biosphere

biosphere
biospheric /buy'euh sfer"ik/, adj.
/buy"euh sfear'/, n.
1. the part of the earth's crust, waters, and atmosphere that supports life.
2. the ecosystem comprising the entire earth and the living organisms that inhabit it.
[1895-1900; < G Biosphäre; see BIO-, -SPHERE]

* * *

Relatively thin life-supporting stratum of the earth's surface, extending from a few miles into the atmosphere to the deep-sea vents of the oceans.

The biosphere is a global ecosystem that can be broken down into regional or local ecosystems, or biomes. Organisms in the biosphere are classified into trophic levels (see food chain) and communities.

* * *

Introduction
 relatively thin life-supporting stratum of the Earth's surface, extending from a few kilometres into the atmosphere to the deep-sea vents of the ocean. The biosphere is a global ecosystem composed of living organisms (biota) and the abiotic (nonliving) factors from which they derive energy and nutrients.

      Before the coming of life, the Earth was a bleak place, a rocky globe with shallow seas and a thin band of gases—largely carbon dioxide, carbon monoxide, molecular nitrogen, hydrogen sulfide, and water vapour. It was a hostile and barren planet. This strictly inorganic state of the Earth is called the geosphere; it consists of the lithosphere (the rock and soil), the hydrosphere (the water), and the atmosphere (the air). Energy from the Sun relentlessly bombarded the surface of the primitive Earth, and in time—millions of years—chemical and physical actions produced the first evidence of life: formless, jellylike blobs that could collect energy from the environment and produce more of their own kind. This generation of life in the thin outer layer of the geosphere established what is called the biosphere, the “zone of life,” an energy-diverting skin that uses the matter of the Earth to make living substance.

      The biosphere is a system characterized by the continuous cycling of matter and an accompanying flow of solar energy in which certain large molecules and cells are self-reproducing. Water is a major predisposing factor, for all life depends on it. The elements carbon, hydrogen, nitrogen, oxygen, phosphorus, and sulfur, when combined as proteins, lipids, carbohydrates, and nucleic acids, provide the building blocks, the fuel, and the direction for the creation of life. Energy flow is required to maintain the structure of organisms by the formation and splitting of phosphate bonds. Organisms are cellular in nature and always contain some sort of enclosing membrane structure, and all have nucleic acids that store and transmit genetic information.

      All life on Earth depends ultimately upon green plants (plant), as well as upon water. Plants utilize sunlight in a process called photosynthesis to produce the food upon which animals feed and to provide, as a by-product, oxygen, which most animals require for respiration. At first, the oceans and the lands were teeming with large numbers of a few kinds of simple single-celled organisms, but slowly plants and animals of increasing complexity evolved. Interrelationships developed so that certain plants grew in association with certain other plants, and animals associated with the plants and with one another to form communities of organisms, including those of forests, grasslands, deserts, dunes, bogs, rivers, and lakes. Living communities and their nonliving environment are inseparably interrelated and constantly interact upon each other. For convenience, any segment of the landscape that includes the biotic and abiotic components is called an ecosystem. A lake is an ecosystem when it is considered in totality as not just water but also nutrients, climate, and all of the life contained within it. A given forest, meadow, or river is likewise an ecosystem. One ecosystem grades into another along zones termed ecotones (ecotone), where a mixture of plant and animal species from the two ecosystems occurs. A forest considered as an ecosystem is not simply a stand of trees but is a complex of soil, air, and water, of climate and minerals, of bacteria, viruses, fungi, grasses, herbs, and trees, of insects, reptiles, amphibians, birds, and mammals.

      Stated another way, the abiotic, or nonliving, portion of each ecosystem in the biosphere includes the flow of energy, nutrients, water, and gases and the concentrations of organic and inorganic substances in the environment. The biotic, or living, portion includes three general categories of organisms based on their methods of acquiring energy: the primary producers, largely green plants; the consumers, which include all the animals; and the decomposers, which include the microorganisms that break down the remains of plants and animals into simpler components for recycling in the biosphere. Aquatic ecosystems are those involving marine environments and freshwater environments on the land. Terrestrial ecosystems are those based on major vegetational types, such as forest, grassland, desert, and tundra. Particular kinds of animals are associated with each such plant province.

      Ecosystems may be further subdivided into smaller biotic units called communities (community). Examples of communities include the organisms in a stand of pine trees, on a coral reef, and in a cave, a valley, a lake, or a stream. The major consideration in the community is the living component, the organisms; the abiotic factors of the environment are excluded.

      A community is a collection of species populations. In a stand of pines, there may be many species of insects, of birds, of mammals, each a separate breeding unit but each dependent on the others for its continued existence. A species, furthermore, is composed of individuals, single functioning units identifiable as organisms. Beyond this level, the units of the biosphere are those of the organism: organ systems composed of organs, organs of tissues, tissues of cells, cells of molecules, and molecules of atomic elements and energy. The progression, therefore, proceeding upward from atoms and energy, is toward fewer units, larger and more complex in pattern, at each successive level.

      This article focuses on the makeup of the biosphere and examines the relationships between its principal components, including man. The characteristics and dynamics of biological populations and communities are dealt with, as are the interactions that constitute the primary stabilizing links among the constituent organisms. Due attention is also given to the distribution patterns of these biotic units and to the processes that produced such patterns. The major aquatic and terrestrial ecosystems of the Earth are treated in some detail. Other points include energy transformations and transfers within the biosphere and the cyclic flow of materials needed for life. For the development, methodology, and applications of the study of interrelations of organisms with their environment and each other, see ecology. Further treatment of the various aquatic and terrestrial environments is provided in ocean, lake, river, continental landform, Arctic, and Antarctica. For a discussion of the origin of life on Earth and the varieties of and commonalities among organisms, see life and Earth, pregeologic history of (Earth, geologic history of). The characteristics and classifications of living organisms are covered in detail in algae, amphibian, angiosperm, animal, annelid, arachnid, arthropod, aschelminth, bacteria, bird, bryophyte, chordate, cnidarian, crustacean, dinosaur, echinoderm, fern (lower vascular plant), fish, flatworm, fungus, gymnosperm, insect, lamp shell (lamp shells), mammal, mollusk, moss animal, plant, protist, protozoa (protozoan), reptile, sponge, and virus.

David M. Gates Ed.

The diversity of life
 The biosphere supports between 3 and 30 million species of plants, animals, fungi, single-celled prokaryotes such as bacteria, and single-celled eukaryotes such as protozoans (Figure 1—>). Of this total, only about 1.4 million species have been named so far, and fewer than 1 percent have been studied for their ecological relationships and their role in ecosystems. A little more than half the named species are insects, which dominate terrestrial and freshwater communities worldwide; the laboratories of systematists are filled with insect species yet to be named and described. Hence, the relationships of organisms to their environments and the roles that species play in the biosphere are only beginning to be understood.

