseman, J. (1992). Project 2061 and evolution. In R. Good, J.E. Trowbridge, S.

Demastes, J., Wandersee, M. Hafner, and C. Cummins (Eds.), Proceedings of the 1992 Evolution Education research conference (pp. 214-229). Baton Rouge: Louisiana State University.

Ross, K. and Shuell, T. (1993). Children’s beliefs about earthquakes. Science Education, 77(2), 191-205 Roth, W.M., & Roychoudhury, A. (1993). The development of science process skillsin authentic contexts. Journal of Research in Science Teaching, 30, 127-152.

Page 77: Revision of Learning Earth Sciences 7-30-04-Nir

77

Rowland, S. M. (1983). Fingernail growth and time-distance rates in geology. Journal of Geological Education, 31, 176-178. Schoon, K. J. (1989). Misconceptions in the earth sciences: A cross-age study. Paper presented at the 62nd Annual Meeting of the National Association for Research in Science Teaching, San Francisco, CA. Schwab, J.J. (1962). The teaching of science as inquiry. In J.J. Schwab & P.F.

Brandweine (Eds.), The teaching of science. Cambridge, MA: HarvardUniversity Press.

Shymansky, J.A., & Yore, L.D. (1980). A study of teaching strategies, student cognitive development, and cognitive style as they relate to student achievement inscience. Journal of Research in Science Teaching, 5, 369-382. Slattery, W., Mayer, V. and Klemm, B. (2002). Using the Internet in earth science systems courses. In V.J. Mayer (ed) Global Science Literacy, pp 93-107. London: Kluwer Academic Publishers Spencer-Cervato, C., and Daly, J.F. (2000). Geological time: An interactive team-oriented introductory geology laboratory. Teaching Earth Sciences, 25 (1), 19-22. Staver, J.R., & Small, L. (1990). Toward a clearer representation of the crisis in science education. Journal of Research in Science Teaching, 27, 79-89. Stow D. A. and McCall G.J. (Eds.), (1996), Geosciences education and training in schools and universities, for industry and public awareness. Proceedings of the conference on geosciences education, Southampton, UK, 1993. AGID special publication series No 19. Rotterdam: Balkema Stofflett, R. (1994). Conceptual change in elementary school teacher candidate knowledge of rock cycle processes. Journal of Geological Education, 42, 494-500. Taiwo, A., Motswiri, M., Ray, H., & Masene, R., (1999). Perceptions of the water cycle among primary school children in Botswana. International Journal of Science Education, 21(4): 413-429. Tamir, p. (1989). Training teachers to teach effectively in the laboratory. Science Education, 73, 59-69. Tank, R. (1983). Environmental Geology. New York, Oxford university press. P. 549. Tomorrow 98, (1993). Report of the superior committee on science, mathematics andtechnology education in Israel. Ministry of Education of Israel. pp115. Jerusalem. Trend, R. D. (1997). An investigation into understanding of geological time among 10and 11-year old children, with a discussion of implications for learning of othergeological concepts. Paper presented at 1st International Conference on GeoscienceEducation and Training, Hilo, Hawaii.

Page 78: Revision of Learning Earth Sciences 7-30-04-Nir

78

Trend, R. D. (1998). An investigation into understanding of geological time among 10-and 11-year old children. International Journal of Science Education, 20 (8), 973-988. Trend, R. D. (2000). Conceptions of geological time among primary teacher trainees,with reference to their engagement with geosciences, history and science. International Journal of Science Education, 22 (5), 539-555. Trend, R. D. (2001a). Deep Time Framework: a preliminary study of UK primaryteachers’ conceptions of geological time and perceptions of geoscience. Journal of Research in Science Teaching. Vol. 38 No. 2: 191-221. Trend, R. D. (2001b). An investigation into the understanding of geological time among 17-year-old students, with implications for the subject matter knowledge of future teachers. International Research in Geographical and Environmental Education, 10, 298-321. Trend, R. D. (2002). Developing the Concept of Deep Time. In V.J. Mayer (ed)Global Science Literacy, Ch 13 pp 187-202. London: Kluwer Academic Publishers. Welch, W.W., Klopfer, L.E., Aikenhead, G.S., and Robinson, J.T. (1981). The role of

inquiry in science education: Analysis and recommendation. Science Education, 65, 33-50

Zohar, A. (1998). Result or conclusion? Students’ differentiation between

experimental results and conclusions. Journal of Biological Education, 32, 53-59.

Zwart, P. J. (1976). About time: A Philosophical Inquiry into the Origin and Nature ofTime. Amsterdam: North-Holland Publishing Company.