CHAPTER 23

Learning Earth Sciences

Nir Orion

Weizmann Institute of Science, Israel

Charles R. Ault, Jr.

Lewis & Clark College

Great news! I've just been accepted into graduate school in geology with the opportunity to work on a terrific project. 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 of 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 excited about getting started, and I can't believe that there is a project in geology so similar to what I have dreamed about doing.

—Message from a twenty-first century graduate student

This young graduate student's excitement echoes the themes and claims developed in this chapter. The message offers a glimpse into the nature of earth science inquiry. The proposed research crosses several disciplines, though housed in the geosciences, and has 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 will guide human actions in response to environmental change. The project has intrinsic appeal to some, social value to many. At the dawn of the twenty-first century, earth scientists are doing multidisciplinary and interdisciplinary research serving the public good.

The image of an optimistic student captures several characteristics proposed in this chapter as representative of the earth sciences. Section 1, “Distinctive characteristics,” introduces these features, arguing that they are crucial for guiding teaching and learning earth sciences. There follows a profile of earth science education worldwide, including trends evident over the past 25 years. This profile focuses on significant reforms in geosciences education undertaken at the very end of the twentieth century: the trend away from disciplinary-based science education toward 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 (American Association for the Advancement of Science [AAAS], 1990).

“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 with a discussion of learning environments for the earth sciences (outdoor and indoor classrooms; the earth science laboratory) and the prospect for cultivating environmental attitudes and insights from learning earth sciences.

Today's ambitious reform agenda, guided by the principle of “science for all,” scaffolds Section 3. Here the concern becomes how well, or how poorly, teachers have adapted to calls for changing their philosophies of teaching. Section 3 deals with the difficulties of reforming earth science education for science teachers who have limited content knowledge and who may lack the motivation to deal with new priorities among subjects, unfamiliar learning environments, and changes in teaching strategies.

The chapter concludes by challenging researchers to study teaching and learning in the earth sciences not only as historically practiced as a discipline-based curriculum, but also as increasingly practiced as integrated study. The conclusion acknowledges that, from a research 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 to embrace the implications of science for all: ambitious integration and social contextualization. At the same time, the next generation of curriculum designers must preserve distinctive characteristics of the earth sciences when setting objectives for student learning.

DISTINCTIVE CHARACTERISTICS

Every subject has something important to offer science for all. The challenge is one of establishing priorities. Essential features of what to teach ought to

  1. Encompass an “intellectually honest” (Bruner, 1960, p. 33) portrait of what scientists do (e.g., date rocks radiometrically).
  2. Emphasize ideas with high conceptual worth or value (Toulmin, 1972), ideas proved to advance thinking and solve problems (e.g., the law of superposition).

The host of individual fields that comprise the earth sciences, the need to integrate these subjects within schools, and the goal of contributing to science for all make characterizing distinctive features imperative. In effect, we are asking, What are the earth sciences about? and What's so important to learn from earth science? The earth sciences are simply about everything beneath our feet and above our heads, with concern for how our collective actions interact with these realms. 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. Learning earth sciences often means learning to think about processes linking the earth's oceans, atmosphere, land, interior, and orbit from a systems perspective. 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 kindergarten to high school.

Let us emphasize the notion of “distinctiveness” on four levels: disciplinary, psychological, pedagogical, and sociohistorical. 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 example, presumably, from learning earth sciences students acquire an intellectually honest understanding of change through earth's history across many scales. Developing a sense of scale is a distinctive feature of learning earth sciences. The concept of scale functions both psychologically and epistemologically. Psychologically, scale may present obstacles to perception and insight. Epistemologically, extrapolation of earth processes in time and space is a goal of explanation. The geologic time scale encompasses durations and changes vastly beyond the scale of human lifetimes; forecasts of global climate change must wrestle with problems of sampling and modeling on various scales. Such inquiry in the earth sciences has distinctive features; our synthesis highlights six:

  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).
  2. The concern for complex systems acting over the Earth as whole (e.g., the several “spheres”: hydro, geo, atmo, 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).
  6. The uniqueness of retrospective scientific thinking. 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 conditions of materials, processes, and environments of the past.

The earth sciences are both similar to and distinct from other fields of science. To the extent that earth sciences serve as concrete contexts for better understanding of basic concepts from physics, chemistry, and biology, they inherit many common challenges, for example, using operational definitions, thinking in terms of direct and inverse proportions, and overcoming pervasive misconceptions about energy, motion, particulate matter, inheritance, and adaptation (Driver, Squires, Rushworth, & Wood-Robinson, 1994; Driver, Guesne, & Tiberghien, 1985; Wandersee, Mintzes, & Novak, 1994). The readily accessible contexts for learning earth sciences may introduce young adolescents to features of scientific reasoning such as observing, hypothesizing, and drawing conclusions from evidence. At the same time, learning about the earth sciences presents distinctive challenges and opportunities.

The history and philosophy of science, when turned toward the examination of geological explanations and the concept of geologic time, reveal features of thought characteristic of the earth sciences (Ault, 1998; Brandon, 1994; Cleland, 2002; Gould, 1986; Kitts, 1977; LeGrand, 1988; Schumm, 1991). So, too, does the psychology of learning earth science concepts unveil what is cognitively distinctive about this field (Ault, 1994; Schoon, 1989; Trend, 1998, 2001b).

A strategy of compare and contrast has proved essential to forming understandings of earth's 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, 1986). The objects of explanation—such as mountain building, ice age onset, seafloor topography, storm generation, magma distillation, planetary coalescence, and 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 (e.g., the Nile River delta is similar to, yet distinct from, other examples of deltas due to similar, thought not identical histories of formation).

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 categories whose members are quite often utterly indistinguishable from each other (Hanson, 1965): protons, atoms of carbon, electromagnetic fields. At an important level, disciplines depart from each other in how they categorize and represent what is most salient about reality, with important consequences for learning from a constructivist standpoint (Driver, Asoko, Leach, Mortimer, & Scott, 1994).

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.

In conclusion, the distinctive features of the earth sciences stem from the centrality of historical methods of inquiry pioneered at the dawn of geology. In addition, investigating the earth depends heavily upon spatial reasoning and visual representation. 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 are combined with basic ideas from physical and life sciences, earth systems thinking emerges, with attention to dynamism on global scales of interest and the realization that human action affects 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.