The organization of the biosphere
Natural groupings
      This tremendous diversity of life is organized into natural ecological groupings. As life has evolved, populations of organisms have become separated into different species that are reproductively isolated from one another. These species are organized through their interrelationships into complex biological communities. The interactions in these communities affect, and are affected by, the physical environments in which they occur, thereby forming ecosystems through which the energy and nutrients necessary for life flow and cycle. The mix of species and physical environments vary across the globe, creating ecological communities, or biomes, such as the boreal forests of North America and Eurasia and the rainforests of the tropics. The sum total of the richness of these biomes is the biosphere.

Processes of evolution
      This hierarchical organization of life has come about through the major processes of evolution— natural selection (the differential success of the reproduction of hereditary variations resulting from the interaction of organisms with their environment), gene flow (the movement of genes among different populations of a species), and random genetic drift (the genetic change that occurs in small populations owing to chance). (See evolution.) Natural selection operates on the expressed characteristics of genetic variants found within populations, winnowing members of the population who are less well suited to their environment from those better suited to it. In this manner, populations become adapted to their local ecosystems, which include both the physical environment and the other species with which they interact in order to survive and reproduce.

      The genetic variation that is necessary for a species to adapt (adaptation) to the physical environment and to other organisms arises from new mutations within populations, the recombination of genes during sexual reproduction, and the migration of and interbreeding with individuals from other populations. In very small populations, however, some of that variation is lost by chance alone through random genetic drift. The combined result of these evolutionary processes is that after many generations populations of the same species have widely divergent characteristics. Some of these populations eventually become so genetically different that their members cannot successfully interbreed, resulting in the evolution of a separate species ( speciation).

      The diversification of life through local adaptation of populations and speciation has created the tremendous biodiversity found on the Earth. In most regions a square kilometre (0.4 square mile) will harbour hundreds—in some places even thousands—of species. The interactions between these species create intricate webs of relationships as the organisms reciprocally evolve, adapting to one another and becoming specialized for their interactions (coevolution; see community ecology: The coevolutionary process (community ecology)). Natural communities of species reflect the sum of these species' interactions and the ongoing complex selection pressures they constantly endure that drive their evolution. The many ecological and evolutionary processes that affect the relationships among species and their environments render ecology one of the most intricate of the sciences. The answers to the major questions in ecology require an understanding of the relative effects of many variables acting simultaneously.

The importance of the biosphere
      The continued functioning of the biosphere is dependent not only on the maintainance of the intimate interactions among the myriad species within local communities but also on the looser yet crucial interactions of all species and communities around the globe. The Earth is blanketed with so many species and so many different kinds of biological communities because populations have been able to adapt to almost any kind of environment on Earth through natural selection. Life-forms have evolved that are able to survive in the ocean depths, the frigid conditions of Antarctica, and the near-boiling temperatures of geysers. The great richness of adaptations found among different populations and species of living organisms is the Earth's greatest resource. It is a richness that has evolved over millions of years and is irreplaceable.

      It is therefore startling to realize that our inventory of the Earth's diversity is still so incomplete that the total number of living species cannot be estimated more closely than between 3 and 30 million species. Decades of continuous research must be carried out by systematists, ecologists, and geneticists before the inventory of biodiversity provides a more accurate count. The research has been slow. Only recently, as the extinction rate of species has been increasing rapidly, have societies begun to realize the interdependence of species. To sustain life on Earth, more than the few animal and plant species used by humans must be preserved. The flow of energy and the cycling of nutrients through ecosystems, the regulation of populations, and the stability of biological communities, all of which support the continued maintenance of life, rely on the diversity of species, their adaptations to local physical conditions, and their coevolved relationships.

      Despite the limited scientific knowledge of most species, ecological studies during the 20th century have made great headway in unraveling the mechanisms by which organisms coevolve with one another and adapt to their physical environment, thereby shaping the biosphere. Each new decade has produced a steady stream of studies showing that the biological and physical elements of the Earth are more interconnected than had been previously thought. Those studies also have shown that often the most seemingly insignificant species are crucial to the stability of communities and ecosystems. Many seemingly obscure species are at risk worldwide of being dismissed as unimportant. The effect that the loss of species will have on ecosystems is appreciated only by understanding the relationships between organisms and their environments and by studying the ecological and evolutionary processes operating within ecosystems.

      The need to understand how the biosphere functions has never been greater. When human population levels were low and technological abilities crude, societies' impact on the biosphere was relatively small. The increase in human population levels and the harvesting of more of the Earth's natural resources has altered this situation, especially in recent decades. Human activities are causing major alterations to the patterns of energy flow and nutrient cycling through ecosystems, and these activities are eliminating populations and species that have not even been described but which might have been of central importance to the maintenance of ecosystems.

      The biologist Edward O. Wilson (Wilson, Edward O.), who coined the term “biodiversity,” estimated conservatively that in the late 20th century at least 27,000 species are becoming extinct each year. The majority of these are small tropical organisms. The impact that this freshet of extinctions would have on the biosphere is akin to receiving a box of engine parts and discarding a portion of them before reading the directions, assuming that their absence will have no negative repercussions on the running of the engine. The following sections describe how many of the biological and physical parts fit together to make the engine of the biosphere run and why many seemingly obscure species are important to the long-term functioning of the biosphere.

Resources of the biosphere

The flow of energy
The photosynthetic process (photosynthesis)
      Life on Earth depends on the harnessing of solar energy by the process of photosynthesis. Photosynthetic plants (plant) convert solar energy into the chemical energy of living tissue, and that stored chemical energy flows into herbivores, predators, parasites, decomposers, and all other forms of life (see also photosynthesis). In the photosynthetic process, light energy is absorbed by the chlorophyll molecules of plants to convert carbon dioxide and water into carbohydrates and oxygen gas. Proteins, fats, nucleic acids, and other compounds also are synthesized during the process, as long as elements such as nitrogen, sulfur, and phosphorus are available.

Efficiency of solar energy utilization
      Most solar energy occurs at wavelengths unsuitable for photosynthesis. Between 98 and 99 percent of solar energy reaching the Earth is reflected from leaves and other surfaces and absorbed by other molecules, which convert it to heat. Thus, only 1 to 2 percent is available to be captured by plants. The rate at which plants photosynthesize depends on the amount of light reaching the leaves, the temperature of the environment, and the availability of water and other nutrients such as nitrogen and phosphorus. The measurement of the rate at which organisms convert light energy (or inorganic chemical energy) to the chemical energy of organic compounds is called primary productivity. Hence, the total amount of energy assimilated by plants in an ecosystem during photosynthesis (gross primary productivity) varies among environments. (Productivity is often measured by an increase in biomass, a term used to refer to the weight of all living organisms in an area. Biomass is reported in grams or metric tons.)

