hydrosphere

hydrosphere
/huy"dreuh sfear'/, n.
the water on or surrounding the surface of the globe, including the water of the oceans and the water in the atmosphere.
[1885-90; HYDRO-1 + -SPHERE]

* * *

Discontinuous layer of water at or near the Earth's surface.

It includes all liquid and frozen surface waters, groundwater held in soil and rock, and atmospheric water vapour. Virtually all of these waters are in constant circulation through the hydrologic cycle. Although the components of the hydrosphere are undergoing continuous change of state and location, the total water budget remains in balance. The components of the hydrosphere have been seriously affected by the water-polluting activities of modern society.

* * *

Introduction
 discontinuous layer of water at or near the Earth's surface. It includes all liquid and frozen surface waters, groundwater held in soil and rock, and atmospheric water vapour.

      Water is the most abundant substance at the surface of the Earth. About 1.4 billion cubic kilometres (326 million cubic miles) of water in liquid and frozen form make up the oceans, lakes, streams, glaciers, and groundwaters found there. It is this enormous volume of water, in its various manifestations, that forms the discontinuous layer, enclosing much of the terrestrial surface, known as the hydrosphere.

      Central to any discussion of the hydrosphere is the concept of the hydrologic cycle. This cycle consists of a group of reservoirs containing water, the processes by which water is transferred from one reservoir to another (or transformed from one state to another), and the rates of transfer associated with such processes. These transfer paths penetrate the entire hydrosphere, extending upward to about 15 kilometres (nine miles) in the Earth's atmosphere and downward to depths on the order of five kilometres in its crust.

      This article examines the processes of the hydrologic cycle and discusses the way in which the various reservoirs of the hydrosphere are related through the hydrologic cycle. It also describes the biogeochemical properties of the waters of the Earth at some length and considers the distribution of global water resources and their utilization and pollution by human society. Details concerning the major water environments that make up the hydrosphere are provided in the articles ocean, lake, river, and ice. See also climate for specific information about the impact of climatic factors on the hydrologic cycle. The principal concerns and methods of hydrology and its various allied disciplines are summarized in Earth sciences. (Earth sciences)

Distribution and quantity of the Earth's waters
       Water masses at the Earth's surface ocean waters and waters trapped in the pore spaces of sediments make up most of the present-day hydrosphere (see Table 1 (Water masses at the Earth's surface)). The total mass of water in the oceans equals about 50 percent of the mass of sedimentary rocks now in existence and about 5 percent of the mass of the Earth's crust as a whole. Deep and shallow groundwaters constitute a small percentage of the total water locked in the pores of sedimentary rocks—on the order of 3 to 15 percent. The amount of water in the atmosphere at any one time is trivial, equivalent to 0.013 × 106 cubic kilometres of liquid water, or about 0.001 percent of the total at the Earth's surface. This water, however, plays an important role in the water cycle.

      At present, ice locks up a little more than 1 percent of the Earth's water and may have accounted for as much as 3 percent or more during the height of the glaciations of the Pleistocene epoch (from 1,600,000 to 10,000 years ago). Although water storage in rivers, lakes, and the atmosphere is small, the rate of water circulation through the rain–river–ocean–atmosphere system is relatively rapid. The amount of water discharged each year into the oceans from the land is approximately equal to the total mass of water stored at any instant in rivers and lakes.

      Soil moisture accounts for only 0.005 percent of the water at the Earth's surface. It is this small amount of water, however, that exerts the most direct influence on evaporation from soils. The biosphere, though primarily H2O in composition, contains very little of the total water at the terrestrial surface, only about 0.00004 percent. Yet, the biosphere plays a major role in the transport of water vapour back into the atmosphere by the process of transpiration.

      As will be seen in the next section, the Earth's waters are not pure H2O but contain dissolved and particulate materials. Thus, the masses of water at the Earth's surface are major receptacles of inorganic and organic substances, and water movement plays a dominant role in the transportation of these substances about the planet's surface.