The general themes of interdisciplinary study, multidisciplinary study, environmental issues, and relationship to social responsibility invariably lie close to the surface when learning earth sciences. “Holistic” properly describes this situation. Learning earth sciences offers holistic perspectives to science for all, and this holism entails a shift from traditional science teaching.

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?

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 contexts. 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 for society to exist in profitable harmony with earth resources.

Increased multidisciplinary and interdisciplinary research within the sciences and across 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 from the perspective of human interaction with natural systems (Tank, 1983; Pickering & Owen, 1994).

The time has come for science education to situate itself squarely within the educational conversation about social justice, poverty, wealth, sustainability, and the human condition. The National Science Education Standards (NSES) for the United States invite science educators to do so in the standard Science in Personal and Social Perspective (National Research Council, 1996). This conversation is at the same time about the nature of democratic institutions for governing the use of earth resources and affecting earth systems. The features distinctive of the earth sciences clearly align with these aims. Systems thinking, hierarchy theory, holistic explanation, and attention to scale and complexity bind learning earth sciences to environmental topics. Indeed, the environmental imperative and the role of learning earth sciences in order to achieve environmental insight has achieved a central position in the field of earth science education (Mayer & Armstrong, 1990; Brody, 1994; Mayer, 1995; Orion, 1996).

In many respects education in earth sciences has, in fact, converged upon environmental education in nations around the globe. 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). Whether from the point of view of integrating multiple disciplines, from the acknowledgement that the phenomena of interest are complex, interacting systems, or simply in response to the imperative of educating citizens for making environmentally responsible decisions, the profile of learning earth science has changed in recent decades. It will continue to change in the direction of holism, in the sense of the multidisciplinary study of complex systems and from the standpoint of environmental concerns.

Reductionist philosophy has historically constrained the introduction of earth sciences within school science curricula by prioritizing 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 paradigm works quite 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 toward a science for all paradigm places the earth sciences in a better position. The new paradigm sets the goal for science education as preparation for citizenship. Science for All Americans (AAAS, 1990) defines minimal levels of scientific literacy. The Benchmarks for Science Literacy (AAAS, 1993), which followed Science for All Americans, advocates balancing 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 National Curriculum for England and Wales (Department of Education and Science [DES], 1989). In the United States, the National Science Education Standards (NSES; National Research Council, 1996) encompass eight categories of content. Four are traditional categories (physical, life, earth and space, science and technology); two incorporate holistic conceptions of science (unifying concepts and processes, science as inquiry) and two examine science within wider contexts (science in personal and social lives, history and nature of science). Reform topics have equal billing with traditional subjects in the NSES.

The shift from direct instruction toward constructivist pedagogy also has influenced the profile of science education (Mintzes & Wandersee, 1998; Driver et al., 1994; Driver et al., 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 recognize, in addition, the importance representations of reality (models, diagrams, equations, and category systems) play within the epistemology of a subject (Driver et al., 1994). Learners must assume personal responsibility to construct these representations and compare their thinking with that of others when in pursuit of “shared meaning” (Gowin, 1981). 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, Strike, Hewson, & Gertzog, 1982; Smith, 1991) has also exerted a strong influence over science teaching. Conceptual change theory recognizes that beliefs about knowledge shape student efforts to learn science. Conceptual change theory, using historical examples of major shifts in scientific conceptualizations, focuses on the adequacy of ideas. 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, 1991).

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 attempts to detect enemy submarines during World War II; see LeGrand, 1988). Sea floor spreading made plausible the notion of drifting continents; the concept has proved enormously fruitful as a component (and precursor) of plate tectonic theory. Now, understandings of geologic hazards due to seismic and volcanic activity depend upon knowledge of plate tectonic theory. Public policy, from building codes to tsunami alerts, has made this knowledge useful in a social context.

Problems, projects, and issues often provide a proper context for promoting meaningful learning. No doubt many students are exposed in their daily lives or through the mass media to earthquakes, volcanoes, global atmospheric changes, journeys to Mars, ocean pollution, fresh water shortages, energy conservation, floods, hurricanes, landslides and avalanches, etc. 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. At least 15 countries have undertaken to reform science teaching by placing greater emphasis on the dynamic systems of the Earth (Mayer, 2002).

Earth Systems Science and Education

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 United States, Australia, and Canada (Stow & McCall, 1996; IGEO, 1997, 2000, 2003). At the first international conference in England participants widely supported the proposal to reinforce the environmental aspect of learning earth sciences (Carpenter, 1996; Orion, 1996; Mayer, 1996). In 1997 in Hawaii earth science educators convened again for an international conference, this time titled “Learning about the Earth as a System” (IGEO, 1997). Now, at the beginning of the twenty-first century, earth science educators accept that the purpose of earth science education for ages 5–19 is both to educate for citizenship and to prepare students to become professional geoscientists.

Orion and Fortner (2003) have argued that the earth systems approach is ideal as a holistic framework for science curricula. The starting point is the four earth systems: geosphere, hydrosphere, atmosphere, and biosphere. The study of cycles organizes earth systems education: the rock cycle, the water cycle, the food chain, and the carbon cycle. The study of these cycles emphasizes relationships among subsystems through the transfer of matter and energy based on the laws of conservation. Such natural cycles should be discussed within the context of their influence on people's daily lives, rather than being isolated to scientific disciplines. The earth systems approach also connects the natural world and technology: technology transforms raw materials that originate from earth systems.

Through the elaboration of cycles, the approach underscores that society is a natural part of the systems of the Earth and that manipulation of one part of this complex system might adversely affect people. In contrast with traditional approaches for teaching science, the earth systems approach does not sequence the curriculum using topics from physics or chemistry. Instead, this approach organizes study in terms of systems and cycles as experienced in peoples’ lives. It does utilize physics and chemistry as tools for understanding science at a deeper and more abstract level within this context. However, the main educational goal is the development of environmental insight in two senses. First, we live in a cycling world that is built upon a series of subsystems (geosphere, hydrosphere, biosphere, and atmosphere) that interact through an exchange of energy and materials. Second, people are a part of nature and thus must act in harmony with its laws of cycling.