       Average net primary production of the Earth's major habitatsMuch of the energy assimilated by plants through photosynthesis is not stored as organic material but instead is used during cellular respiration. In this process organic compounds such as carbohydrates, proteins, and fats are broken down, or oxidized, to provide energy (in the form of adenosine triphosphate [ATP]) for the cell's metabolic needs. The energy not used in this process is stored in plant tissues for further use and is called net primary productivity. About 40 to 85 percent of gross primary productivity is not used during respiration and becomes net primary productivity. The highest net primary productivity in terrestrial environments occurs in swamps (swamp) and marshes (marsh) and tropical rainforest (tropical rainforest)s; the lowest occurs in deserts (Table 1 (Average net primary production of the Earth's major habitats)). In aquatic environments, the highest net productivity occurs in estuaries, algal beds, and reefs. Consequently, these environments are especially critical for the maintenance of worldwide biological productivity.

Energy transfers and pyramids
      A small amount of the energy stored in plants, between 5 and 25 percent, passes into herbivores (plant eaters) as they feed, and a similarly small percentage of the energy in herbivores then passes into carnivores (animal eaters). The result is a pyramid of energy, with most energy concentrated in the photosynthetic organisms at the bottom of food chains (food chain) and less energy at each higher trophic level (a division based on the main nutritional source of the organism; see community ecology: Biotic elements of communities: Trophic pyramids and the flow of energy (community ecology)). Some of the remaining energy does not pass directly into the plant-herbivore-carnivore food chain but instead is diverted into the detritus food chain. Bacteria, fungi, scavengers, and carrion eaters that consume detritus (detritivores) are all eventually consumed by other organisms.

 The rate at which these consumers convert the chemical energy of their food into their own biomass is called secondary productivity. The efficiency at which energy is transferred from one trophic level to another is called ecological efficiency. On average it is estimated that there is only a 10 percent transfer of energy (Figure 2—>).

      Energy is lost in several ways as it flows along these pathways of consumption. Most plant tissue is uneaten by herbivores, and this stored energy is therefore lost to the plant-herbivore-carnivore food chain. In terrestrial communities less than 10 percent of plant tissue is actually consumed by herbivores. The rest falls into the detritus pathway, although the detritivores consume only some of this decaying tissue. Oil and coal deposits are major repositories of this unused plant energy and have accumulated over long periods of geologic time.

      The efficiency by which animals convert the food they ingest into energy for growth and reproduction is called assimilation efficiency. Herbivores (herbivore) assimilate between 15 and 80 percent of the plant material they ingest, depending on their physiology and the part of the plant that they eat. For example, herbivores that eat seeds and young vegetation high in energy have the highest assimilation efficiencies, those that eat older leaves have intermediate efficiencies, and those that feed on decaying wood have very low efficiencies. Carnivores generally have higher assimilation efficiencies than herbivores, often between 60 and 90 percent, because their food is more easily digested.

      The overall productivity of the biosphere is therefore limited by the rate at which plants convert solar energy (about 1 percent) into chemical energy and the subsequent efficiencies at which other organisms at higher trophic levels convert that stored energy into their own biomass (approximately 10 percent). Human-induced changes in net primary productivity in the parts of the biosphere that have the highest productivity, such as estuaries and tropical moist forests, are likely to have large effects on the overall biological productivity of the Earth.

Nutrient cycling
      The cells of all organisms are made up primarily of six major elements that occur in similar proportions in all life-forms. These elements— hydrogen, oxygen, carbon, nitrogen, phosphorus, and sulfur—form the core protoplasm of organisms, and the first four of these elements make up about 99 percent of the mass of most cells. Additional elements, however, are also essential to the growth of organisms. Calcium and other elements help to form cellular support structures such as shells, internal or external skeletons, and cell walls. Chlorophyll molecules, which allow photosynthetic plants (plant) to convert solar energy into chemical energy, are chains of carbon, hydrogen, and oxygen compounds built around a magnesium ion. Altogether, 16 elements are found in all organisms; another eight elements are found in some organisms but not in others.

      These bioelements combine with one another to form a wide variety of chemical compounds. They occur in organisms in higher proportions than they do in the environment because organisms capture them, concentrating and combining them in various ways in their cells, and release them during metabolism and death. As a result, these essential nutrients alternate between inorganic and organic states as they rotate through their respective biogeochemical cycles (biogeochemical cycle). These cycles can include all or part of the following: the atmosphere, which is made up largely of gases including water vapour; the lithosphere, which encompasses the soil and the entire solid crust of the Earth; and the hydrosphere, which includes lakes, rivers, and oceans (marine ecosystem).

 A portion of the elements are bound up in limestone and in the minerals of other rocks and are unavailable to organisms. The slow processes of weathering and erosion eventually release these elements to enter the cycle. For most of the major nutrients, however, organisms not only intercept the elements moving through the biosphere, but they actually drive the biogeochemical cycles (Figure 3—>).

      The movement of nutrients through the biosphere is different from the transfer of energy because, whereas energy flows through the biosphere and cannot be reused, elements are recycled. The same molecule of carbon or nitrogen may, over the course of eons, move repeatedly between organisms, the atmosphere, the soil, and the oceans. A molecule of carbon released as carbon dioxide by an animal may remain in the atmosphere for 5 or 10 years before being taken up by another organism, or it may cycle almost immediately back into a neighbouring plant and be used during photosynthesis.

The carbon cycle (carbon cycle)
      Life is built on the conversion of carbon dioxide into the carbon-based organic compounds of living organisms. The carbon cycle illustrates the central importance of carbon in the biosphere. Different paths of the carbon cycle recycle the element at varying rates. The slowest part of the cycle involves carbon that resides in sedimentary (sedimentation) rocks, where most of the Earth's carbon is stored. When in contact with water that is acidic (pH is low), carbon will dissolve from bedrock; under neutral conditions, carbon will precipitate out as sediment such as calcium carbonate (limestone). This cycling between solution and precipitation is the background against which more rapid parts of the cycle occur.

      Short-term cycling of carbon occurs in the continual physical exchange of carbon dioxide (CO2) between the atmosphere and hydrosphere. Carbon dioxide from the atmosphere becomes dissolved in water (H2O), with which it reacts to form carbonic acid (H2CO3), which dissociates into hydrogen ions (H+) and bicarbonate ions (HCO3-), which further dissociate into hydrogen and carbonate ions (CO32-). The more alkaline the water (pH above 7.0 is alkaline), the more carbon is present in the form of carbonate, as is shown in the following reversible reactions:

      At the same time, carbon dioxide in the water is continually lost to the atmosphere. The exchange of carbon between the atmosphere and hydrosphere links the remaining parts of the cycle, which are the exchanges that occur between the atmosphere and terrestrial organisms and between water and aquatic organisms.

      The biological cycling of carbon begins as photosynthetic organisms assimilate carbon dioxide or carbonates from the surrounding environment. In terrestrial communities, plants convert atmospheric carbon dioxide to carbon-based compounds through photosynthesis (see above The flow of energy: The photosynthetic process (biosphere)). During this process, plants cleave the carbon from the two oxygen molecules and release the oxygen back into the surrounding environment. Plants are thus primarily responsible for the presence of atmospheric oxygen. In aquatic communities, plants use dissolved carbon in the form of carbonates or carbon dioxide as the source of carbon for photosynthesis. Once carbon has been assimilated by photosynthetic organisms, as well as by the animals that eat them, it is released again in the form of carbon dioxide as these organisms respire. The release of carbon dioxide into the atmosphere or hydrosphere completes the biological part of the carbon cycle.

