Living in the Environment: Principles, Connections, and Solutions / G. Tyler Miller, Jr. SYSNO 6379018, přírůstkové číslo 6543 Seznam obrázků v dokumentu: Environmental science: Concepts and Connections Developed by Jane Heinze-Fry with assistance from G. Tyler Miller, Jr. (For assistance in creating your own concept maps, see the webside for this book.) Schema: Reactactants: carbon (C, black solid) + oxygen (O2, coloress gas) -> Product(s): carbon dioxide (CO2, coloress gas) + energy Figure_1–01.jpg The J-shaped curve of past exponential world population growth, with projections beyond 2100. Figure_1–02.jpg Linear and exponential growth. If you save $1,000 a year for a lifetime of 70 years, the resulting linear growth will allow you to save $70,000 (lower curve). Figure_1–03.jpg World population milestones. (Data from United Nations Population Division, World Population Prospects, 1998) Figure_1–04.jpg Human disturbance of the earth’s land area. Figure_1–05.jpg Degree of economic development as measured by per capita GNP in 1998. (Data from United Nations and the World Bank) Figure_1–06.jpg Past and projected population size for developed countries, developing countries, and the world, 1950–2120. More than 95% of the addition of 3.6 billion people between 1990 and 2030 is projected to occur in developing countries. (Data from United Nations) Figure_1–07.jpg Growth of the gross world product, 1900–2000. (Data from United Nations and the World Bank) Figure_1–08.jpg Types of decision making in traditional and sustainable societies. The traditional decision making in most societies involves treating social, economic, and environmental issues separately (left). Environmentally sustainable economic development calls for integrating social, economic, and environmental issues and concepts to find sustainable solutions to problems (right). Figure_1–09.jpg One in every three children under age 5, such as this Brazilian child, suffers from malnutrition. Figure_1–10.jpg Relative ecological footprints of the United States, the Netherlands, and India. Figure_1–11.jpg Major types of material resources. Figure_1–12.jpg Full production and exhaustion cycle of a nonrenewable resource such as copper, iron, oil, or coal. Usually, a nonrenewable resource is considered economically depleted when 80 % of its total supply has been extracted and used. Normally, it costs too much to extract and process the remaining 20 % Figure_1–13.jpg Major environmental and resource problems. Figure_1–14.jpg Environmentalists have identified five root causes of the environmental problems we face. Figure_1–15.jpg Simplified model of how three factors – population, affluence, and technology – affect the environmental impact of population in developing countries (top) and developed countries (bottom). Figure_1–16.jpg Major components and interactions within and between the earth’s life-support system and the human sociocultural system (culture sphere). The goal of environmental science is to learn as much as possible about these complex interactions. Figure_2–01.jpg The historical range of the bison shrank severely between 1500 and 1906, mostly because of unregulated and deliberate over hunting. Figure_2–02.jpg Technological innovations have led to greater human control over the rest of nature and an expanding human population. Dashed lines represent three alternative population futures: (1) continued growth (top), (2) stabilization (middle), and (3) a crash and stabilization at a much lower level. Figure_2–03.jpg The first crop-growing technique may have been a combination of slash-and-burn and shifting cultivation in tropical forests. Figure_2–04.jpg Henry David Thoreau (1817–1862) was an American writer and naturalist who kept journals about his excursions into wild nature throughout parts of the northeastern United States and Canada and at Walden Pond in Massachusetts. He sought self-sufficiency, a simple lifestyle, and a harmonious coexistence with nature.& Figure_2–05.jpg Examples of the increased role of the federal government in resource conservation and public health and establishment of key private environmental groups, 1870–1930. Figure_2–06.jpg John Muir (1838–1914) was a geologist, explorer, and naturalist. He spent 6 years studying, writing journals, and making sketches in the wilderness of California’s Yosemite Valley and then went on to explore wilderness areas in Utah, Nevada, the Northwest, and Alaska. He was largely responsible for the establishment of Yosemite National Park in 1890. He also founded the Sierra Club and spent 22 years lobbying actively for conservation laws. Figure_2–07.jpg Theodore („Teddy“) Roosevelt (1858–1919) was a writer, explorer, naturalist, avid birdwatcher, and 26th president of the United States. He was the first national political figure to bring the issues of conservation to the attention of the American public. According to many historians, he has contributed more than any other president to the natural resource conservation in the United States. Figure_2–08.jpg Alice Hamilton (1869–1970) was the first and foremost expert on industrial disease in the United States. (Schliesinger Library, Radcliffe College) Figure_2–09.jpg Some important conservation and environmental events between 1930 and 1960 Figure_2–10.jpg Some important environmental events during the 1960s. Figure_2–11.jpg Biologist Rachel Carson (1907–64) was a pioneer in increasing public awareness of the importance of nature and the threat of pollution. She died without knowing that her efforts were a key in beginning the modern era of environmentalism in the United States. (©1962 Eric Hartmann/Magnum Photos) Figure_2–12.jpg Some important environmental events during the 1970s, sometimes called the environmental decade. Figure_2–13.jpg Some important environmental events during the 1980s. Figure_2–14.jpg Some important environmental events during the 1990s. Figure_2–15.jpg Aldo Leopold (1887–1948) was a forester, writer, and conservationist. His book, A Sand County Almanac (published after his death) is considered an environmental classic that has inspired the modern environmental movement. His land ethic expanded the role of humans as protectors of nature. (Robert McCabe, University of Wisconsin-Madison