Ten years after introducing the earth systems approach, Mayer introduced Global Science Literacy (GSL; Mayer, 1997, 2002, 2003). GSL expands the argument for new science curricula for secondary schools. Instead of presenting major disciplines, Mayer argued the importance of organizing curricula with the “Earth System” concept. This approach includes teaching the methodology of system sciences and capitalizing on the cross-cultural characteristics of science.

The Earth System concept embraces holism and extends learning earth sciences into environmental, 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 illustrates.

The year 2003 witnessed in the United States the inauguration of an unprecedented multidisciplinary, 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 2001, pp. 1–2)

EarthScope organizers fully expect to affect school and museum science in substantial ways, 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 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.

LEARNING EARTH SCIENCES

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.

Alternative Frameworks of Learners Concerning Earth Sciences Concepts

The constructivist paradigm has dominated the field of science education in recent decades, producing studies of misconceptions, preconceptions, naive ideas, and alternative frameworks. Although there are relatively few published studies of students’ alternative frameworks in earth sciences education, findings and patterns have emerged in four areas (see Ault, 1994, for an earlier review of this literature and related studies of “expert and novice” styles of solving earth science problems):

  1. Students’ conceptions of processes and mechanisms of geospheric change, including plate tectonics, the rock cycle, earthquakes, and erosion (Ault, 1984; Happs, 1985; Ross & Shuell, 1993; Bezzi & Happs, 1994; Lillo, 1994; Marques & Thompson, 1997a,b; Schoon, 1989; Gobert & Clement, 1998; Stofflett, 1994; Dove, 1997, 1999; Gobert, 2000; Kali, Orion, & Elon, 2003; Libarkin et al., 2005).
  2. Students’ and teachers’ understanding and conceptions of the Earth's interior (DeLaughter, Stein, Stein, & Bain, 1998; Gobert & Clement, 1998; Marques & Thompson, 1997a,b; Lilio, 1994; Nottis & Ketter, 1999; King, 2000; Beilfuss, Dickerson, Boone, & Libarkin, 2004).
  3. Students’ and teachers’ perceptions of geological deep time (Happs, 1982a,b; Marques, 1988; Oversby, 1996; Schoon, 1989; Marques Thompson, 1997a; Noonan-Pulling & Good, 1999; Trend, 1997, 1998, 2000; Dodick & Orion, 2003a, 2003b).
  4. Students’ and teachers’ conceptions of hydrospheric processes and the water cycle (Meyer, 1987; Fetherstonhaugh & Bezzi, 1992; Brody, 1994; Taiwo, Ray, Motswiri, & Masene, 1999; Agelidou, Balafoutas, & Gialamas, 2001; Dickerson, 2003; Ben-zvi-Assaraf & Orion, 2005; Beilfuss et al., 2004).

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 emerge as students encounter difficult abstractions about the Earth in conflict with the scale of their everyday perceptions. For example, 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.

Review of these studies leads to 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 nonproductive cycle.

Visualization and Spatial Reasoning

Teaching and learning earth sciences at all levels relies upon spatial reasoning. The phenomena of interest sometimes have simple geometries, though 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.

Often earth science phenomena have challenging geometries. As a result, earth scientists use visual representations 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. 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.

Although the basic dependence of geoscientists on spatial abilities has long been recognized (Chadwick, 1978), the geoscience education community has only begun to explore the array of spatial reasoning abilities for learning earth sciences (McAuliffe, Hall-Wallace, Piburn, Reynolds, & Leedy, 2000). These spatial reasoning abilities may, in fact, be quite distinct from those commonly associated with tasks in learning chemistry (Dori & Barak, 2001; Pribyl & Bodner, 1987), physics (Pallrand & Seeber, 1984), and engineering (Hsi, Linn, & Bell, 1997).

The spatial objects that are studied in the geological sciences are usually large enough to walk in physically (the field learning environment). Block models can also readily represent them, as can more sophisticated renderings in a virtual setting. In the earth sciences, these blocks are visualized, but rotated, inspected, and modified to reflect temporal changes.

An understanding of deep geologic time also is associated with spatial cognition (Dodick & Orion, 2003a, 2003b). There is additional evidence that the outdoor field learning environment enhances the ability to construct a coherent narrative for layers of sedimentary rocks as experienced in the field (Orion, Ben-Chaim, & Kali, 1997; Riggs & Tretinjak, 2003).

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 requires a special type of spatial visualization, which they named VPA (Visual Penetration Ability). Spatial visualization is 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. Hsi, Linn, and Bell (1997) have also demonstrated the advantage of virtual worlds and computer tools in improving learners’ capability to solve problems that require spatial skills. They found that students acquired spatial skills in relatively short time with the use of technological tools.

The NSF-funded Hidden Earth Project (Reynolds et al., 2002) 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 by 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.

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.

Spatial cognition in a geoscientific problem-solving context must address more than two-dimensional representations of three-dimensional objects. In a geoscientific problem, students inspect three-dimensional objects and infer temporal histories from spatial features.

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). There is probably some robust connection between place-based, indigenous cultures and their success in field-based learning spatial reasoning. Clearly, experience plays an essential role in developing spatial reasoning ability.

Temporal Thinking

In the history of geology two discoveries, plate tectonics and geological time, have determined how geologists view the Earth. Geological time means the understanding (aptly referred to by John McPhee in 1980 as “deep time”) that the universe has existed for countless millennia, and that humanity's earthly dominion is confined to the last milliseconds of a metaphorical geological clock.

The understanding of geological time has shaped numerous disciplines, especially geology, cosmology, and evolutionary biology (Dodick & Orion, 2003c). Rose-man (1992) noted in a review of the literature of science education that there “was next to nothing about … how kids’ understanding of notions of systems, scale or models develop over time” (p. 218). Since that time, there have been several large-scale studies of how students understand this concept. They divide roughly 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 the subject is asked to justify his or her reasons for the proposed temporal order. Such studies include: Noonan-Pulling and Good's (1999) research on the understanding of the origins of Earth and life among junior high students; a similar study by Marques and Thompson (1997a) with Portuguese students; and Trend's studies on the conception of geological time among 10–11-year-old children (Trend, 1997, 1998), 17-year-old students (Trend, 2001b), and primary teacher trainees (Trend 2000, 2001a).