      The pathways of the global carbon cycle, however, are never completely balanced. That is to say, carbon does not move in and out of all parts of the biosphere at equal rates. Consequently, over time some parts of the biosphere accumulate more carbon than others, thereby serving as major accessible carbon reservoirs. In preindustrial times the major reservoirs of carbon were the deep and shallow portions of the ocean; the soil, detritus, and biota of the land; and the atmosphere. The oceans were, and still are, the greatest reservoirs of carbon. Because marine phytoplankton have such short life cycles, the carbon in the ocean cycles rapidly between inorganic and organic states.

      In terrestrial environments, forests (forest) are the largest carbon reservoirs. Up to 80 percent of the aboveground carbon in terrestrial communities and about a third of belowground carbon are contained within forests. Unlike the oceans, much of this carbon is stored directly in the tissues of plants. High-latitude forests include large amounts of carbon not only in aboveground vegetation but also in peat deposits. Forests at high and low latitudes particularly are important reservoirs of carbon. An estimated one-half of the carbon in forests occurs in high-latitude forests, and a little more than one-third occurs in low-latitude forests. The two largest forest reservoirs of carbon are the vast expanses in Russia, which hold roughly 25 percent of the world's forest carbon, and the Amazon basin, which contains about 20 percent.

      Until recent centuries, the equilibrium between the carbon in the world's forests and in the atmosphere remained constant. Samples of carbon dioxide trapped in ice during the past 1,000 years and direct measurements of carbon dioxide in the atmosphere had remained fairly constant until the 18th century. However, the cutting of much of the world's forest is, along with the increase in consumption of fossil fuels attendant on the Industrial Revolution, has resulted in a disruption of the balance between the amount of carbon dioxide in the forests and in the atmosphere. The concentration of atmospheric carbon dioxide has been increasing steadily (Figure 4); currently the rate of increase is about 4 percent per decade (see global warming and climatic variation and change (climate change)). If human activities continue to alter the relative sizes of the carbon reservoirs worldwide, they are likely to have large effects on the carbon cycle and other biogeochemical cycles. Large-scale deforestation in Russia and the Amazon basin are likely to have particularly significant effects on global (global warming) carbon storage and cycling.

      The warming of global temperatures also is changing which ecosystems act as long-term sinks for carbon and which act as sources for carbon dioxide in the atmosphere. For example, the Arctic tundra, with large amounts of carbon stored in its soils, has been a net sink for carbon dioxide during long periods of geologic time. The recent warming of many Arctic regions, however, has accelerated the rate of soil decomposition, transforming these Arctic areas into potential sources of atmospheric carbon dioxide.

      The complex cycle of carbon and the varying sizes of carbon reservoirs illustrate some of the reasons it has been so difficult to predict the effects that increased atmospheric carbon will have on global change.

The nitrogen cycle (nitrogen cycle)
 Nitrogen is one of the elements most likely to be limiting to plant growth. Like carbon, nitrogen has its own biogeochemical cycle, circulating through the atmosphere, lithosphere, and hydrosphere (Figure 5—>). Unlike carbon, which is stored primarily in sedimentary rock, most nitrogen occurs in the atmosphere as an inorganic compound (N2). It is the predominant atmospheric gas, making up about 79 percent of the volume of the atmosphere. Plants, however, cannot use nitrogen in its gaseous form and are able to assimilate it only after it has been converted to ammonia (NH3) and nitrates (NO3). This reductive process, called nitrogen fixation, is a chemical reaction in which electrons are picked up from another molecule. A small amount of nitrogen is fixed by lightning, but most of the nitrogen harvested from the atmosphere is removed by nitrogen-fixing bacteria and cyanobacteria (formerly called blue-green algae).

 Certain species of nitrogen-fixing bacteria can coexist intimately (symbiotically) with legumes and other plants, providing the plants with necessary nitrogen (Figure 6—>). In this symbiotic (symbiosis) association, the bacteria become encased in nodules that grow on the roots of plants, through which nitrogen that has been fixed by the resident bacteria is obtained. Cyanobacteria have developed similar relationships with various life-forms, such as liverworts, hornworts, cycads, and at least one genus of flowering plants (Gunnera). Their symbiotic relationship with fungi has earned its own designation—the coexistent species are called lichen.

      Other microorganisms perform important tasks that propel the nitrogen cycle along. Although plants can assimilate ammonia as well as nitrates, most of the ammonia in the soil is converted to nitrites (NO2) and then to nitrates by certain aerobic bacteria through the oxidative process of nitrification. Once nitrogen has been assimilated by plants, it can be converted to organic forms, such as amino acids and proteins. Animals can use only organic nitrogen, which they obtain by consuming plants or other animals. As these organisms die, certain microbes such as detritivores are able to participate in the decomposition of organic nitrogen into ammonia (ammonification), providing a constant supply of ammonia to be used in the process of nitrification. Although the fixation of atmospheric nitrogen is an essential part of the nitrogen cycle, ammonification and nitrification are the predominant methods by which organic nitrogen is prevented from returning to the atmosphere and is kept cycling through the biosphere.

      Some nitrogen does return to the atmosphere, however, as denitrifying bacteria break down nitrates to obtain oxygen, thereby releasing gaseous N2. Nitrogen is also lost from plants and soil in terrestrial environments via other routes, including erosion, runoff, volatilization of ammonia into the atmosphere, and leaching from soils into lakes and streams. Eventually some of these nutrients reach the oceans as rivers flush them onto the ocean surface.

The sulfur cycle (sulfur cycle)
      Sulfur is found in all living organisms as a constituent of some proteins, vitamins, and hormones. Like carbon and nitrogen, sulfur cycles between the atmosphere, lithosphere, and hydrosphere; but, unlike these two other elements, it has major reservoirs in both the atmosphere and the lithosphere. As is true in the nitrogen cycle, the activities of microorganisms are crucial in the global cycling of this nutrient.

      The process begins with geochemical and meteorologic processes such as the weathering of rock. When sulfur is released from the rock and comes in contact with air, it is converted into sulfate (SO4), which is taken up by plants and microorganisms and converted into organic forms. Animals acquire these organic forms of sulfur from their foods. When organisms die and decompose, some of the sulfur enters the tissues of microorganisms and some is released again as sulfate. There is, however, a continual loss of sulfur from terrestrial ecosystems as some of it drains into lakes and streams and eventually into the ocean as runoff. Additional sulfur enters the ocean through fallout from the atmosphere.