Archives) Figure_3–01.jpg These massive stone figures on Easter Island are the remains of the technology created by an ancient civilization of Polynesians. Figure_3–02.jpg What scientists do. In the scientific process, a form of critical thinking, (1) facts (data) are gathered and verified by repeated experiments, (2) data are analyzed to see whether there is a consistent pattern of behavior that can be summarized as a scientific law, (3) hypotheses are proposed to explain the data, and (4) deductions or predictions are made and tested to evaluate each hypothesis. Figure_3–04.jpg Coupled negative and positive feedback loops involved in temperature control of the human body. Homeostasis works in a limited range only: Above a certain body temperature, body metabolic rates get out of control and generate large amounts of heat. This positive (runaway) feedback loop generates more heat than the negative feedback loop can get rid of, and body temperature increases out of control, resulting in death. Figure_3–05.jpg Comparison of the solid, liquid, and gaseous physical states of matter. Figure_3–06.jpg Isotopes of hydrogen and uranium. Figure_3–07.jpg The pH scale, used to measure acidity and alkalinity of water solutions. Values shown are approximate. A solution with a pH less than 7 is acidic, a neutral solution has a pH of and one with a pH greater than 7 is basic. Each whole-number drop in pH represents a 10-fold increase in acidity. (From Biology: Concepts and Applications , 4th ed., by Cecie Starr, ©2000. Reprinted with permission of Brooks/Cole, a division of Thomson Learning. Fax 800 730–2215.) Figure_3–08.jpg Relationships between cells, nuklei, chromosomes, DNA, and genes. Figure_3–09.jpg Examples of differences in matter quality. High-quality matter (left-hand column) is fairly easy to extract and is concentrated; low-quality matter (right-hand column) is more difficult to extract and is more dispersed than high-quality matter. Figure_3–10.jpg The electromagnetic spectrum: the range of electromagnetic waves, which differ in wavelength (distance between successive peaks or troughs) and energy content. Cosmic rays, gamma rays, X rays, and ultraviolet radiation are called ionizing radiation because they have enough kinetic energy to knock electrons from atoms and change them to positively charged ions. The resulting highly reactive electrons and ions can (1) disrupt living cells, (2) interfere with body processes, and (3) cause many types of sickness, including various cancers. The other forms of electromagnetic radiation (right side) do not contain enough kinetic energy to form ions and are called nonionizing radiation. Figure_3–11.jpg Categories of the quality (usefulness for performing various energy tasks) of different sources of energy. High-quality energy is concentrated and has great ability to perform useful work; low-quality energy is dispersed and has little ability to do useful work. To avoid unnecessary energy waste, it is best to match the quality of an energy source with the quality of energy needed to perform a task. Figure_3–12.jpg The three principal types of ionizing radiation emitted by radioactive isotopes differ greatly in their penetrating power. Figure_3–13.jpg The radioactive decay of plutonium-239, which is produced in nuclear reactors and used as the explosive in some nuclear weapons, has a half-life of 240,000 years. Figure 3–14.jpg Natural and human sources of the average annual dosage of ionizing radiation received by people in the United States. Most studies indicate that there is no safe dosage of ionizing radiation. (Data from National Council on Radiation Protection and Measurements). Figure_3–15.jpg Fission of a uranium-235 nucleus by a neutron (n). Figure_3–16.jpg A nuclear chain reaction initiated by one neutron triggering fission in a single uranium-235 nucleus. This figure illustrates only a few of the trillions of fissions caused when a single uranium-235 nucleus is split within a critical mass of uranium-235 nuclei. The elements krypton (Kr) and barium (Ba), shown here as fission fragments, are only two of many possibilities. Figure_3–17.jpg The deuterium-tritium (D-T) and deuterium-deuterium (D-D) nuclear fusion reactions, which take place at extremely high temperatures. Figure_3–18.jpg The second energy law in action in living systems. Each time energy is changed from one form to another, some of the initial input of high-quality energy is degraded, usually to low-quality heat that disperses into the environment. Figure_3–19.jpg The high-throughput (high-waste) economies of most developed countries are based on maximizing the rates of energy and matter flow. This process rapidly converts high-quality matter and energy resources into waste, pollution, and low-quality heat. Figure_3–20.jpg Lessons from nature. A low-throughput (low-waste) economy, based on energy flow and matter recycling, works with nature to reduce throughput. This is done by (1) reusing and recycling most nonrenewable matter resources, (2) using renewable resources no faster than they are replenished, (3) using matter and energy resources efficiently, (4) reducing unnecessary consumption, (5) emphasizing pollution prevention and waste reduction, and (6) controlling population growth. Figure_4–1.jpg Insects play important roles in helping sustain life on earth. The bright green caterpillar moth feeding on pollen in a crocus (right) and other insects pollinate flowering plants that serve as food for many plant eaters. The praying mantis eating a monarch butterfly (left) and many other insect species help control the populations of at least half of the insect species we classify as pests. (Left: Pat Andersen, Visuals Unlimited; right: Stephen Hopkin/Planet Earth Pictures, Ltd.) Figure_4–2.jpg Model of levels of organization of matter in nature. Note that ecology focuses on five levels of this hierarchical model. Figure_4–3.jpg (a) Generalized structure of a eukaryotic cell. Figure_4–4.jpg A population of monarch butterflies wintering in Michoacan, Mexico. The geographic distribution of this butterfly coincides with that of the milkweed plant, on which monarch larvae and caterpillars feed. (Frans Lanting/Bruce Coleman Collection.) Figure_4–5.jpg The genetic diversity