Qualitative research (structured interviews) with small sample groups dominates the literature. There has never been a large-scale, quantitative study of older students’ (junior high to senior high) understanding of geological time.

In logic-based studies, the researcher is interested in the cognitive processes undergone by students when they are 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 his or her 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 40 students from grades kindergarten, two, four, and six, using a series of puzzles that 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 to their understanding of geologic events. Many of the children in his test group were successful at solving puzzles involving skills necessary to understanding the logic, though not the extent, of geological time. 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 their being dark or crumbly—not based upon their position in a series of strata.

Piaget's (1969) work on time cognition influenced and restricted Ault's (1981) research design. 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, 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 among strata should necessarily allow one to both conceptualize and internalize the entirety of geological time.

Dodick and Orion (2003a, 2003b) conducted a large-scale study with junior high and high school students using 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. A 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.

Montagnero (1996) argues that there are four schemes, which are activated when one attempts to reconstruct transformational sequences. Dodick and Orion translated three of these schemes into the logical skills needed to solve temporal problems about 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 among strata (ex: superposition) are used in determining temporal organization.
  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, Montagnero designed a specialized (validated) instrument, the GeoTAT, consisting of a series of open puzzles that tested the subject's understanding of diachronic schemes as applied to geological settings.

In addition, two other questionnaires were distributed to subunits 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 and (b) a Stratigraphic Factors Test (SFT), which tested the influence of (geological strata) dimensions on students’ temporal understanding. In addition, researchers pursued qualitative research in the classroom and field by studying and interviewing students who were studying geology and paleontology as part of their matriculation studies.

From this study Montagnero constructed a model of temporal thinking. This model identified abilities needed to reconstruct geological features in time:

  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. Extracognitive factors, such as spatial-visual ability, that influence how a subject temporally organizes three-dimensional structures such as geological strata.

Among 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 ninth-grade students (on the one hand) and seventh-grade students (on the other) in their ability to understand geological phenomena with the use of diachronic thinking. This suggests that somewhere between grades 7 and 8 it should be possible to start teaching some of the 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 latter 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 three-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 show 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 fieldwork 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 among temporal/spatial/and field abilities.

In addition to the studies mentioned above, one might add the small body of research that catalogues general misconceptions in geology and includes within its parameters problems related to geological time (Happs, 1982a,b; Marques, 1988; Oversby, 1996; Schoon, 1989). Finally, one might note those works that have focused on the practical elements of teaching the scale of time (Everitt, Good, & Pankiewicz, 1996; Hume, 1978; Metzger, 1992; Ritger & Cummins, 1991; Rowland, 1983; Spencer-Cervato & Day, 2000). Unfortunately, these teaching models have never been critically evaluated, so they are of untested value to the pedagogic literature.

Systems Thinking

Current earth science education is characterized by a shift toward a systems approach to teaching and curriculum development (Mayer, 2002). Earth science educators call for reexamination of 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 an effect in another part, sometimes in ways that are quite unexpected.

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, 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 1990s and the beginning of this decade three studies were conducted at 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, 1977) 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; 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 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 seventh-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.

Kali, Orion, and Elon (2003) reported that while answering an open-ended questionnaire about the rock cycle, 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. At the low end of this continuum they located students who expressed a lack of connectedness between parts of the system, indicated poor dynamic thinking, and represented a completely static view of the rock cycle system. At the opposite end of the continuum there were the students who thought dynamically 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.

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 the basis for developing more sophisticated models, until the highest level of understanding the cyclic nature of the system is reached.

Ben-zvi-Assaraf and Orion (2005) used a large battery of qualitative and quantitative research tools in order to explore the development of systems-thinking skills of junior high school students who studied the water cycle as part of the “Blue Planet” program. The pre-test findings indicated that most of the students sampled experienced substantial difficulties in all aspects of systems thinking. They even struggled to identify basic system components. They entered the eighth grade holding an incomplete and naive perception of the water cycle and were only acquainted with the atmospheric component of the cycle (i.e., evaporation, condensation, and rainfall). They 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 as stages linked through processes. Ben-zvi-Assaraf and Orion (2005) found the same phenomenon of disconnected “islands of knowledge” that Kali, Orion, and Elon (2003) reported in reference to how students conceived of the rock cycle. Most of the students were not able to link the various components of the water cycle together into a coherent network of processes and stages. Some of them demonstrated an ability to create a relationship between several components, but even those students were not able to draw a complete network of relationships.

The post-test findings indicated that most of the students shifted from a fragmented conception of the water cycle toward a more holistic view. About 70% of the students who initially grasped only the atmospheric component of the water cycle significantly increased their familiarity with the other stages and processes of the water cycle. For about half of the students, this knowledge improved their ability to identify relationships among the stages and processes of the water cycle. The classification of students’ achievements indicated that the development of systems thinking in the context of the earth systems consists of several sequential stages arranged in a hierarchical pyramid structure. The findings of a hierarchical notion and the interrelationships between dynamic perception and cyclic perception are in accordance with the studies of Kali, Orion, and Elon (2003) and Gudovitch (1997). 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 to improvement of students’ abilities to develop systems thinking:

  1. Focusing on inquiry-based learning.
  2. Using the outdoor learning environment for the construction of a concrete model of a natural system.
  3. Using knowledge integration activities throughout the learning process.

There are interesting connections among the several cognitive studies mentioned above. For example, Dodick and Orion (1993a) reported an interrelationship between temporal thinking ability and spatial thinking ability. Orion, Ben-Chaim, and Kali (1997) and Riggs and Tretinjak (2003) determined that geological outdoor experiences tended to increase students’ spatial thinking abilities. Ben-zvi-Assaraf and Orion (1994) found systems thinking about the Earth to be 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 turned out to be a very effective tool for developing a concrete, realistic perception of nature serving as a cognitive bridge for the development of abstract thought: temporal, spatial, and systems thinking. Moreover, all of the above studies acknowledged the significance of alternative frameworks and experiences that most students bring to earth science classes (no matter what age), thus indicating the need to respond to preconceptions and misconceptions with appropriate instruction, whether in the laboratory, the outdoors, or the classroom, or when working with computers.