      Once in the ocean, some of the sulfur cycles through marine communities as it moves through food chains, some reenters the atmosphere, and some is lost to the ocean depths as it combines with iron to form ferrous sulfide (FeS), which is responsible for the black colour of marine sediments. Sulfur reenters the atmosphere naturally in three major ways: sea spray releases large amounts of the element from the ocean into the atmosphere; anaerobic respiration by sulfate-reducing bacteria causes the release of hydrogen sulfide (H2S) gas especially from marshes, tidal flats, and similar environments in which anaerobic microorganisms thrive; and volcanic activity releases additional but much smaller amounts of sulfur gas into the atmosphere.

      Since the Industrial Revolution, human activities have contributed significantly to the movement of sulfur from the lithosphere to the atmosphere as the burning of fossil fuels and the processing of metals have occasioned large emissions of sulfur dioxide. Oxides of sulfur and nitrogen contribute to the acid rain that is common downwind from these industrial activities (see acid rain).

The cycling of phosphorus (phosphorus cycle) and other essential nutrients
      Most other major nutrients such as phosphorus, potassium, magnesium, iron, and calcium enter terrestrial communities through the weathering of bedrock. These nutrients lack a volatile gaseous state. Consequently, they cycle through the biosphere differently from carbon, nitrogen, and sulfur, all of which sometimes occur as volatile gases. Of the nonvolatile nutrients, phosphorus is the one that most often limits plant growth, especially in aquatic environments.

      Phosphorus and the other nonvolatile elements move unidirectionally from land, through aquatic environments, into ocean sediments. Most phosphorus cycling occurs between the surface and depths of the ocean. When near the surface, phosphorus is taken up by the plankton and passed through the food chain. It cycles back to the ocean bottom as individuals die and fall to the ocean floor, releasing assimilated phosphorus. Much of this element gradually accumulates in the ocean sediment as particulate phosphorus and is eventually brought back to the surface only through ocean upwelling and tectonic activity. The ocean sediments are therefore by far the greatest reservoirs of phosphorus.

      In terrestrial ecosystems, much of the available phosphorus moves in a closed cycle between living organisms and the organic debris in the soil. phosphate (PO43-) is the only important inorganic form involved in this cycle. Microorganisms in the soil break down litter and other organic matter, thereby releasing phosphate, which is then taken up by plants and released again when they die and decompose. Soils differ in the amount of phosphorus they contain, and in some phosphorus-poor soils almost all the available phosphorus resides in living organisms and organic debris. In some tropical forests, such as those in parts of Brazil, so much of the phosphorus is contained in living organisms that the clearing of forests eliminates most of this element. As a result, the plant communities cannot recover, and crops cannot be grown.

      The addition of phosphorus to soils poor in this nutrient and the use of phosphorus-rich detergents have had a great impact on the phosphorus cycle in many ecosystems. Runoff (pollution) from agricultural fields and the emptying of sewage has introduced so much extra phosphorus to rivers and lakes that populations of plants and microorganisms have grown explosively, sometimes creating a solid mat that extends over the surface of the water. This growth increases the amount of organic debris in the water, which can lead to a decrease in the available oxygen, resulting in suffocation of fish and other animals.

The hydrologic cycle (hydrologic cycle)
      A portion of the biogeochemical cycle of all elements involves time spent cycling through the hydrosphere. Water itself cycles within the biosphere. (For a detailed discussion of the hydrologic cycle see hydrosphere: The hydrologic cycle (hydrosphere).) Unlike the cycles of the other major nutrients, however, the hydrologic, or water, cycle would continue in some form even in the absence of living organisms. Most of the Earth's water is in its core, in the sedimentary rocks near its surface, and in the ocean (marine ecosystem). A minute percentage of the water, however, continually cycles through the atmosphere, oceans, and terrestrial environments mainly by the processes of evaporation (vaporization) and precipitation.

      This part of the hydrologic cycle is driven by solar energy. Water evaporates from both the aquatic and terrestrial environments as it is heated by the Sun's energy. The rates of evaporation and precipitation depend on solar energy, as do the patterns of circulation of moisture in the air and currents in the ocean. Evaporation exceeds precipitation over the oceans, and this water vapour is transported by the wind over land, where it returns to the land through precipitation. The water falling onto terrestrial environments seeps into the ground or runs off into lakes and streams and eventually empties into the oceans, carrying with it many of the other major nutrients. Water also reenters the atmosphere through the evaporative loss of water by plants (plant) (transpiration).

Links among the cycles
      Although the overall pattern of cycling of all the major elements is now known, there is still much to learn about the relative importance of the different stages of each cycle. For example, there is considerable debate concerning which ecosystems act as the major sources of carbon for the atmosphere and which act as sinks by accumulating more carbon than they release. The ways in which the different cycles interact with one another also must be minutely studied. It has been discovered that sulfur availability influences the rate of nitrogen accumulation in plants and nitrogen availability influences phosphorus uptake. All three elements are thought to influence the rate of carbon accumulation by plants. As a result, changes in any one of these nutrient cycles influence the other cycles as well.

      The effects that disruptions in these cycles may have within the biosphere are not clearly understood. Natural geologic phenomena, such as ice ages and major periods of volcanic activity, have repeatedly disturbed these cycles throughout Earth history. Many human activities may have impacts of similar scope. Deforestation, the burning of fossil fuels, and increased fertilization are disturbing these cycles. These anthropogenic disturbances have increased atmospheric levels of carbon dioxide (Figure 4), decreased ozone (O3) levels, and undermined the natural equilibrium of streams and lakes by excessive nutrient enrichment from sewage, fertilizers, and factory waste (cultural eutrophication). Gleaning more information about the biogeochemical cycles and their interactions has become increasingly important now that the effects of human activities are becoming more apparent.

      Another potential effect that may result from human intrusions in the environment is global warming. Carbon dioxide in the atmosphere has the ability to act as an insulator, preventing some of the Earth's heat, absorbed from solar radiation, from escaping back into space. This process, known as the greenhouse effect, is suspected of being enhanced by rising levels of atmospheric carbon dioxide (Figure 4), which have resulted in part from the combustion of fossil fuels and the clearing and burning of tropical forests. This increase in atmospheric carbon dioxide and other so-called greenhouse gases could raise the overall global temperature, causing the polar ice caps to melt, sea levels to rise, and the Earth's precipitation to be redistributed.

      The complexity and interconnectedness of each of the biogeochemical cycles make it difficult to pinpoint how any one human activity is altering the cycles; nevertheless, the majority of those who study these fluctuations agree that this is happening. Disagreements generally concern the extent to which various activities affect particular cycles and what the long-term consequences of these disturbances will be.

John N. Thompson

Environmental (environment) conditions
      Most organisms are limited to either a terrestrial or an aquatic environment. An organism's ability to tolerate local conditions within its environment further restricts its distribution. One parameter, such as temperature tolerance, may be important in determining the limits of distribution, but often a combination of variables, such as temperature tolerance and water requirements, is important. Extreme environmental variables can evoke physiological and behavioral responses from organisms. The physiological response helps the organism maintain a constant internal environment (homeostasis), while a behavioral response allows it to avoid the environmental challenge—a fallback strategy if homeostasis cannot be maintained.