among individuals of one species of Caribbean snail is reflected in the variations in shell color and banding patterns. (Alan Solem) Figure_4–6.jpg The general structure of the earth. The atmosphere consists of several layers, including the tropo-sphere (innermost layer) and the strato-sphere (second layer). Figure_4–7.jpg Life on the earth depends on (1) the one-way flow of energy (dashed lines) from the sun through the biosphere, (2) the cycling of crucial elements (solid lines around circles), and (3) gravity, which keeps atmospheric gases from escaping into space and draws chemicals downward in the matter cycles. This simplified model depicts only a few of the many cycling elements. Figure_4–8.jpg The flow of energy to and from the earth. The ultimate source of energy in most ecosystems is sunlight. Figure_4–9.jpg Major biomes found along the 39th parallel across the United States. The differences reflect changes in climate, mainly differences in average annual precipitation and temperature (not shown). Figure 4–10 Ecosystems rarely have sharp boundaries. Two adjacent ecosystems such as dry land and an open lake often contain a marsh – an ecotone or transitional zone – between them. This zone contains a mixture of species found in each ecosystem and contains some species not found in either ecosystem. Figure_4–11.jpg Major components of a freshwater ecosystem. Figure_4–12.jpg Major components of an ecosystem in a field. Figure_4–13.jpg Key physical and chemical or abiotic factors affecting terrestrial ecosystems (left) and aquatic life zones (right). Figure_4–14.jpg Range of tolerance for a population of organisms to an abiotic environmental factor-in this case, temperature. Figure_4–15.jpg Some detritivores, called detritus feeders, directly consume tiny fragments of this log. Other detritivores, called decomposers (mostly fungi and bacteria), digest complex organic chemicals in fragments of the log into simpler inorganic nutrients. These nutrients can be used again by producers if they are not washed away or otherwise removed from the system. Figure_4–16.jpg Model showing how an ecosystem’s main structural components (energy, chemicals, and organisms) are linked by matter recycling and the flow of energy from the sun, through organisms, and back to the environment as low-quality heat. Each type of organism in an ecosystem plays a unique role in the processes of energy flow and matter cycling. Figure_4–17.jpg Two species found in tropical forests are part of the earth’s biodiversity. Figure_4–18.jpg Model of a food chain. The arrows show how chemical energy in food flows through various trophic levels or energy transfers; most of the energy is degraded to heat, in accordance with the second law of energy. Food chains rarely have more than four trophic levels. Can you explain why? Figure_4–19.jpg Model of a greatly simplified food web in the Antarctic. Many more participants in the web, including an array of decomposer organisms, are not depicted here. Figure_4–20.jpg Generalized pyramid of energy flow showing the decrease in usable energy available at each succeeding trophic level in a food chain or web. Figure_4–21.jpg Annual pyramid of energy flow (in kilocalories per square meter per year) for an aquatic ecosystem in Silver Springs, Florida. The pyramid is constructed by using the data on energy flow through this ecosystem shown in the top drawing. Figure_4–22.jpg Generalized graphs of biomass of organisms in the various trophic levels for two ecosystems: The size of each tier in this conceptual model represents the dry weight per square meter of all organisms at that trophic level. Figure_4–23.jpg Generalized graphs of numbers of organisms in the various trophic levels for two ecosystems. Figure_4–24.jpg Three years of satellite data on the earth’s gross primary productivity. Figure_4–25.jpg Estimated annual average net primary productivity (NPP) per unit of area in major life zones and ecosystems, expressed as kilocalories of energy produced per square meter per year (kcal/m2/yr). (Data from Communities and Ecosystems, 2/E by R. H. Whittaker, 1975, New York: Macmillan, ) Figure_4–26.jpg Human use of the biomass produced by photosynthesis. Humans destroy, alter, and directly use about 27% of the earth’s total net primary productivity and about 40% of the net primary productivity of the earth’s terrestrial ecosystems. (Data from Peter Vitousek) Figure_4–27.jpg Major nonliving and living storehouses of elemental nutrients. Figure_4–28.jpg Simplified model of the hydrologic cycle. Figure_4–29.jpg Simplified model of the global carbon cycle. Figure_4–30.jpg Greatly simplified model of the nitrogen cycle in a terrestrial ecosystem. Nitrogen reservoirs are shown as boxes and processes changing one form of nitrogen to another are shown in unboxed print. (From Biology: The Unity and Diversity of Life, 9/e by Starr & Taggart ©2001) Figure_4–31.jpg Loss of nitrate ions (NO3-) from a deforested watershed in the Hubbard Brook Experimental Forest in New Hampshire. Figure_4–32.jpg Simplified model of the phosphorus cycle. Figure_4–33.jpg Simplified model of the sulfur cycle. Green shows the movement of sulfur compounds in living organisms, blue in aquatic systems, and orange in the atmosphere. Figure_4–34.jpg Geographic information systems (GISs) provide the computer technology for organizing, storing, and analyzing complex data collected over broad geographical areas. GISs enable scientist to overlay many layers of data (such as soils, topography, distribution of endangered populations, and land protection status). Figure_4–35.jpg Major stages of systems analysis. (Modified data from Charles Southwick) Figure_4–36.jpg Ecosystem services. Energy from the sun (solar capital) and natural resources (natural capital) provide ecological services that support and sustain all life and all economies on the earth. Figure_5–01.jpg The earth is a blue and white planet in the black void of space Currently, it has the right physical and chemical conditions to allow the development of life as we know it today. (NASA). Figure_5–02.jpg Summary of the chemical and biological evolution of the earth. This drawing is not to scale. Note that the time span for biological evolution is almost four