Despite the limited amount of research about learning earth sciences, a holistic framework has emerged to guide teachers who work within an earth systems approach. Holism in this sense refers not only to the systems approach, but also to the interconnectedness of spatial reasoning and temporal thinking and the cultivation of environmental insight.

The Integration of Learning Environments within the Earth Sciences

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 site, museum, or industrial site), and the virtual worlds of computers.

The outdoor learning environment. Review of the proceedings of the three International Geosciences Education Organization international conferences on geo-science education (IGEO, 1997, 2000, 2003) indicates 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. The guiding principle of this model is a gradual progression from the concrete levels of the curriculum toward 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).

Orion's holistic model combines indoor and outdoor environments. In this model the learning process begins 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.

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. Familiarity with properties and possibilities is the principal outcome—the raw material for forming concepts and posing questions. Kempa and Orion (1996) add that the outdoor learning environment may introduce the methodology of field research from 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.

One point is most crucial to understand: the outdoor learning environment addresses phenomena and processes that cannot be cultivated indoors. The outdoors, however, is a very complicated learning environment and includes a large number of stimuli that can easily distract students from meaningful learning.

Consider a location where students find that an outcrop reveals an anticline. They 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 at 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 & Orion, 1997).

The main aim of the initial indoor phase is to prepare the students for their outdoor learning activities. This preparation reduces what Orion and Hofstein (1994) term the “novelty space” of an outdoor setting (Fig. 23.1). 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.

images

FIGURE 23–1. The three dimensions, which identify the novelty space of an outdoor learning activity.

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 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, and maps, and second 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.

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 that might be answered according to the evidence uncovered at the specific outdoor site. Otherwise time and resources, including the students’ attention, are 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. Furthermore, they found a positive influence of this learning environment on students’ learning. However, their study also highlighted the difficulties teachers faced in adapting to the novel, outdoor learning environment.

Geo3D software (Kali & Orion, 1996) clearly 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 a computer simulation for the development of spatial visualization (Kali & Orion, 1996). However, without previous concrete outdoors experience with geological structures, such software loses much of its relevance for many students.

Integrating inquiry and the laboratory learning environment. 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 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.

The earth science laboratory environment has great potential to contribute to the development of the skills of scientific inquiry reasoning. Inquiry in the geo-sciences 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 that “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 “observational sciences” that emphasize comparisons and contrasts among features of the Earth in different times and places. 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 by rules of logic. They offer different milieus for illustrating the meaning of some of the most basic constructs of scientific thinking: for example, observations, hypotheses, and conclusions.

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). With such methods, inquiry-based curricula can be placed anywhere on a 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 to design, conduct, and analyze their own investigations, then communicate their findings. However, the appropriate stages for engaging students in open inquiry are 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 Veen-man (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, and 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).

Learning earth sciences has a role to play in remedying this situation. Orion and Kali (2005) suggest that 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.

Orion and Kali (2005) 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 seventh and eighth 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 seventh- and eighth-grade students included in this study had considerable difficulty 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, Orion and Kali 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 their traditional manner. Appropriate curriculum materials are not sufficient in themselves for inducing cognitive development among students. Sometimes teachers are the limiting factor in students’ ability to exploit the potential of “The Rock Cycle” in developing scientific thinking skills.

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. In this section we report in detail about a curriculum for teaching the water cycle from an earth systems and environmental insight perspective. The curriculum “The Blue Planet” emerged from a “design research” effort.

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.

Pre-development of “The Blue Planet” Curriculum

Based upon Orion's helical model, research preceded and followed development for eighth-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 student misconceptions and to validate the quantitative tools.

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 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 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.

Transcription and qualitative analysis of the questionnaires from the predevelopment study for the Blue Planet curriculum indicated that most of the students demonstrated an incomplete picture of the water cycle and held many misconceptions about it. Children who drew the water cycle usually represented the upper part of the water cycle (evaporation, condensation, and rainfall) and ignored the groundwater system. More than 50% of the students could not identify components of the groundwater system even when they were familiar with the associated terminology. In their mind, underground water was a static, subsurface 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.”

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 among the various earth systems. 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 (Edmondson, 1999). In addition, regular observations were conducted in the classes.

Observations indicated that teachers concentrated primarily on scientific principles and only very little on the cognitive aspects of the connections between the water cycle and 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.

A significant improvement was found in the student's level of knowledge, namely acquaintance with the components of the water cycle. 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.

The analysis of the cyclic and systemic thinking questionnaires showed some improvements in students’ understanding of interrelationship among 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 demonstrated 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 falls short of the development of environmental insight. For environmental insight, students must develop cyclic and systemic thinking about what happens to water in the air, on the Earth's surface, and underground. Teachers should not overlook activities developed directly 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.

PEDAGOGICAL INERTIA AND THE STRUGGLE FOR PARADIGM SHIFT

The science for all paradigm holds promise around the globe. However, its implementation will take decades—perhaps until the return of Haley's Comet in 2061. For learning earth sciences, science for all means cultivating environmental insight through the study of earth systems.

For teachers to move from traditional science teaching to proper 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 toward holism and attention to educating students for lives of social responsibility within democratic societies. The shift demands something more: that properly trained 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 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.

This decade of investigation has produced four Ph.D. dissertations (Kali & Orion, 2003; Dodick & Orion, 2003b; Ben-zvi-Assaraf & Orion, 2005; Kapulnick, Orion, & Gniel, 2004) and one master's thesis (Midyan, 2003). Altogether, these different studies examined the practice of science 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.

From each of these studies came the conclusion 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. Results indicated a clear gap between teachers’ perceptions of their development as expressed through questionnaires and interviews and their actual teaching practice.