      The ways in which modern living organisms tolerate environmental conditions reflect the aquatic origins of life. With few exceptions, life cannot exist outside the temperature range at which water is a liquid. Thus, liquid water, and temperatures that maintain water as a liquid, are essential for sustaining life. Within those parameters, the concentrations of dissolved salts and other ions, the abundance of respiratory gases, atmospheric or hydrostatic pressure, and rate of water flow all influence the physiology, behaviour, and distribution of organisms.

      Temperature has the single most important influence on the distribution of organisms because it determines the physical state of water. Most organisms cannot live in conditions in which the temperature remains below 0° C or above 45° C for any length of time. Adaptations have enabled certain species to survive outside this range—thermophilic bacteria have been found in hot springs in which the temperatures may approach the boiling point, and certain polar mosses and lichens can tolerate temperatures of -70° C—but these species are the exceptions. Few organisms can remain for long periods at temperatures above 45° C, because organic molecules such as proteins will begin to denature. Nor are temperatures below freezing conducive to life: cells will rupture if the water they contain freezes.

      Most organisms are not able to maintain a body temperature that is significantly different from that of the environment. Sessile organisms, such as plants and fungi, and very small organisms and animals that cannot move great distances, therefore, must be able to withstand the full range of temperatures sustained by their habitat. In contrast, many mobile animals employ behavioral mechanisms to avoid extreme conditions in the short term. Such behaviours vary from simply moving short distances out of the Sun or an icy wind to large-scale migrations.

      Some types of animals employ physiological mechanisms to maintain a constant body temperature, and two categories are commonly distinguished: the term “cold-blooded (cold-bloodedness)” is understood to refer to reptiles and invertebrates, and “warm-blooded” is generally applied to mammals and birds. These terms, however, are imprecise; the more accurate terms, ectotherm for cold-blooded and endotherm for warm-blooded, are more useful in describing the thermal capabilities of these animals. Ectotherms rely on external sources of heat to regulate their body temperatures, and endotherms thermoregulate by generating heat internally.

 Terrestrial ectotherms utilize the complex temperature profile of the terrestrial environment to derive warmth. They can absorb solar radiation, thus raising their body temperatures above that of the surrounding air and substrate (Figure 7—>), unlike the aquatic ectotherm, whose body temperature is usually very close to that of the environment. As this solar radiation is taken up, physiological mechanisms contribute to the regulation of heat—peripheral blood vessels dilate and heart rate increases. The animal also may employ behavioral mechanisms, such as reorienting itself toward the Sun or flattening its body and spreading its legs to maximize its surface area exposure. At night, loss of heat may be reduced by other behavioral and physiological mechanisms—the heart rate may slow, peripheral blood vessels may constrict, surface area may be minimized, and shelter may be sought.

      Endotherms maintain body temperature independently of the environment by the metabolic production of heat. They generate heat internally and control passive heat loss by varying the quality of their insulation or by repositioning themselves to alter their effective surface area (i.e., curling into a tight ball). If heat loss exceeds heat generation, metabolism increases to make up the loss. If heat generation exceeds the rate of loss, mechanisms to increase heat loss by evaporation occur. In either case, behavioral mechanisms can be employed to seek a more suitable thermal environment.

      To survive for a limited period in adverse conditions, endotherms may employ a combination of behavioral and physiological mechanisms. In cold weather, which requires an increase in energy consumption, the animal may enter a state of torpor in which its body temperature, metabolism, respiratory rate, and heart rate are depressed. Long-term winter hypothermia, or hibernation, is an extended state of torpor that some animals use as a response to cold conditions. Torpor and hibernation free the animals from energetically expensive maintenance of high body temperatures, saving energy when food is limited.

      Another form of torpor, estivation, is experienced by animals in response to heat stress. This state is seen more often in ectothermic animals than in endotherms, but in both the stimulus for estivation is usually a combination of high temperatures and water shortage.

      Most terrestrial organisms must maintain their water content within fairly narrow limits. Water commonly is lost to the air through evaporation (vaporization) or, in plants, transpiration. Because most water loss occurs by diffusion and the rate of diffusion is determined by the gradient across the diffusion barrier such as the surface of a leaf or skin, the rate of water loss will depend on the relative humidity of the air. Relative humidity is the percent saturation of air relative to its total saturation possible at a given temperature. When air is totally saturated, relative humidity is said to be 100 percent. Cool air that is completely saturated contains less water vapour than completely saturated warm air because the water vapour capacity of warm air is greater (see climate: Atmospheric humidity (climate)). Diffusion gradients across skin or leaves, therefore, can be much steeper in summer when the air is warm, rendering evaporative water loss a much more serious problem in warm environments than in cool environments. Nevertheless, rates of water loss are higher in dry air (conditions of low relative humidity) than in moist air (conditions of high relative humidity), regardless of the temperature.

      Water loss from evaporation must be compensated by water uptake from the environment. For most plants, transpirational water loss is countered by the uptake of water from the soil via roots. For animals, water content can be replenished by eating or drinking or by uptake through the integument. For organisms living in dry environments, there are many morphological and physiological mechanisms that reduce water loss. Desert plants, or xerophytes, typically have reduced leaf surface areas because leaves are the major sites of transpiration. Some xerophytes shed their leaves altogether in summer, and some are dormant during the dry season.

      Desert animals typically have skin that is relatively impervious to water. The major site of evaporation is the respiratory exchange surface, which must be moist to allow the gaseous exchange of oxygen and carbon dioxide. A reduction in amount of water lost through respiration can occur if the temperature of the exhaled air is lower than the temperature of the body. As many animals, such as gazelles, inhale warm air, heat and water vapour from the nasal passages evaporate, cooling the nose and the blood within it. The cool venous blood passes close to and cools the warm arterial blood traveling to the brain. If the brain does not require cooling, the venous blood returns to the heart by another route. The nasal passages also cool the warm, saturated air from the lungs so that water condenses in the nose and is reabsorbed rather than lost to the environment.

      The relative acidity or alkalinity of a solution is reported by the pH scale, which is a measure of the concentration of hydrogen ions in solution. Neutral solutions have a pH of 7. A pH of less than 7 denotes acidity (an increased hydrogen ion concentration), and above 7 alkalinity (a decreased hydrogen ion concentration). Many important molecular processes within the cells of organisms occur within a very narrow range of pH. Thus, maintenance of internal pH by homeostatic mechanisms is vital for cells to function properly. Although pH may differ locally within an organism, most tissues are within one pH unit of neutral. Because aquatic organisms generally have somewhat permeable skins or respiratory exchange surfaces, external conditions can influence internal pH. These organisms may accomplish the extremely important task of regulating internal pH by exchanging hydrogen ions for other ions, such as sodium or bicarbonate, with the environment.