times longer than that for chemical evolution. Figure_5–03.jpg Synthesis of organic compounds necessary for life from the gases believed to be in the earth’s primitive atmosphere. Figure_5–04.jpg Greatly simplified overview of the biological evolution of life on the earth, which was preceded by about 0.5–1 billion years of chemical evolution. Figure_5–05.jpg Two varieties of peppered moths found in England illustrate one kind of adaptation: camouflage. Figure_5–06.jpg Three ways in which natural selection can occur, using the trait of coloration in a population of snails. Figure_5–07.jpg Overlap of the niches of two different species: a specialist and a generalise In the overlap area the two species compete for one or more of the same resources. Figure_5–08.jpg How geographic isolation can lead to reproductive isolation, divergence, and speciation. Figure_5–09.jpg Continental drift, the extremely slow movement of continents over millions of years on several gigantic plates (discussed in more detail in Section 10–2, p. 212). Figure_5–10.jpg Over millions to hundreds of millions of years, macroevolution has consisted of dramatic exits (extinctions) and entrances (speciation and radiations) of large groups of species. Figure_5–11.jpg Adaptive radiation of mammals began in the first 10–12 million years of the Cenozoic era (which began about 65 million years ago) and continues today. Figure_6–1.jpg Some of the dust shown here blowing from Africa’s Sahara Desert can end up as soil nutrients in Amazonian rain forests. (NOAA.USGS/MND EROS Data Center) Figure_6–2.jpg Areas of North America most susceptible to tornadoes and hurricanes. (Data from NOAA and U.S. Geological Survey) Figure_6–3.jpg Climate and its effects. (Data from National Oceanic and Atmospheric Administration) Figure_6–4.jpg Generalized map of global climate zones, showing the major contributing ocean currents and drifts. Large variations in climate are dictated mainly by temperature (with its seasonal variations) and by the quantity and distribution of precipitation. Figure_6–5.jpg The effects of the earth’s tilted axis on climate. As the planet makes its annual revolution around the sun on an axis tilted about 23.5°, various regions are tipped toward or away from the sun. The resulting variations in solar energy reaching the earth create the seasons. Figure_6–6.jpg Formation of prevailing surface winds, which disrupt the general flow of air from the equator to the poles and back to the equator. Figure_6–7.jpg Transfer of energy by convection in the atmosphere. Figure_6–8.jpg Model of global air circulation and biomes. Figure_6–9.jpg A shore upwelling (shown here) occurs when deep, cool, nutrient-rich waters are drawn up to replace surface water moved away from a steep coast by wind-driven currents. Figure_6–10.jpg Normal surface winds blowing westward cause shore upwellings of cold, nutrient-rich bottom water in the tropical Pacific Ocean near the coast of Peru (left). Figure_6–11.jpg Typical global climatic effects of an El Niño -Southern Oscillation. During the 1996–98 ENSO, huge waves battered the California coast, and torrential rains caused widespread flooding and mudslides. Figure_6–12.jpg El Niño and La Niña conditions between 1950 and 1999. (Data from U.S. National Weather Service) Figure_6–13.jpg The greenhouse effect. Without the atmospheric warming provided by this natural effect, the earth would be a cold and mostly lifeless planet. According to the widely accepted greenhouse theory, when concentrations of greenhouse gases in the atmosphere rise, the average temperature of the troposphere also rises. (From Biology: Concepts and Principles, 4th ed. by Cecie Starr © 2000) Figure_6–14.jpg The rain shadow effect is a reduction of rainfall on the side of high mountains facing away from prevailing surface winds. Figure_6–15.jpg Sea and land breezes. Figure_6–16.jpg The earth’s major biomes – the main types of natural vegetation in different undisturbed land areas-result primarily from differences in climate. Figure_6–17.jpg Average precipitation and average temperature, acting together as limiting factors over a period of 30 or more years, determine the type of desert, grassland, or forest biome in a particular area. Figure_6–18.jpg Generalized effects of latitude and altitude on moisture and biomes in North America. Parallel changes in vegetation type occur when we travel from the equator to the poles or from lowlands to mountain-tops. Plant types in these areas also vary with mean annual air temperature and soil types. (From The Unity and Diversity of Life, 4th ed. by Cecie Starr and Ralph Taggart © 2001) Figure_6–19.jpg Climate graphs showing typical variations in annual temperature and precipitation in tropical, temperate, and polar (cold) deserts. Figure_6–20.jpg Some components and interactions in a temperate desert biome. Figure_6–21.jpg Climate graphs showing typical variations in annual temperature and precipitation in tropical, temperate, and polar (arctic tundra) grasslands. Figure_6–22.jpg Some of the grazing animals found in different parts of the African savanna. Each species has a different feeding niche that allows them to share vegetation resources. Figure_6–23.jpg Some components and interactions in a temperate tall-grass prairie ecosystem in North America. Figure_6–24.jpg Some components and interactions in an arctic tundra (polar grassland) ecosystem. Figure_6–25.jpg Replacement of a temperate grassland with a monoculture crop near Blythe, California. Figure_6–26.jpg Climate graphs showing typical variations in annual temperature and precipitation in tropical, temperate, and polar forests. Figure_6–27.jpg Some components and interactions in a tropical rain forest ecosystem. Figure_6–28.jpg Stratification of specialized plant and animal niches in various layers of a tropical rain forest. Figure_6–29.jpg Some components and interactions in a temperate deciduous forest ecosystem. Figure_6–30.jpg Some components and interactions in an evergreen coniferous (boreal or taiga) forest ecosystem. Figure_6–31.jpg Tree farm in North Carolina. Some diverse virgin (old-growth) or second-growth forests are cleared and replanted with a single tree species (monoculture), often for harvest as Christmas