In addition to teachers’ reluctance to implement new teaching methods and incorporate new scientific topics, the interviews uncovered four additional factors preventing them from genuinely implementing reform. They felt, in general, apprehension toward change and that professional training institutes did not provide them with the practical tools needed to overcome their apprehension. Teachers believed that school administrators failed to provide them with the resources necessary for reform, such as laboratory equipment, smaller class sizes in the laboratory, computers, and access to outdoor learning environments. Reform placed, in their judgment, inordinate demands upon their time. Finally, teachers faulted the Ministry of Education and its science education inspectors for a double standard. On the one hand, the Ministry initiated reform and inspectors encouraged participation. On the other hand, resources were not forthcoming and the Ministry called upon the inspectors to implement a national testing regime. The focus of testing tended to institutionalize objectives antithetical to the Science for All paradigm and the earth systems approach.

The world is complicated and diverse, and the Israeli example is that of just one nation. Movement toward an earth systems approach in keeping with the spirit of the Science for All paradigm is a change many teachers cannot or do not really want to undergo. Yet the conflict between reform efforts and testing priorities is worrisome and is 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.

CONCLUSION

The first decade of the twenty-first 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 educators to establish earth science 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 persistently 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 virtual worlds of the interior earth, its surface features, its motion in space, 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, and 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 humans 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 as environmental insight.

This chapter presents a holistic view of earth sciences education and a holistic perspective for achieving meaningful learning of the earth sciences. This perspective combines an educational vision (development of environmental insight through adopting the earth systems approach) together with a research agenda (curriculum development for outdoor and indoor, laboratory and computer, as well as classroom, learning environments). This vision and agenda acknowledge the challenge of preparing teachers for the implementation of new curriculum materials and adoption of teaching strategies and tactics appropriate for each learning environment.

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 the obstacles are and how to overcome them.

The good news that emerges from this chapter is that there are sound studies that demonstrate the way for progress. The better news is that these studies are still few, and there is room for many young researchers to join the groundswell and make their mark in earth science education and on the future of humankind on Earth.

ACKNOWLEDGMENTS

Thanks to Eugene Chiappetta and Gerald Krockover, who reviewed this chapter.

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 the geologic past. Unpublished doctoral dissertation, Cornell University, Ithaca.

Ault, C. R., Jr. (1982). Time in geological explanations as perceived by elementary school students. 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. (1994). Research on problem solving: earth science. In D. L. Gabel (Ed.), Handbook of research on science teaching and problem solving. New York: Macmillan.

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., & Travis, A. S. (1991). Children's views concerning phase changes. Journal of Research in Science Teaching, 28, 363–382.

Barker, M. (1998). Understanding transpiration—more than meets the eye. Journal of Biological Education, 33, 17–20.

Beilfuss, M., Dickerson, D. L., Boone, W., & Libarkin, J. (2004). Exploring conceptual understandings of groundwater through students’ interviews and drawings. Proceedings of the 77th Annual Meeting of the National Association for Research in Science Teaching, Vancouver, BC.

Ben-zvi-Assaraf, O., & Orion, N. (2005). The development of system thinking skills in the context of Earth System education. Journal of Research in Science Teaching, 42, 1–43.

Bezzi, A. (1995). Personal construct psychology and the teaching of petrology at undergraduate level. Journal of Research in Science Teaching, 33, 179–204.

Bezzi, A., & 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 (pp. 196). New York: David McKay.

Brandon, R. (1994). Theory and experiment in evolutionary biology. Synthese, 99, 59–73.

Brody, M. J. (1994). Student science knowledge related to ecological crises. International Journal of Science Teaching, 16, 421–435.

Bruner, J. (1960). The process of education. New York: Vintage Books.

Bybee, R. W. (1993). Reforming science education—Social perspectives & personal reflections. New York: Teachers College Press, Columbia University.

Bybee, R. W., & Deboer, G. E. (1994). Research on goals for the science curriculum. In D. L. Gabel (Ed.), Handbook of research in science teaching and learning (pp. 357–388). New York: Macmillan.

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.

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.

Cleland, C. (2002). Methodological and epistemic differences between historical science and experimental science. Philosophy of Science, 69, 474–496.

De Jong, T., & 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., & 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. London: Her Majesty's Stationery Office.

Dickerson, D. L. (2003, March). Naïve conceptions about groundwater among pre-service science teachers and secondary students. Paper presented at the Annual Conference of the North Carolina Association for Research in Education, Holly Springs, NC.

Dodick, J. T., & Orion, N. (2003a). Cognitive factors affecting student understanding of geological time. Journal of Research in Science Teaching, 40(4), 415–442.

Dodick, J. T., & Orion, N. (2003b). Measuring student understanding of “deep time.” Science Education, 87(5), 708–731.

Dodick, J. T., & 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. (1997). Student ideas about weathering and erosion. International Journal of Science Education, 19(8), 971–980.

Dove, J. E. (1999). Exploring a hydrological concept through children's drawings. International Journal of Science Education, 21(5), 485–497.

Driver, R., Asoko, H., Leach, J., Mortimer, E., & Scott, P. (1994). Constructing scientific knowledge in the classroom. Educational Researcher, 23(7), 5–12.

Driver, R., Guesne, E., & Tiberghien, A. (1985). Children's ideas in science. Milton Keynes, UK: Open University Press.

Driver, R., Squires, A., Rushworth, P., & Wood-Robinson, V. (1994). Making sense of secondary science: Research into children's ideas. London: Routledge.

Edelson, D. C., Gordin, D. N., & Pea, R. D. (1999). Addressing the challenges of inquiry based learning through technology and curriculum design. Journal of the Learning Sciences, 8, 391–449.

Edmondson, K. M. (1999). Assessing science understanding through concept maps. In J. J. Mintzes, J. H. Wandersee, & J. D. Novak (Eds.), Assessing science understanding: A human constructivist view. San Diego: Academic Press.

Elshout, J., & Veenman, M. (1992). Relations between intellectual ability and working method as predictors of learning. Journal of Educational Research, 85, 134–143.

Emery, R. E. (1992). Parenting in context: Systemic thinking about parental conflict and its influence on children. Journal of Consulting and Clinical Psychology, 60, 909–912.

Everitt, C. L., Good, S. C., & Pankiewicz, P. R. (1996). Conceptualizing the inconceivable by depicting the magnitude of geological time with a yearly planning calendar. Journal of Geo-science 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 Medicine, 73, 766–769.