Biogeochemical properties of the hydrosphere

Rainwater (rain)
      About 110,300 cubic kilometres of rain fall on land each year. The total water in the atmosphere is 0.013 × 106 cubic kilometres, and this water, owing to precipitation and evaporation, turns over every 9.6 days. Rainwater is not pure but rather contains dissolved gases and salts, fine-ground particulate material, organic substances, and even bacteria. The sources of the materials in rainwater are the oceans, soils, fertilizers, air pollution, and fossil-fuel combustion.

      It has been observed that rains over oceanic islands and near coasts have ratios of major dissolved constituents very close to those found in seawater. The discovery of the high salt content of rain near coastlines was somewhat surprising because sea salts are not volatile, and it might be expected that the process of evaporation of water from the sea surface would “filter” out the salts. It has been demonstrated, however, that a large percentage of the salts in rain is derived from the bursting of small bubbles at the sea surface due to the impact of rain droplets or the breaking of waves, which results in the injection of sea aerosol into the atmosphere. This sea aerosol evaporates, with resultant precipitation of the salts as tiny particles that are subsequently carried high into the atmosphere by turbulent winds. These particles may then be transported over continents to fall in rain or as dry deposition.

      Assuming equilibrium with the atmospheric carbon dioxide partial pressure (PCO2) of 10−3.5 atmosphere, the approximate mean composition of rainwater is in parts per million (ppm): sodium (Na+), 1.98; potassium (K+), 0.30; magnesium (Mg2+), 0.27; calcium (Ca2+), 0.09; chloride (Cl), 3.79; sulfate (SO 2/4), 0.58; and bicarbonate (HCO3), 0.12. In addition to these ions, rainwater contains small amounts of dissolved silica—about 0.30 ppm. The average pH value of rainwater is 5.7. (The term pH is defined as the negative logarithm of the hydrogen ion concentration in moles per litre. The pH scale ranges from 0 to 14, with lower numbers indicating increased acidity.) On a global basis, as much as 35 percent of the sodium, 55 percent of the chlorine, 15 percent of the potassium, and 37 percent of the sulfate in river water may be derived from the oceans through sea aerosol generation.

      A considerable amount of data has become available for marine aerosols. These aerosols are important because (1) they are vital to any description of the global biogeochemical cycle of an element; (2) they may have an impact on climate; (3) they are a sink, via heterogeneous chemical reactions, for trace atmospheric gases; and (4) they influence precipitation of cloud and rain droplets. For many trace metals the ratio of the atmospheric flux to the riverine flux for coastal and remote oceanic areas may be greater than one, indicating the importance of atmospheric transport. Figures have been prepared that illustrate the enrichment factors (EF) of North Atlantic marine aerosols and suspended matter in North Atlantic waters relative to the crust, where

      and (X/Al)air and (X/Al)crust refer, respectively, to the ratio of the concentration of the element X to that of Al, aluminum, in the atmosphere and in average crustal material. The similarity in trend of enrichment factors for marine aerosols and suspended matter indicates qualitatively the importance of the marine aerosol to the composition of marine suspended matter and, consequently, to deep-sea sedimentation.

      In some instances the ratios of ions in rainwater deviate significantly from those in seawater. Mechanisms proposed for this fractionation are, for example, the escape of chlorine as gaseous hydrogen chloride (HCl) from sea-salt aerosol with a consequent enrichment in sodium and bubbling and thermal diffusion. In addition, release of gases like dimethyl sulfide (DMS) from the sea surface and its subsequent reaction in the oceanic atmosphere to sulfate can change rainwater ion ratios with respect to seawater. Soil particles also can influence rainwater composition. Rainfall over the southwestern United States contains relatively high sulfate concentrations because of sulfate-bearing particles that have been blown into the atmosphere from desert soils. Rain near industrial areas commonly contains high contents of sulfate, nitrate, and carbon dioxide (CO2) largely derived from the burning of coal and oil. There are two main processes leading to the conversion of sulfur dioxide (SO2) to sulfuric acid (H2SO4). These are reactions with hydroxyl radicals (OH) and with hydrogen peroxide (H2O2) in the atmosphere:

      and

      The sulfuric acid then dissociates to hydrogen and sulfate ions:

      For the nitrogen gases nitric oxide (NO) and nitrogen dioxide (NO2) released from fossil-fuel burning, their atmospheric reactions lead to the production of nitric acid (HNO3) and its dissociation to hydrogen ions (H+) and nitrate (NO3). These reactions are responsible for the acid rain conditions highly evident in the northeastern United States, southeastern Canada, and western Europe (see below Acid rain (hydrosphere)). The high sulfate values of the rain in the northeastern United States reflect the acid precipitation conditions of this region.