trees, timber, or wood converted to pulp to make paper. (Gene Alexander/USDA)& Figure_7–01.jpg A healthy coral reef in the Philippines covered by colorful algae (bottom) and a bleached coral reef in the Bahamas that has lost most of its algae (top) because of changes in the environment (such as cloudy water or too warm temperatures). Figure_7–02.jpg Distribution of the world’s major saltwater oceans, coral reefs, mangroves, and freshwater lakes and rivers. Figure_7–03.jpg Variations in concentrations of dissolved oxygen (O2) and carbon dioxide (CO2) in parts per million (ppm) with water depth. Figure_7–04.jpg The ocean planet. Figure_7–05.jpg Major life zones, in an ocean. Ecologists classify ocean habitats and their organisms on the basis of light levels, depth, and bottom type. (Not drawn to scale Actual depths of zones may vary) Figure_7–06.jpg Mangroves in Columbia, South America. Mangroves (Figure 7–2) are important coastal systems that protect coastlines and provide important habitats for aquatic species. Figure_7–07.jpg View of an estuary taken from space. Figure_7–08.jpg Salt marsh of an estuary in a temperate area consists of several connected coastal life zones. (From Biology: Concepts and Applications, 4th ed. by Cecie Starr © 2000) Figure_7–09.jpg Some components and interactions in a salt marsh ecosystem in a temperate areas such as the United States. Figure_7–10.jpg Living between the tides. Some organisms with specialized niches found in various zones on rocky shore beaches (top) and barrier or sandy beaches (bottom). Organisms are not drawn to scale. Figure_7–11.jpg Primary and secondary dunes on gently sloping sandy beaches play an important role in protecting the land from erosion by the sea. Figure_7–12.jpg A developed barrier island: Ocean City, Maryland, host to 8 million visitors a year. Figure_7–13.jpg Some components and interactions in a coral reef ecosystem. Figure_7–14.jpg The distinct zones of life in a fairly deep temperate zone lake. Figure_7–15.jpg An oligotrophic, or nutrient-poor, lake (top) and a eutrophic, or nutrient-rich, lake (bottom). Figure_7–16.jpg During the summer and winter, the water in deep temperate zone lakes becomes stratified into different temperature layers, which do not mix. Figure_7–17.jpg The three zones in the downhill flow of water: (1) source zone containing mountain (headwater) streams, (2) transition zone containing wider, lower-elevation streams, and (3) floodplain zone containing rivers, which empty into the ocean. Figure_7–18.jpg Some components and interactions in a river in a tropical forest. Figure_7–19.jpg Inland wetlands are not always wet. The graphs show the fluctuating water levels of three types of inland wetlands. (Data from Jon A. Kusler, William J. Mitsch, and Joseph S. Larson, 1994) Figure_8–1.jpg Flying foxes (top right) are bats that play key ecological roles in tropical rain forests in Southeast Asia by pollinating (bottom left) and spreading the seeds of durian trees that produce durians, a highly prized tropical fruit, (bottom right). (Top and bottom right, Dr. Merlin Tuttle/Photo Researchers, Inc.; Bottom left, Christer Friedriksson/Bruce Coleman Collection) Figure_8–2.jpg Generalized types, relative sizes, and stratification of plant species in various terrestrial communities or ecosystems. Figure_8–3.jpg Changes in species diversity at different latitudes (distances from the equator) in terrestrial communities for (a) ants and (b) breeding birds of North and Central America. Figure_8–4.jpg Changes in species diversity with depth in marine environments for (a) snails and (b) tube worms. As a general rule, species diversity in the ocean increases from the surface to a depth of about 2,000 meters and then decreases until the sea bottom where species diversity is usually high. Figure_8–5.jpg Changes in the species diversity and species abundance of diatom species in an unpolluted stream and a polluted stream. Note that both species diversity and species abundance decrease with pollution. Figure_8–6.jpg The species equilibrium model or theory of island biogeography, developed by Robert MacArthur and Edward 0. Wilson, (a) The equilibrium number of species (blue triangle) on an island is determined by a balance between the immigration rate of new species and the extinction rate of species already on the island. Figure_8–7.jpg Research data supporting the theory of island biogeography with (a) the right graph showing that species diversity increases with island size and (b) the left graph showing that the species diversity of birds occupying lowland areas of South Pacific islands decreases with distance from New Guinea. Figure_8–8.jpg The results of G. F. Gause’s classic laboratory experiment with two similar single-celled, bacteria-eating Paramecium species (which reproduce asexually) support the competitive exclusion principle that similar species cannot occupy the same ecological niche indefinitely. Figure_8–9.jpg Specialized feeding niches of various bird species in a coastal wetland. Such resource partitioning reduces competition and allows sharing of limited resources. Figure_8–10.jpg Resource partitioning and niche specialization as a result of competition between two species. The left diagram shows the overlapping niches of two competing species. The right diagram shows that through evolution the niches of the two species become separated and more specialized (narrower) so that they avoid competing for the same resources. Figure_8–11.jpg Resource partitioning of five species of common insect-eating warblers in spruce forests of Maine. Each species minimizes competition with the others for food by spending at least half its feeding time in a distinct portion (shaded areas) of the spruce trees; each also consumes somewhat different insect species. (After „Population Ecology of Some Warblers in Northeastern Coniferous Forests,“ by R. H. MacArthur, 1958, Ecology, Vol. 36, 533–36) Figure_8–12.jpg Some ways in which prey species avoid their predators by (1) camouflage (a and b), (2) chemical warfare (c and e), (3) warning coloration (d and e), (4) mimicry (f), (5) deceptive looks (g), and (6) deceptive behavior (h). Figure_8–13.jpg Two examples of mutualism, (a) Oxpeckers (or tickbirds) feed on the parasitic ticks that infest large, thick-skinned animals such as a black rhinoceros, and (b) a clownfish lives among deadly stinging sea anemones.