      The pH of naturally occurring waters can range from very acidic conditions of about 3 in peat swamps to very alkaline conditions of about 9 in alkaline lakes. Naturally acidic water may result from the presence of organic acids, as is the case in a peat swamp, or from geologic conditions such as sulfur deposits associated with volcanic activity. Naturally occurring alkaline waters usually result from inorganic sources. Most organisms are unable to live in conditions of extreme alkalinity or acidity.

Salinity
      The term “salinity” refers to the amount of dissolved salts that are present in water. Sodium and chloride are the predominant ions (ion) in seawater, and the concentrations of magnesium, calcium, and sulfate ions are also substantial. Naturally occurring waters vary in salinity from the almost pure water, devoid of salts, in snowmelt to the saturated solutions in salt lakes such as the Dead Sea. Salinity in the oceans is constant but is more variable along the coast where seawater is diluted with freshwater from runoff or from the emptying of rivers. This brackish water forms a barrier separating marine and freshwater organisms.

      The cells of organisms also contain solutions of dissolved ions, but the range of salinity that occurs in tissues is more narrow than the range that occurs in nature. Although a minimum number of ions must be present in the cytoplasm for the cell to function properly, excessive concentrations of ions will impair cellular functioning. Organisms that live in aquatic environments and whose integument is permeable to water, therefore, must be able to contend with osmotic pressure. This pressure arises if two solutions of unequal solute concentration exist on either side of a semipermeable membrane such as the skin. Water from the solution with a lower solute concentration will cross the membrane diluting the more highly concentrated solution until both concentrations are equalized. If the salt concentration of an animal's body fluids is higher than that of the surrounding environment, the osmotic pressure will cause water to diffuse through the skin until the concentrations are equal unless some mechanism prevents this from happening.

      Many marine invertebrates have the same osmotic pressure as seawater. When the salt concentration of their surroundings changes, however, they must be able to adjust. Two means of contending with this situation are employed, and, depending on how they regulate the salt concentrations of their tissues, organisms are classified as osmoregulators or osmoconformers. The osmotic concentration of the body fluids of an osmoconformer changes to match that of its external environment, whereas an osmoregulator (osmoregulation) controls the osmotic concentration of its body fluids, keeping them constant in spite of external alterations. Aquatic organisms that can tolerate a wide range of external ion concentrations are called euryhaline; those that have a limited tolerance are called stenohaline.

      Even if aquatic organisms have an integument that is relatively impermeable to water, as well as to small inorganic ions, their respiratory exchange surfaces are permeable. Hence, organisms occurring in water that has a lower solute concentration than their tissues (e.g., trout in mountain streams) will constantly lose ions to the environment as water flows into their tissues. In contrast, organisms in salty environments face a constant loss of water and an influx of ions.

 Many mechanisms have evolved that deal with these problems. Because water cannot be readily pumped across cell membranes, salinity balance is usually maintained by actively transporting inorganic ions, usually sodium and chloride. This process consumes energy and can usurp a large portion of the energy budget of animals in very saline environments. In marine fish, gill cells pump ions out of the body into the sea, while in freshwater fish gill cells pump ions in the opposite direction. Passive water loss in marine fish is compensated primarily in one of two ways. Most bony fish drink copiously and excrete salt across the gills, while the majority of sharks artificially elevate the salt concentration of their tissues above that of seawater with urea and other organic molecules, allowing water to slowly and passively enter the body. Through their food and across their gills, freshwater fish replenish most of the ions they lose. They also produce large quantities of very dilute urine to excrete excess water that diffuses into their bodies (Figure 8—>).

Water currents
      The flow of water presents special problems for aquatic organisms. Flow is associated with rivers, oceanic currents (ocean current), and waves and can be laminar (streamlined) or turbulent. Many organisms are specialized to live in flowing environments; the main obstacle to this lifestyle is the constant threat of being washed away. Both plants and animals have evolved mechanisms that help to anchor them to the substratum in flowing water (e.g., the holdfast of kelp or the byssus threads of mussels). If anchorage can be assured, there are many advantages to living in this environment. Flowing water generally is well oxygenated, and the supply is continuous; nutrients and food are constantly replenished as well. The very precariousness of the environment also affords some protection from predation because the number of predators that make this type of habitat home is limited.

      Variations in atmospheric pressure can present special problems for the respiratory systems of animals because atmospheric pressure affects the exchange of oxygen and carbon dioxide that occurs during animal respiration. Normal atmospheric pressure at sea level is the total pressure that a column of air above the surface of the Earth exerts (760 millimetres of mercury, or 1 atmosphere). The total pressure is the sum of the pressures that each gas—mainly nitrogen, oxygen, and carbon dioxide—would exert alone (the partial pressure of that gas; see respiration: The gases in the environment (respiration)). As an animal breathes, oxygen moves from the environment across the respiratory surfaces into the blood; carbon dioxide moves in the reverse direction. This process occurs primarily by passive diffusion; each gas moves from an area of greater to lesser partial pressure, driven by the differential that exists across the respiratory surface. At higher altitudes, where the atmospheric pressure is lower, the partial pressure of oxygen is also lower. The partial pressure differential of oxygen, therefore, is also lower, and the organism effectively receives less oxygen when it breathes, even though the percentage of oxygen in the air remains constant. This lack of oxygen is why humans carry oxygen when ascending to high altitudes. Humans who live in mountainous regions, however, can become acclimatized (acclimatization) to the lowered availability of oxygen, and certain animals such as llamas have adaptations of the blood that allow them to live at high altitudes. Birds have very efficient lungs, and many apparently have no problems flying to high altitudes, even for extended flights (see respiration: High altitudes (respiration, human)).

Hydrostatic pressure
      Because air and water have vastly different densities, the pressures experienced in terrestrial and aquatic habitats differ markedly. A column of water, so much denser than air, exerts a greater amount of pressure than a column of air. With each 10-metre (32.8-foot) increase in depth, there is an increase in hydrostatic pressure equivalent to one atmosphere. Mean ocean depth is about 3,800 metres and has a pressure of about 380 atmospheres. To surmount this environmental challenge, animals that live at great depths lack air compartments such as lungs or swim bladders. Surface-dwelling animals that dive to great depths meet this challenge differently. As pressure increases during a dive, air compartments compress, returning to their former volume when the animal surfaces. Air is forced into the trachea, bronchi, and bronchioles, where no gas uptake occurs. Thus, the increased pressure cannot drive more gases into the bloodstream, and, as the animal rises, it does not experience the “bends” (decompression sickness resulting from a rapid reduction of air pressure). In contrast, sea snakes avoid the bends by excreting nitrogen across the skin to offset the uptake of this gas from the lungs.