River and ocean waters
       river discharge constitutes the main source for the oceans. Seawater has a more uniform composition than river water. It contains, by weight, about 3.5 percent dissolved salts, whereas river water has only 0.012 percent. The average density of the world's oceans is roughly 2.75 percent greater than that of typical river water. Of the average 35 parts per thousand salts of seawater, sodium and chlorine make up almost 30 parts, and magnesium and sulfate contribute another four parts. Of the remaining one part of the salinity, calcium and potassium constitute 0.4 part each and carbon, as carbonate and bicarbonate, about 0.15 part. Thus, only eight elements (oxygen, sulfur, chlorine, sodium, magnesium, calcium, potassium, and carbon) make up 99 percent of seawater, though most of the 92 naturally occurring elements have been detected therein. Of importance are the nutrient elements phosphorus, nitrogen, and silicon, along with such essential micronutrient trace elements as iron, cobalt, and copper. These elements strongly regulate the organic production of the world's oceans.

      In contrast to ocean water, the average salinity of the world's rivers is low—only about 0.012 percent, or 120 ppm by weight. Of this salt content, carbon as bicarbonate constitutes 58 parts, or 48 percent, and calcium, sulfur as sulfate, and silicon as dissolved monomeric silicic acid make up a total of about 39 parts, or 33 percent. The remaining 19 percent consists predominantly of chlorine, sodium, and magnesium in descending importance. It is obvious that the concentrations and relative proportions of dissolved species in river waters contrast sharply with those of seawater. Thus, even though seawater is derived in part by the chemical differentiation and evaporation of river water, the processes involved affect every element differently, indicating that simple evaporation and concentration are entirely secondary to other processes.

Water-rock interactions as determining river water composition
      Generally speaking, the composition of river water, and thus that of lakes, is controlled by water–rock interactions. The attack of carbon dioxide-charged (carbon dioxide) rain and soil waters on the individual minerals in continental rocks leads to the production of dissolved constituents for lakes, rivers, and streams. It also gives rise to solid alteration products that make up soils or suspended particles in freshwater aquatic systems. The carbon dioxide content of rain and soil waters is of particular importance in weathering processes. The pH of rainwater equilibrated with the atmospheric carbon dioxide partial pressure of 10−3.5 atmosphere is 5.7. In industrial regions, rainwater pH values may be lower because of the release and subsequent hydrolysis of acid gases—namely, sulfur dioxide and nitrogen oxides (NOx) from the combustion of fossil fuels. After rainwater enters soils, its characteristics change markedly. The usual few parts per million of salts in rainwater increase substantially as the water reacts. The upper part of the soil is a zone of intense biochemical activity. The bacterial population near the surface is large, but it decreases rapidly downward with a steep gradient. One of the major biochemical processes of the bacteria is the oxidation of organic material, which leads to the release of carbon dioxide. Soil gases obtained above the zone of water saturation may contain 10 to 40 times as much carbon dioxide as the free atmosphere, and in some cases carbon dioxide has been shown to make up 30 percent of the soil gases as opposed to 0.03 percent of the free atmosphere. In addition to the acid effects of carbon dioxide, a highly acidic microenvironment is created by the roots (root) of living plants. Values of pH as low as 2 have been measured immediately adjacent to root hairs. Plants may have several kilometres of root hairs, and so their chemical effects are formidable.