& Figure_8–14.jpg Commensalism between a white orchid (an epiphyte or air plant from the tropical forests of Latin America) that roots in the fork of a tree rather than the soil without penetrating or harming the tree. In this interaction, the epiphytes gain access to water, nutrient debris, and sunlight; the tree apparently remains unharmed unless it contains a large number of epiphytes.& Figure_8–15.jpg Starting from ground zero. Primary ecological succession over several hundred years of plant communities on bare rock exposed by a retreating glacier on Isle Royal in northern Lake Superior. Figure_8–16.jpg Natural restoration of disturbed land. Secondary ecological succession of plant communities on an abandoned farm field in North Carolina. Figure_8–17.jpg Examples of wildlife species typically found at different stages of ecological succession in areas of the United States with a temperate climate. Figure_8–18.jpg According to the intermediate disturbance hypothesis, moderate disturbances in communities promote greater species diversity than small or major disturbances. Figure_8–19.jpg Examples of how some of the earth’s natural resources are being depleted and degraded at an accelerating rate as a result of the exponential growth of the human population and resource use by humankind. Figure_9–01.jpg (a) A southern sea otter. (Jeff Foott Productions/Bruce Coleman Collection) (b) A sea urchin. (Jane Burton/Bruce Coleman Collection) Figure_9–02.jpg Generalized dispersion patterns for individuals in a population throughout their habitat. The most common pattern is one in which members of a population exist in clumps throughout their habitat (left), mostly because resources usually are found in patches. Figure_9–03.jpg Factors that tend to increase or decrease the size of a population. Whether the size of a population grows, remains stable, or decreases depends on infractions between its growth factors (biotic potential) and decrease factors (environmental resistance). Figure_9–04.jpg Theoretical population growth curves, (a) Exponential growth, in which the population’s growth rate increases with time. Figure_9–05.jpg Logistic growth of a sheep population on the island of Tasmania between 1800 and 1925. Figure_9–06.jpg Exponential growth, overshoot, and population crash of reindeer introduced to a small island off the southwest coast of Alaska. Figure_9–07.jpg General types of simplified population change curves found in nature, (a) The population size of a species with a fairly stable population fluctuates slightly above and below its carrying capacity, (b) Figure_9–08.jpg Population cycles for the snowshoe hare and Canadian lynx. Figure_9–09.jpg Changes in moose and wolf populations on Isle Royale from 1900 to 1999. (Data from Rolf O. Peterson, 1995) Figure_9–10.jpg Generalized characteristics of r-selected or opportunist species and K-selected or competitor species. Many species have characteristics between these two extremes. Figure_9–11.jpg Three general survivorship curves for populations of different species, obtained by showing the percentages of the members of a population surviving at different ages. Figure_9–12.jpg Some effects of environmental stress on organisms, populations, communities, and ecosystems. Figure_10–1.jpg The San Andreas Fault as it crosses part of the Carrizo Plain between San Francisco and Los Angeles, California. Figure_10–2.jpg The earth’s internal zones. (Surface features are not to scale.) Figure_10–3.jpg Major features of the earth’s crust and upper mantle. The lithosphere, composed of the crust and outermost mantle, is rigid and brittle. The asthenosphere, a zone in the mantle, can be deformed by heat and pressure. Figure_10–4.jpg Composition by weight of the earth’s crust. Various combinations of only eight elements make up the bulk of most minerals. Figure_10–5.jpg Earthquake and volcano sites are distributed mostly in bands along the planet’s surface, (a) These bands correspond to the patterns for the types of lithospheric plate boundaries (b) shown in Figure 10–6. Figure_10–6.jpg Types of boundaries between the earth’s lithospheric plates. All three boundary types occur both in oceans and on continents. Figure_10–7.jpg The variety of land-forms and sedimentary environments depicted here result mainly from external processes. They are powered primarily by solar energy (as it drives the hydrologic cycle and wind) and gravity, with some assistance from organisms such as reef-building corals Figure_10–8.jpg The rock cycle, the slowest of the earth’s cyclic processes. Figure_10–9.jpg Major features and effects of an earthquake. Figure_10–10.jpg Expected damage from earthquakes in Canada and the contiguous United States. Figure_10–11.jpg A volcano erupts when molten magma in the partially molten asthenosphere rises in a plume through the lithosphere to erupt on the surface as lava that can spill over or be ejected into the atmosphere. Chains of islands can be created by the action of volcanoes that then become inactive. Figure_10–12.jpg Soil formation and generalized soil profile. Figure 10–13.jpg Greatly simplified food web of living organisms found in soil. Figure 10–14.jpg Pathways of plant nutrients in soils. Figure 10–15.jpg Soil profiles of the principal soil types typically found in five different biomes. Figure 10–16.jpg Soil texture depends on the proportions of clay, silt, and sand particles in the texture affects soil porosity, the average nu i and spacing of pores in a given volume of Loams – roughly equal mixtures of clay, sand, silt, and humus – are the best soils for grow most crops. (Data from Natural Resources Conservation Service) Figure 10–17.jpg The size, shape, and degree of clumping of soil particles determine the number and volume of spaces for air and water within a soil. Soils with more pore spaces (left) contain more air and are more permeable to water flows than soils with fewer pores (right). Figure_10–18.jpg Rill and gully erosion of vital topsoil from irrigated cropland in Arizona. (Natural Resources Conservation Service) Figure_10–19.jpg Global soil erosion. (Data