Fetherstonhaugh, A., & Bezzi, A. (1992, August–September). Public knowledge and private understanding: Do they match? An example with the water cycle. Paper presented at the 29th International Geological Congress, Kyoto, Japan.

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., & Mayer, V. J. (Eds.). (1998, August). Learning about the Earth as a system. Proceedings of the 2nd International Conference on Geoscience Education. Columbus, OH: Earth Systems Education, the Ohio State University. ERIC Document ED422163.

Friedman, W. (1982). Conventional time concepts and children's structuring of time. In W. Friedman (Ed.), The developmental psychology of time. New York: Academic Press.

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.

Geosciences Professional Services. (Eds.). (2001). EarthScope: A new view into earth (project plan). Washington, DC: EarthScope. Report accessed July 30, 2004 at http://www.earthscope.org/links_pubs/index.html.

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 (p. 335). Boston: Houghton Mifflin.

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

Glasersfeld, E. von. (1987). The construction of knowledge—Contributions to conceptual semantics. Salinas, CA: Intersystems Publications.

Gobert, D. J. (2000). A typology of casual models for plate tectonics: Inferential power and barriers to understanding. International Journal of Science Education, 22, 937–977.

Gobert, J., & Clement, J. (1998). Effects of student-generated diagrams versus student-generated summaries on conceptual understanding of causal and dynamic knowledge in plate tectonics. Journal of Research in Science Teaching, 36, 39–53.

Gould, S. J. (1986). Evolution and the triumph of homology, or why history matters. American Scientist, 74, 60–89.

Gowin, D. B. (1981). Educating. Ithaca, New York: Cornell University Press.

Graczyk, S. L. (1993). Get with the system: General systems theory for business officials. School Business Affairs, 59, 16–20.

Greeno, J., Collins, A., & Resnick, L. (1996). Cognition and learning. In Handbook of Educational Psychology. New York: Macmillan.

Gudovitch, Y., & Orion, N. (2001, July). The carbon cycle and the Earth systems—Studying the carbon cycle in an environmental multidisciplinary context. Proceedings of the First IOSTE Symposium in Southern Europe, Paralimni, Cyprus.

Hanson, N. R. (1965). Patterns of discovery: An inquiry into the conceptual foundations of science. Cambridge, UK: Cambridge University Press.

Happs, J. C. (1982a). Some aspects of student understanding of two New Zealand landforms. New Zealand Science Teacher, 32, 4–12.

Happs, J. C. (1982b). 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). Learning about the Earth as a system. R. W. Fortner & V. J. Mayer (Eds.). Proceedings of the Second International Conference on Geoscience Education (1998, August). Columbus, OH: Earth Systems Education, the Ohio State University. ERIC Document ED422163.

IGEO. (2000, January). Conference proceedings of the 2nd meeting of the International Geosciences Education Organization, Sydney, Australia. I. Clark (Ed.).

IGEO. (2003, August). Conference proceedings of the 3rd meeting of the International Geosciences Education Organization, Calgary, Canada.

Kali, Y., & Orion, N. (1996). Relationship between Earth science education and spatial visualization. Journal of Research in Science Teaching, 33, 369–391.

Kali, Y., & Orion, N. (1997). Software for assisting high school students in the spatial perception of geological structures. Journal of Geoscience Education, 45, 10–21.

Kali, Y., & Orion, N. (2005). The effect of an Earth science learning program on students’ scientific thinking skills. Journal of Geoscience Education, 53, 387–393.

Kali, Y., Orion, N., & 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,40, 545–565.

Kapulnick, E., Orion, N., & Ganiel, U. (2004). In-service “science & Technology” training programs:

What changes do teachers undergo? Paper accepted for publication in the NARST annual meeting, Vancouver, 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.

Kitts, D. B. (1977). The structure of geology. Dallas: Southern Methodist University Press.

Krajcik, J., Blumenfeld, P. C., Marx, R. W., Kristin M. B., & Fredricks, J. (1998). Inquiry in project-based science classrooms: Initial attempts by middle school students. The Journal of the Learning Sciences (7), 313–350.

LeGrand, H. (1991). Drifting continents and shifting theories. Cambridge, UK: Cambridge University Press.

Leighton, J., & 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.

Lewis, J. P. (1998). Mastering project management: Applying advanced concepts of systems thinking, control and evaluation, resource allocation. New York: McGraw-Hill.

Libarkin, J. C., Anderson, S., Dahl, J., Beilfuss, M., & Boone, W. (2005). Qualitative analysis of college students’ ideas about earth: Interviews and open-ended questionnaires. Journal of Geo-science Education, 53, 17–26.

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.

Lovelock, J. (1991). Healing Gaia—Practical medicine for the planet (p. 192). New York: Harmony Books.

Marques, L., Paria, J., & Kempa, R. (2003). A study of students’ preconceptions of the organization and effectiveness of fieldwork in earth sciences education. Research in Science and Technological Education, 21, 265–278.

Marques, L. F., & Thompson, D. B. (1997a). Portuguese students’ understanding at age 10/11 and 14/15 of the origin and nature of the Earth and the development of life. Research in Science and Technology Education, 15, 29–51.

Marques, L. F., & Thompson D. (1997b). Misconceptions and conceptual changes concerning continental drift and plate tectonics among Portuguese students aged 16–17. Research in Science and Technological Education, 15(2), 195.

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. (Ed.). (2003). Implementing global science literacy. Columbus, OH: Ohio State University.

Mayer, V. J., & 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., & Fortner, R. W. (1995). Science is a study of earth. Columbus, OH: Earth Systems Education Program, the Ohio State University.

Mayer, V. J., & Fortner, R. W. (2002). A case history of science and science education policies. In V. J. Mayer (ed.), Global science literacy. Dordrecht, the Netherlands: Kluwer.

McAuliffe, C., Hall-Wallace, M., Piburn, M., Reynolds, S., & Leedy, D. E. (2000). Visualization and Earth science education. GSA Abstracts with Programs, 32, A-266.

McPhee, J. (1980). Basin and range. New York: Farrar, Straus & Giroux.

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.