Congruent and incongruent weathering reactions
      These acid solutions in the soil environment attack the rock minerals, the bases of the system, producing neutralization products of dissolved constituents and solid particles. Two general types of reaction occur: congruent and incongruent. In the former, a solid dissolves, adding elements to the water in their proportions in the mineral. An example of such a weathering reaction is the solution of calcite (CaCO3) in limestones: (limestone)

      Here, one of the HCO3 ions comes from calcite and the other from CO2(g) in the reacting water. The amount of carbon dioxide dissolved according to reaction (4) depends on temperature, pressure, original bicarbonate content of the weathering solution, and the partial pressure of the carbon dioxide. The carbon dioxide and the temperature are the most important variables. Increases in one or both of these variables lead to increases in the amount of calcite dissolved. For example, for a carbon dioxide pressure of 10−3.5 atmosphere, the amount of calcium that can be dissolved until saturation is about 10−3.3 mole, or 20 ppm, at 25° C (77° F). For an atmospheric carbon dioxide pressure of 10−2 atmosphere and for a soil atmosphere of nearly pure carbon dioxide, the values are 65 and 300 ppm, respectively. The weathering of calcite leads to the release of calcium and bicarbonate ions into soil waters and groundwaters, and these constituents eventually reach lake and river systems. The insoluble residue of quartz (SiO2), clay minerals (cation, aluminum, silicon, oxygen, hydrogen phases), and iron oxides (e.g., FeOOH) in the limestone rock make up the deep-red soils that form from limestone weathering. These particles may be carried into streams by runoff and hence to lakes and the oceans and become part of the suspended load of these systems.

      An example of an incongruent weathering reaction is that involving aluminosilicates. One such reaction is the aggressive attack of carbon dioxide-charged soil water on the mineral K-spar (KAlSi3O8), an important phase found in continental rocks. The reaction is

      It should be noted that the K-spar changes into a new mineral—kaolinite (a clay mineral) in this case—plus solution, and acid is consumed. The total dissolved material per litre of soil solution released is about 60 ppm for a solution initially equilibrated with a typical soil carbon dioxide content. The water resulting from reaction (5) would contain bicarbonate, potassium, and dissolved silica in the ratios 1:1:2, and the new solid, kaolinite, would be a weathering product. These dissolved constituents and the solid alteration product would eventually reach rivers to be transferred possibly to lakes and ultimately to the sea. It has been demonstrated that the composition of river water is the product of a variety of mineral–water reactions such as (4) and (5). The dissolved load of the world's rivers comes from the following sources: 7 percent from beds of halite (NaCl) and salt disseminated in rocks; 10 percent from gypsum (CaSO4·2H2O) and anhydrite (CaSO4) deposits and sulfate salts disseminated in rocks; 38 percent from limestones and dolomites; and 45 percent from the weathering of one silicate mineral to another. Of the bicarbonate ions in river water, 56 percent stems from the atmosphere, 35 percent from carbonate minerals, and 9 percent from the oxidative weathering of fossil organic matter. Reactions involving silicate minerals account for 30 percent of the riverine bicarbonate ions.

      Besides dissolved substances, rivers also transport solids in traction (i.e., bed load) and, most importantly, suspended load. The present global river-borne flux of solids to the oceans is estimated as 155 × 1014 grams per year. Most of this flux comes from Southeast Asian rivers. The composition of this suspended material resembles soils and shales and is dominated by silicon and aluminum. Present elemental fluxes are estimated in 1012 grams per year as silicon, 4,420; aluminum, 1,460; iron, 740; calcium, 330; potassium, 310; magnesium, 210; and sodium, 110. The total load of particulate organic carbon of the world's rivers is 180 × 1012 grams per year. The riverine fluxes of trace metals to the oceans are dominated by their occurrence in the particulate phase as opposed to the dissolved phase. The particulate matter in river water is an important source of silicon, aluminum, ion, titanium, rubidium, scandium, vanadium, the lanthanides, and other elements for deep-sea sediments.

lake waters
      Although lake waters constitute only a small percentage of the water in the hydrosphere, they are an important ephemeral storage reservoir for fresh water. Aside from their recreational use, lakes constitute a source of water for household, agricultural, and industrial uses. Lake waters are also very susceptible to changes in chemical composition due to these uses and to other factors.