from UN Environment Programme and the World Resources Institute) Figure_10–20.jpg The Dust Bowl of the Great Plains, where a combination of extreme drought and poor soil conservation practices led to severe wind erosion of topsoil in the 1930s. Figure_10–21.jpg Desertification of arid and semiarid lands. (Data from UN Environmental Programme and Harold E. Drengue) Figure 10–22.jpg Salinization and waterlogging of soil on irrigated land without adequate drainage lead to decreased crop yields. Figure_10–23.jpg Severe salinization. Because of high evaporation, poor drainage, and severe salinization, white alkaline slats have displaced crops that once grew on this heavily irrigated land in Colorado. (Natural Resources Conservation Service) Figure_10–24.jpg Soil conservation methods: (a) terracing, (b) contour planting and strip cropping, (c) alley cropping, and (d) windbreaks. Figure_11–01.jpg This policeman in Bangkok, Thailand, is wearing a mask to reduce his intake of air polluted mainly by automobiles. Figure_11–02.jpg Average crude birth and death rates for various groupings of countries in 2000. (Data from Population Reference Bureau) Figure_11–03.jpg Average annual rate of population change (natural increase) in 2000. (Data from Population Reference Bureau) Figure_11–04.jpg Average annual increase in the world’s population, 1950–2000, and projected increase 2000–2050 (dotted line). (Data from United Nations) Figure_11–05.jpg The world’s 10 most populous countries in 2000, with projections of their population size in 2025. In 2000, more people lived in China than in all of Europe, Russia, North America, Japan, and Australia combined. (Data from World Bank and Population Reference Bureau) Figure_11–06.jpg Population projections by region, 2000–2025. (Data from United Nations and Population Reference Burea). Figure_11–07.jpg Decline in total fertility rates for various groupings of countries, 1950–2000. (Data from United Nations) Figure_11–08.jpg Total fertility rates in 2000. (Data from Population Reference Bureau) Figure_11–09.jpg United Nations world population projections to 2050, assuming that the world’s total fertility rate is 2.5 (high), 2.0 (medium), or 1.6 (low) children per woman. (Data from United Nations) Figure_11–10.jpg Total fertility rates for the United States between 1917 and 2000. (Data from Population Reference Bureau and U.S. Census Bureau) Figure_11–11.jpg Birth rates in the United States from 1910 to 2000. (Data from U.S. Bureau of Census and U.S. Commerce Department) Figure_11–12.jpg Typical effectiveness of birth control methods in the United States. Figure_11–13.jpg Changes in crude birth and death rates for developed and developing countries between 1775 and 2000. (Data from Population Reference Bureau and United Nations) Figure_11–14.jpg Infant mortality rates in 2000. (Data from Population Reference Bureau) Figure_11–15.jpg Infant mortality rates in the United States, 1917–2000. Figure_11–16.jpg Generalized population age structure diagrams for countries with rapid (1.5–3%), slow (0.3–1.4%), zero (0–0.2%), and negative population growth rates. (Data from Population Reference Bureau) Figure_11–17.jpg Population structure by age and sex in developing countries and developed countries, 2000. In 2000, there were 1 billion young people in their prime reproductive years of 15–24 and 1.9 billion people under age 15, moving into their reproductive years. (Data from United Nations Population Division and Population Reference Bureau) Figure_11–18.jpg Comparison of key demographic indicators in a highly developed (United States), moderately developed (Brazil), and less developed country (Nigeria) in 2000. (Data from Population Reference Bureau) Figure_11–19.jpg Tracking the baby-boom generation in the United States. (Data from Population Reference Bureau and U.S. Census Bureau) Figure_11–20.jpg Number of workers supporting each beneficiary of the Social Security program in the United States, 1945–2075 (projected). Social Security taxes from workers in each current generation are used to pay current beneficiaries. (Data from United Nations) Figure_11–21.jpg Where are Japan, Thailand, Indonesia, India, China, and Bangladesh? Some of the countries highlighted here are discussed in other chapters. Figure_11–22.jpg Global aging. Projected percentage of world population under age 15, age 60 or over, and age 80 or over, 1950–2150, assuming the medium fertility projection shown in Figure 11–9. Between 1998 and 2050 the number of people over age 80 is projected to increase from 66 million to 370 million. The cost of supporting a much larger elderly population will place enormous strains on the world’s economy. (Data from the United Nations) Figure_11–23.jpg Legal immigration to the United States, 1820–1999. The large increase in immigration since 1989 resulted mostly from the Immigration Reform and Control Act of 1986, which granted legal status to illegal immigrants who could show that they had been living in the country for several years. (Data from U.S. Immigration and Naturalization Service) Figure_11–24.jpg Plots of a computer model projecting what might happen if the world’s population and economy continue growing exponentially at 1990 levels, assuming no major policy changes or technological innovations. Figure_11–25.jpg Computer-generated scenario projecting how we can avoid overshoot and collapse and make a fairly smooth transition to a sustainable future. Figure_11–26.jpg Generalized model of the demographic transition. (Data from United Nations) Figure_11–27.jpg Estimated global use of contraceptive methods in the 1990s by married women, ages 15–49. (Data from United Nations) Figure_11–28.jpg Typical workday for a woman in rural Africa. In addition to their domestic work, rural African women perform about 80% of all agricultural work. (Data from United Nations) Figure_11–29.jpg Comparison of basic demographic data for India and China. (Data from United Nations and Population Reference Bureau). Figure_12–1.jpg The Land Institute in Salina, Kansas, is a farm, a prairie laboratory, and a school dedicated to changing the way we grow food. It advocates growing a diverse mixture of edible perennial plants to supplement traditional annual monoculture