Michael B. Thompson

Additional Reading

General works
Mitchell B. Rambler, Lynn Margulis, and René Fester (eds.), Global Ecology: Towards a Science of the Biosphere (1989), describes and interprets the biosphere and the processes that occur within it. Richard J. Huggett, Climate, Earth Processes, and Earth History (1991), discusses the changing of the biosphere over time. William K. Purves, Gordon H. Orians, and H. Craig Heller, Life: The Science of Biology, 4th ed. (1994), treats such topics as general ecology, the biosphere, and the origin of life. Paul R. Ehrlich and Jonathan Roughgarden, The Science of Ecology (1987), describes organisms in their environments. Lawrence E. Joseph, Gaia: The Growth of an Idea (1990), is a simple explanation of the Gaia hypothesis. Leslie A. Real and James H. Brown (eds.), Foundations of Ecology: Classic Papers with Commentaries (1991), provides a historical perspective of the major issues in ecology. Nigel Pears, Basic Biogeography, 2nd ed. (1985), examines the distribution and abundance of life on the Earth. Lynn Margulis and Lorraine Olendzenski (eds.), Environmental Evolution (1992), discusses the interaction of life and the abiotic components of the Earth, as well as the evolution of life as a consequence of changes to the biosphere over time.The environmental physiology of animals is covered in F. Harvey Pough, John B. Heiser, and William N. McFarland, Vertebrate Life (1989); Knut Schmidt-Nielsen, Animal Physiology: Adaptation and Environment, 4th ed. (1990); and Philip C. Withers, Comparative Animal Physiology (1992). J. Prothero and K.D. Jurgens, “An Energetic Model of Daily Torpor in Endotherms,” Journal of Theoretical Biology, 121(4):403–416 (1986), explores the advantages to endothermic animals of entering torpor. Cynthia Carey et al. (eds.), Life in the Cold: Ecological, Physiological, and Molecular Mechanisms (1993), investigates the interactions of organisms with cold environments from ecological and physiological perspectives. Harold Heatwole and Janet Taylor, Ecology of Reptiles (1987), sets forth the major environmental parameters that influence the lives of animals, using reptilian examples. Knut Schmidt-Nielsen, F.R. Hainsworth, and D.E. Murrish, “Counter-Current Heat Exchange in the Respiratory Passages: Effects on Water and Heat Balance,” Respiration Physiology, 9(7):263–276 (May 1970), describes animals' utilization of the nasal passages as a means to allow thermoregulation and simultaneously conserve water. R.S. Seymour, “How Sea Snakes May Avoid the Bends,” Nature, 250(5466):489–490 (Aug. 9, 1974), outlines the physical and physiological influences affecting animals that breathe air on the surface but dive to great depths. David W. Goodall, Ecosystems of the World (1977– ), discusses terrestrial and aquatic ecosystems.Michael B. Thompson

The organism and the environment
Michael Begon, John L. Harper, and Colin R. Townsend, Ecology: Individuals, Populations, and Communities, 2nd ed. (1990); Robert E. Ricklefs, Ecology, 3rd ed. (1990); Charles J. Krebs, Ecology: The Experimental Analysis of Distribution and Abundance, 4th ed. (1994); and the work by Ehrlich and Roughgarden, cited above, are well-written textbooks that provide good general descriptions of energy flow and nutrient cycling through ecosystems, population biology, and community ecology.D.L. DeAngelis, Dynamics of Nutrient Cycling and Food Webs (1992), is a mathematical treatment of the rates of energy flow and nutrient cycling through ecosystems. B. Bolin and R.B. Cook (eds.), The Major Biogeochemical Cycles and Their Interactions (1983), discusses the major global patterns of nutrient cycling for all the major nutrients. An in-depth treatment of how the major nutrients necessary for plant growth move in cycles between plants and the soil can be found in F.J. Stevenson, Cycles of Soil: Carbon, Nitrogen, Phosphorus, Sulfur, Micronutrients (1986). A review of the groups of microorganisms and their various roles in the nitrogen cycle is Janet I. Sprent and Peter Sprent, Nitrogen Fixing Organisms: Pure and Applied Aspects (1990).David A. Dunnette and Robert J. O'brien (eds.), The Science of Global Change: The Impact of Human Activities on the Environment (1992), reports on the ways in which human activities are influencing biogeochemical cycles and climate change. Robert L. Peters and Thomas E. Lovejoy (eds.), Global Warming and Biological Diversity (1992); and Peter M. Kareiva, Joel G. Kingsolver, and Raymond B. Huey (eds.), Biotic Interactions and Global Change (1993), investigate how human-induced global changes affect organisms, population, species, communities, and ecosystems.John N. Thompson

* * *


Universalium. 2010.

Игры ⚽ Нужна курсовая?

Look at other dictionaries:

  • Biosphere 2 — is a 3.14 acre (1.27 ha) [ [http://www.b2science.org/about facts.html Biosphere 2 Where science lives ] ] structure originally built to be an artificial closed ecological system in Oracle, Arizona (USA) by Space Biosphere Ventures, a company… …   Wikipedia

  • Biosphere 2 — Biosphère II 32°34′43.60″N 110°51′02.14″O / 32.5787778, 110.8505944 …   Wikipédia en Français

  • Biosphere II — Biosphère II 32°34′43.60″N 110°51′02.14″O / 32.5787778, 110.8505944 …   Wikipédia en Français

  • Biosphère 2 — Biosphère II 32°34′43.60″N 110°51′02.14″O / 32.5787778, 110.8505944 …   Wikipédia en Français

  • BIOSPHÈRE — L’idée de biosphère, ensemble des êtres vivants qui peuplent notre planète, remonte à J. B. Lamarck. C’est ensuite le géologue autrichien Suess qui, en 1875, a mis en parallèle le terme de biosphère avec ceux d’hydrosphère, d’atmosphère et de… …   Encyclopédie Universelle

  • Biosphere — Основная информация Полное имя Геир …   Википедия

  • Biosphere — Cet article concerne le musicien, pour les autres significations, voir biosphère Biosphere est le nom d artiste de Geir Jenssen (né en 1962 à Tromsø, Norvège), qui réalise des albums de musique électronique ambient. Sommaire 1 Discographie 1.1… …   Wikipédia en Français

  • Biosphere — (englisch für Biosphäre) steht für: Biosphere, Pseudonym für den Künstler Geir Jenssen Biosphere 2, ökologisches Experiment Biosphère (Montreal), ein Museum in Kanada Biosphere Conference, 1968 von der Unesco in Paris veranstaltet …   Deutsch Wikipedia

  • Biosphere — es el principal alias artístico de Geir Jenssen (Tromsø, 1962), músico noruego responsable de algunos de los discos más relevantes de música ambient. Su estilo es conocido como ambient techno o arctic ambient, y sus señas son el uso de loops y… …   Wikipedia Español

  • biosphere — 1899, on model of Ger. Biosphäre (1875), coined by German geologist Eduard Suess (1831 1914); see BIO (Cf. bio ) + SPHERE (Cf. sphere) …   Etymology dictionary

  • biosphere — ► NOUN ▪ the regions of the surface and atmosphere of the earth occupied by living organisms. DERIVATIVES biospheric adjective …   English terms dictionary

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”