      In general, fresh waters at the continental surface evolve from their rock sources by enrichment in calcium and sodium and by depletion in magnesium and potassium. In very soft waters the alkalies may be more abundant than the alkaline earths, and in the more concentrated waters of open river systems Ca > Mg > Na > K. For the anions, in general, HCO3 exceeds SO 2/4, which is greater in concentration than Cl. It is worthwhile at this stage to consider some major mechanisms that control global surface water composition. These mechanisms are atmospheric precipitation, rock reactions, and evaporation-precipitation.

      The mechanism principally responsible for waters of very low salinity is precipitation. These waters tend to form in tropical regions of low relief and thoroughly leached source rocks. In these regions rainfall is high, and water compositions are usually dominated by salts brought in by precipitation. Such waters constitute one end-member of a series of water compositions for which the other end-member represents water compositions dominated by contributions of dissolved salts from the rocks and soils of their basins. These waters have moderate salinity and are rich in dissolved calcium and bicarbonate. They are, in turn, the end-member of another series that extends from the calcium-rich, medium-salinity fresh waters to the high-salinity, sodium chloride-dominated waters of which seawater is an example. Seawater composition, however, does not evolve directly from the composition of fresh waters and the precipitation of calcium carbonate; other mechanisms that control its composition are involved. Such factors as relief and vegetation also may affect the composition of the world's surface waters, but atmospheric precipitation, water–rock reactions, and evaporation–crystallization processes appear to be the dominant mechanisms governing continental surface water chemistry.

   Chemistry of representative closed-basin, saline lake watersContinental fresh waters evaporate once they have entered closed basins, and their constituent salts precipitate on the basin floors. The composition of these waters may evolve along several different paths, depending on their initial chemical makeup. Table 3 (Chemistry of representative closed-basin, saline lake waters) shows a number of brine compositions from North American saline lakes, and Figure 1—> illustrates possible evaporation paths that led to these compositions. For example, Surprise Valley Lake, California, is a body of sodium chloride water that may have evolved from what was initially fresh water which precipitated calcite during evaporation, resulting in water with a dissolved-calcium concentration to alkalinity ratio of less than 2. Further evaporation of such water and precipitation of sepiolite [MgSi3O6(OH)2] would give rise to water enriched in alkalinity with respect to dissolved magnesium and hence to sodium chloride–bicarbonate–carbonate water such as that of Surprise Valley Lake. The paths shown in Figure 1—> are an oversimplification of the actual processes, which are far more diverse and complex in nature, but they do provide a reasonable picture of how fresh waters evolve into saline lake waters in closed basins.

      Biological processes strongly affect the composition of lake waters and are responsible to a significant degree for the compositional differences between the upper water layer (the epilimnion) and the lower water layer (the hypolimnion) of lakes. The starting point is photosynthesis, represented by the following reaction:

      The reversal of this reaction is oxidation-respiration leading to the release of the nutrients nitrogen and phosphorus, as well as carbon dioxide. In a stratified lake, carbon, nutrients, and silica are extracted from the epilimnion during photosynthesis. This process leads to reduced concentrations of nitrate, phosphate, and silica in these waters and, during times of maximum daylight organic production, supersaturation of epilimnion waters with respect to dissolved oxygen. The organic matter produced by phytoplankton may be either grazed upon by zooplankton and other organisms or decomposed by bacteria. Some of it, however, sinks into the hypolimnion. There, it is further decomposed, especially by bacteria, resulting in the release of dissolved phosphorus and nitrogen and the consumption of oxygen. Oxygen concentrations therefore are reduced in these lower lake waters, because stratification prevents oxygen exchange with the atmosphere. Furthermore, the inorganic carbonate and siliceous skeletons of the dead organisms sinking into the hypolimnion may dissolve, giving rise to increased concentrations of dissolved silica and inorganic carbon in the deep waters of stratified lakes. This dissolution is a result of undersaturation of the hypolimnion waters with respect to the opaline silica and calcium carbonate that make up the skeletons of the dead and sinking plankton. These natural biological processes have been accelerated in some lakes because of excess nutrient input by human activity, resulting in the eutrophication of lake waters (and marine systems; see below Eutrophication (hydrosphere)).