crops. (Terry Evans) Figure_12–2.jpg Locations of the world’s principal types of food production. Excluding Antarctica and Greenland, agricultural systems cover almost one third of the earth’s land surface and account for an annual output of food worth about $1.3 trillion. Figure_12–3.jpg Relative inputs of land, labor, financial capital, and fossil fuel energy in four agricultural systems. An average of 60% of the people in developing countries are involved directly in producing food, compared with only 8% in developed countries (2% in the United States). Figure_12–4.jpg Countries whose crop yields per unit of land area increased during the two green revolutions. The first took place in developed countries between 1950 and 1970; the second has occurred since 1967 in developing countries with enough rainfall or irrigation capacity. Several agricultural research centers and gene or seed banks play a key role in developing high-yield crop varieties. Figure_12–5.jpg A high-yield, semidwarf variety of rice called IR-8 (left), a part of the second green revolution, was produced by crossbreeding two parent strains of rice: PETA from Indonesia (center) and DGWG from China (right). The shorter and stiffer stalks of the new variety allow the plants to support larger heads of grain without toppling over and increase the benefit of applying more fertilizer. (International Rice Research Institute, Manila) Figure_12–6.jpg General uses of land in the United States. (Data from U.S. Geological Survey and U.S. Department of Agriculture) Figure_12–7.jpg In the United States, industrialized agriculture uses about 17% of all commercial energy. On average, a piece of food eaten in the United States has traveled 2,100 kilometers (1,300 miles). Figure_12–8.jpg Total worldwide grain production of wheat, corn, and rice, and per capita grain production, 1950–2000. In order, the world’s three largest grain-producing countries in 2000 were China, the United States, and India. (Data from U.S. Department of Agriculture, WorldWatch Institute, and UN Food and Agriculture Organization) Figure_12–9.jpg Interactions between poverty, malnutrition, and disease form a tragic cycle that can perpetuate such conditions in succeeding generations of families Figure_12–10.jpg Major environmental effects of food production. Figure_12–11.jpg Traditional crossbreeding of species that are fairly close to one another genetically. Figure_12–12.jpg Steps in genetically modifying a plant. Figure_12–13.jpg Insects are important food items in many parts of the world. Mopani – emperor moth caterpillars – are among several insects eaten in South Africa. However, this food is so popular that the caterpillars (known as mopane worms) are being overharvested. Kalahari Desert dwellers eat cockroaches, lightly toasted butterflies are a favorite food in Bali, and French-fried ants are sold on the streets of Bogota, Colombia. Most of these insects are 58–78% protein by weight – three to four times as protein-rich as beef, fish, or eggs. (Anthony Bannister/Natural History Photographic Agency) Figure_12–14.jpg World irrigated area of cropland per person, 1900–2000, with projections to 2050. (Data from United Nations Food and Agriculture Organization, U.S. Census Bureau, and the WorldWatch Institute) Figure_12–15.jpg Classification of the earth’s land. Theoretically, we could double the amount of cropland by clearing tropical forests and irrigating arid lands. However, converting these lands into cropland would (1) destroy valuable forest resources, (2) reduce the earth’s biodiversity, (3) affect water quality and quantity, and (4) cause other serious environmental problems, usually without being cost-effective. Figure_12–16.jpg Average grain area per person worldwide, 1950–2000, with projections to 2030. (Data from U.S. Department of Agriculture and WorldWatch Institute) Figure_12–17.jpg Arid and semiarid regions of the world in which livestock (mostly cattle, sheep, and goats) can be raised (1) on open ranges where rainfall is low but fairly regular and (2) by nomadic herding where rainfall is so sparse and irregular that livestock must be moved to find adequate grass. Figure_12–18.jpg Rangeland grasses grow from the bottom up and are renewable as long as the bottom half of the plant (metabolic reserve), where photosynthesis takes place, is not eaten (top). If the metabolic reserve is eaten, the plant is weakened and can die (bottom) (From Environmental Science, 5/E by Chiras, p.208. Copyright (c) 1998 by Wadsworth.) Figure_12–19.jpg Effects of various degrees of grazing on the relative amounts of three major types of grassland plants. Decreasers are grass species that decline in abundance with moderate grazing; increasers are those that increase with moderate to heavy grazing pressure. Invaders are plants that colonize an area because of overgrazing or other changes in rangeland conditions. Figure_12–20.jpg Large-scale, factorylike production of livestock – mostly cattle, pigs, and poultry – in the United States and the resulting pollution problems are concentrated mostly in a few states. (Data from the U.S. Department of Agriculture) Figure_12–21.jpg Efficiency of converting grain into animal protein. Data in kilograms of grain per kilogram of body weight added. (Data from U.S. Department of Agriculture) Figure_12–22.jpg Rangeland: overgrazed (left) and lightly grazed (right). (USDA, Natural Resources Conservation Service) Figure_12–23.jpg Cattle on a riparian zone of a public rangeland along Arizona’s San Pedro River (left) in the mid-1980s just before this section of waterway was protected by banning domestic livestock grazing for 15 years, eliminating sand and gravel operations and water pumping rights in nearby areas, and limiting access by off-highway vehicles. The photo on the right shows the recovery of this riparian area at the same time of year after 10 years of protection. (Bureau of Land Management) Figure_12–24.jpg Deferred grazing plan in which during a 6-year rotation cycle each fenced in field gets nearly a 2-year rest from grazing. (From Environmental Science, 5/E by Chiras, 208. Copyright (c) 1998 by Wadsworth.) Figure_12–25.jpg Some major types of commercially harvested marine fish and shellfish. ... Celkem 599 obrázků