Mintzes, J. J., Wandersee, J. H., & Novak, J. D. (1998). Research in science teaching and learning: a human constructivist view. In J. J. Mintzes, J. H. Wandersee, & J. D. Novak (Eds.), Teaching science for understanding: A human constructivist view. San Diego: Academic Press.

Montagnero, J. (1992). The development of the diachronic perspective in children. In F. Macar (Ed.), Time action and cognition (pp. 55–65). Amsterdam: Kluwer Academic.

Montagnero, J. (1996). Understanding changes in time. London: Taylor & Francis.

National Research Council. (1996). National science education standards. Washington, DC: National Academy Press.

Noonan-Pulling, L. C., & 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.

Nottis, K., & Ketter, K. (1999). Using analogies to teach plate tectonics concepts. Journal of Geo-science Education, 47(5), 449–454.

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. (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 and public awareness. AGID special publication series No 19. Rotterdam: Balkema.

Orion, N. (1998, July). Earth science education + environmental education = Earth systems education. In R. W. Fortner & V. J. Mayer (Eds.), Learning about the Earth as a system. Proceedings of the second International Conference on Geoscience Education (pp. 134–137). Columbus, OH: Earth Systems Education, the Ohio State University. ERIC Document ED422163.

Orion, N. (2002). An earth systems curriculum development model. In V. Mayer (Ed.), Global science literacy (pp. 159–168). Dordecht, the Netherlands: Kluwer.

Orion, N. (2003a). The outdoor as a central learning environment in the global science literacy framework: From theory to practice. In V. Mayer (Ed.), Implementing global science literacy (pp. 33–66). Columbus, OH: Ohio State University.

Orion, N. (2003b). Teaching global science literacy: a professional development or a professional change. In V. Mayer (Ed.), Implementing global science literacy (pp. 279–286). Columbus, OH: Ohio State University.

Orion, N., Ben-Chaim, D., & Kali, Y. (1997). Relationship between Earth science education and spatial visualization. Journal of Geoscience Education, 45, 129–132.

Orion, N., Dubowski, Y., & Dodick, J. (2000). The educational potential of multimedia authoring as a part of earth science curriculum—A case study. Journal of Research in Science Teaching, 37, 1121–1153.

Orion, N., & 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.

Orion, N., & 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., & Kali, Y. (2005). The effect of an Earth-science learning program on students’ scientific thinking skills. Journal of Geosciences Education, 53, 387–393.

Osborne, R., & 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 skill. Web site: http:www-sci.uni-klu.ac.at/~gossimit.

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. New York: Basic Books.

Pickering, K., & Owen, L. (1994). An introduction to global environmental issues. London and New York: Routledge.

Posner, G. J., Strike, K. A., Hewson, P. W., & Gertzog, W. (1982). Accomodation of a scientific conception: toward a theory of conceptual change. Science Education, 66(2), 211–227.

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, April). The hidden Earth: Visualization of geologic features and their subsurface geometry. Paper presented at the National Association for Research in Science Teaching annual meeting, New Orleans.

Riggs, E. M. (2003). Field-based education and indigenous knowledge: Essential components of geoscience education for native American communities. Science Education, 89, 296–313.

Riggs, E. M., & Semken, S. C. (2001). Earth science for Native Americans. Geotime, 49(9), 14–17.

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.

Ritger, S. D., & Cummins, R. H. (1991). Using student created metaphors to comprehend 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, & C. Cummins (Eds.), Proceedings of the 1992 evolution education research conference (pp. 214–229). Baton Rouge: Louisiana State University.

Ross, K., & 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 skills in authentic contexts. Journal of Research in Science Teaching, 30, 127–152.

Rowland, S. M. (1983). Fingernail growth and time-distance rates in geology. Journal of Geological Education, 31, 176–178.

Schoon, K. J. (1989, April). 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.

Schumm, S. A. (1991). To interpret the earth: Ten ways to be wrong. Cambridge, UK: Cambridge University Press.

Schwab, J. J. (1962). The teaching of science as inquiry. In J. J. Schwab & P. F. Brandweine (Eds.), The teaching of science. Cambridge, MA: Harvard University Press.

Senge, P. M. (1998). The practice of innovation. Leader to Leader, Summer.

Shymansky, J. A., & Yore, L. D. (1980). A study of teaching strategies, student cognitive development, and cognitive style as they relate to student achievement in science. Journal of Research in Science Teaching, 5, 369–382.

Slattery, W., Mayer, V., & 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.

Smith, E. L. (1991). Students’ conceptual frameworks: Consequences for learning science. In S. M. Glynn, R. H. Yeany, & B. K. Britton (Eds.), The psychology of learning science (pp. 65–85). Hillsdale, NJ: Lawrence Erlbaum Associates.

Solomon, J., Duveen, J., & Hall, S. (1994). What's happened to biology investigations? Journal of Biological Education, 28(4), 261.

Spencer-Cervato, C., & 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.

Stofflett, R. (1994). Conceptual change in elementary school teacher candidate knowledge of rock cycle processes. Journal of Geological Education, 42, 494–500.

Stow, D. A., & McCall, G. J. (Eds.). (1993, August). 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.

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 (p. 549). New York: Oxford University Press.

Tomorrow 98. (1993). Report of the superior committee on science, mathematics and technology education in Israel (p. 115). Jerusalem: Ministry of Education of Israel.

Toulmin, S. E. (1972). Human understanding. Princeton, NJ: Princeton University Press.

Trend, R. D. (1997, July). An investigation into understanding of geological time among 10 and 11-year old children, with a discussion of implications for learning of other geological concepts. Paper presented at 1st International Conference on Geoscience Education and Training, Hilo, Hawaii.

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 primary teachers’ conceptions of geological time and perceptions of geoscience. Journal of Research in Science Teaching, 38(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 (pp. 187–202). London: Kluwer Academic.

Wandersee, J. H., Mintzes, J. J., & Novak, J. D. (1994). Research on alternative conceptions in science. In D. L. Gabel (Ed.), Handbook of research on science teaching and learning. New York: Macmillan.

Welch, W. W., Klopfer, L. E., Aikenhead, G. S., & 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 of time. Amsterdam: North-Holland.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.142.136.226