Groundwaters (groundwater)
  Processes affecting the major chemical components of groundwaterThese waters derive their compositions from a variety of processes, including dissolution, hydrolysis, and precipitation reactions; adsorption and ion exchange; oxidation and reduction; gas exchange between groundwater and the atmosphere; and biological processes (see Table 4 (Processes affecting the major chemical components of groundwater)). The biological processes of greatest importance are microbial metabolism, organic production, and respiration (oxidation). By far the most important overall process for the major constituents of groundwater is that of mineral–water reactions, which were briefly described above in River and ocean waters (hydrosphere). Thus, the composition of groundwaters strongly reflects the types of rock minerals that the waters have encountered in their movement through the subsurface. Table 5—> shows the waters found in limestones, crystalline rocks, and two types of fine-grained rocks. The mineralogy of the Wissahickon schist is dominated by aluminosilicate compositions, whereas the Ecca shale contains significant carbonate and dispersed salts. The latter minerals are more soluble than aluminosilicate minerals, and their dissolution gives rise to the high salinity of Ecca shale waters. The waters of the Wissahickon schist have low salinity partly because of the low chemical reactivity of the silicate minerals in this rock. In contrast, the Miocene limestone waters are dominated by dissolved calcium and bicarbonate, a characteristic reflecting the higher solubility and rate of dissolution of the calcite that makes up this rock. The groundwater from the granite is rich in the Ca2+ and Na+ cations derived from dissolution of the plagioclase feldspar (a sodium, calcium aluminosilicate), which is a major mineral found in this rock type. In general, the most mobile elements in groundwater—i.e., those most easily liberated by the weathering of rock minerals—are calcium, sodium, and magnesium. Silicon and potassium have intermediate mobilities, and aluminum and iron are essentially immobile and locked up in solid phases.

      Groundwaters are highly susceptible to contamination because of human activities and the fact that their dissolved constituents are derived to a large extent from the leaching of surface materials. Some of the nitrogen and phosphorus applied to soils as fertilizers and organic pesticides may be leached and leak into groundwater systems, leading to increased concentrations of ammonium and phosphate. Radioactive wastes, industrial chemicals, household materials, and mine refuse are other anthropogenic sources of dissolved substances that have been detected in groundwater systems.

      Ice is nearly a pure solid and, as such, accommodates few foreign ions in its structure. It does contain, however, particulate matter and gases, which are trapped in bubbles within the ice. The change in composition of these materials through time, as recorded in the successive layers of ice, has been used to interpret the history of the Earth's surface environment and the impact of human activities on this environment. The increase in the lead content of continental glacial ice with decreasing age of the ice, for example, reflects the progressive input of tetraethyl lead into the global environment from gasoline burning. Also, atmospheric carbon dioxide and methane concentrations, which have increased significantly during the past century because of anthropogenic activities, are faithfully recorded in ice bubbles of the thick continental ice sheets. Present-day atmospheric carbon dioxide and methane concentrations are 25 percent and 167 percent, respectively, higher than their concentrations 200 years ago; the latter concentration values were obtained from measurements of the gases in air trapped in ice.

The hydrologic cycle

General nature of the cycle
 The present-day hydrologic cycle at the Earth's surface is illustrated in Figure 2—>. Some 496,000 cubic kilometres of water evaporate (vaporization) from the land and ocean surface annually, remaining for about 10 days in the atmosphere before falling as rain or snow. The amount of solar radiation necessary to evaporate this water is half of the total solar radiation received at the Earth's surface. About one-third of the precipitation falling on land runs off to the oceans primarily in rivers, while direct groundwater discharge to the oceans accounts for only about 0.6 percent of the total discharge. A small amount of precipitation is temporarily stored in the waters of rivers and lakes. The remaining precipitation over land, 0.073 × 106 cubic kilometres per year, returns to the atmosphere by evaporation. Over the oceans, evaporation exceeds precipitation, and the net difference represents transport of water vapour over land, where it precipitates as rain and returns to the oceans as river runoff and direct groundwater discharge.

      The various reservoirs (reservoir) in the hydrologic cycle have different water residence times. Residence time is defined as the amount of water in a reservoir divided by either the rate of addition of water to the reservoir or the rate of loss from it. The oceans have a water residence time of 37,000 years; this long residence time reflects the large amount of water in the oceans. In the atmosphere the residence time of water vapour relative to total evaporation is only 10 days. Lakes, rivers, ice, and groundwaters have residence times lying between these two extremes and are highly variable.

  There is considerable variation in evaporation and precipitation over the globe. In order to have precipitation, there must be sufficient atmospheric water vapour and enough rising air to carry the vapour to an altitude where it can condense and precipitate. Figure 3—> shows the latitudinal variation of precipitation and evaporation and their gross relation to the global wind belts. The trade winds (trade wind), for example, are initially cool, but they warm up as they blow toward the equator. These winds pick up moisture from the ocean, increasing ocean surface salinity and causing seawater to sink. When the trade winds reach the equator, they rise, and the water vapour in them condenses and forms clouds. Net precipitation is high near the equator and also in the belts of the prevailing westerlies where there is frequent storm activity. Evaporation exceeds precipitation in the subtropics where the air is stable and near the poles where the air is both stable and has a low water vapour content because of the cold. The Greenland and Antarctic ice sheets formed due to the very low evaporation rates at the poles that result in precipitation exceeding evaporation in these local regions. The strong link between wind belts, the water balance, and the salinity of ocean water is apparent in Figure 3—>.

Processes involved in the cycle
 The hydrologic cycle consists of various complicated processes of precipitation, evaporation, interception, transpiration, infiltration, percolation, retention, detention, overland flow, throughflow, and runoff. Figure 4—> provides a schematic representation of the details of the hydrologic cycle, illustrating the processes involved. In the following sections, some of these processes are discussed in detail.

Water vapour and precipitation
      As noted above, water exists in the atmosphere in gaseous form. Its liquid form, either as water droplets in clouds or as rain, and its solid form, as ice crystals in clouds, snowflakes, or hail, occur only momentarily and locally.

      Water vapour performs two major functions: (1) it is important to the radiation balance of the Earth as its presence keeps the planetary surface warmer than would otherwise be the case; and (2) it is the principal phase of the ascending part of the hydrologic cycle.

      The mass of water vapour in the atmosphere, which represents only 0.001 percent of the hydrosphere, is highest in the tropics and decreases toward the poles. At a mean temperature of the Earth's surface of 15° C, the partial pressure of water vapour at equilibrium with pure water is 0.017 atmosphere. The addition of salts to pure water lowers its vapour pressure. The equilibrium, or saturation, water vapour pressure of a saturated solution of sodium chloride is 22 percent lower than that of pure water. Precipitable water vapour has, on the average, a vapour pressure of 0.0025 atmosphere, which amounts to 15 percent of the saturation vapour pressure. The ratio of observed water vapour pressure to the saturation vapour pressure at the same temperature is the relative humidity of the air. Thus, the mean relative humidity of the atmosphere is only 15 percent, a value that is low in human and biological terms. The relative humidity of the air, however, varies greatly from one geographic region to another and also vertically in the atmosphere. Atmospheric water vapour decreases rapidly with increasing altitude relative to its surface value. The amount of water required to saturate a volume of air depends on the temperature of the air. Air at high temperature can hold more water vapour at saturation than can air at low temperature. Because the temperature of the lower atmosphere (the troposphere) decreases rapidly with increasing altitude to about 15 kilometres, the upper levels of the troposphere contain little water vapour; most of the vapour is found within a few kilometres of the Earth's surface. The average relative humidity of tropospheric air is about 50 percent. Above 15 kilometres, water vapour is essentially frozen out of the atmosphere, amounting to less than 0.1 percent of its concentration at the Earth's surface.

      Aside from temperature, other factors determine the water vapour content of the air and are particularly important in the lower troposphere. These factors include local evaporation and the horizontal atmospheric transportation of moisture, which varies with altitude, latitude, season, and topography. During a period of 10 days (i.e., the residence time of water in the atmosphere), horizontal eddy turbulence may disperse vapour over distances up to 1,000 kilometres.

      When a mass of air at the Earth's surface is exposed to a body of water, it gains water by evaporation or loses water by precipitation, depending on its relative humidity. If the air is undersaturated, with a relative humidity of less than 100 percent, it gains water vapour because the rate of evaporation exceeds the rate of condensation. If the air is supersaturated, with a relative humidity greater than 100 percent, the air mass loses water vapour because the rate of precipitation exceeds that of evaporation. This interaction between air masses and surface water bodies drives the atmosphere toward a state of saturation, which is not achieved for the entire atmosphere because of the variability in weather and because not all air masses are in contact with water bodies. In general, the level of atmospheric water vapour is higher in the summer, since temperatures are higher at this time of year. Also, atmospheric water vapour content is higher near the source of moisture than in distant regions. Over the oceans, the air is almost always near saturation, whereas over the deserts, where the supply of moisture is limited, the air is far below water vapour saturation values. In most cases, atmospheric water vapour content decreases inland over continents, but this decrease is modified by rainfall conditions, by the presence or absence of high mountains, large lakes, extensive forests, and swamps, and by the prevailing wind directions. Horizontal winds and air mass movements transfer water vapour from the ocean to the land. Although the processes are not completely separable, the horizontal transfer of water vapour seldom causes the vapour to undergo condensation, whereas vertical movements are most important in the condensation process.

      Condensation depends strongly on the average temperature of the Earth's surface because the water vapour content of the air is strongly dependent on temperature. In figures that show the states of water as a function of the variables of pressure and temperature, the slope of the phase boundary between liquid water and water vapour is positive, implying that with increasing temperature the air at equilibrium will hold increasing amounts of water vapour. Cooling or mixing of this air results in condensation of the vapour and precipitation as water droplets or as ice crystals if the air temperature is below 0° C. When first formed, the water droplets or ice crystals are very small, on the order of 10−2 to 10−3 centimetre in diameter, and they float freely in the atmosphere. In large quantities, these water droplets and ice crystals produce a cloud. All clouds are formed as a result of cooling below the dew point, the temperature at which condensation begins when air is cooled at constant pressure and constant water vapour content. When the droplets or crystals coalesce to a size of about 10−2 centimetre in diameter, the drops become heavy enough to fall as raindrops or snowflakes. Hailstones measure about 10−1 centimetre in diameter or much larger. Water vapour condensing in the atmosphere contains strongly soluble salts (mostly of oceanic origin), weakly soluble or insoluble solids (dust), and dissolved gases. The dust and sea-salt aerosol particles in the air may act as sites of condensation by serving as nuclei for bringing initially a few water molecules together and inducing condensation from supersaturated air.

Distribution of precipitation
      Precipitation falling toward the Earth's surface may suffer several fates. It may be evaporated during its fall or after it reaches the ground surface. If the surface is covered with dense vegetation, much of the precipitation may be held on leaves and plant limbs and stems. This process is termed interception and may result in little water reaching the ground because the water may be directly evaporated from plant surfaces back into the atmosphere. If precipitation reaches the ground in the form of snow, it may remain there for some time. On the other hand, if precipitation falls as rain, it may evaporate, infiltrate the soil, be detained in small catchment areas, or become overland flow—a form of runoff. Overland flow (Ro) may be expressed in terms of intensity units, water depth per unit of time (e.g., centimetres per hour, or inches per hour), as

      where P is precipitation rate and I is infiltration rate (rate of entry and downward movement of water into the soil profile). Infiltration rate will equal precipitation rate until the limit of the infiltration rate, or infiltration capacity, is reached. Soil infiltration rates are usually high at the beginning of a rain preceded by a dry spell and decrease as the rainfall continues. This change in rate is due to the clogging of soil pores by particles brought from above by the infiltrating rain and to the swelling of colloidal soil particles as they absorb water. Thus, rapid decreases in infiltration rates during a rain are more likely to occur in clay-rich soils than in sandy soils.

      Between rainfall periods, water held in the soil as soil moisture is gradually lost by direct evaporation or by withdrawal by plants. Evaporation into the open atmosphere occurs at the surface of the soil, and the soil dries progressively downward with time. Water vapour in the soil diffuses upward, replenishing the evaporated water, and in turn is evaporated. The pumping of air in and out of the soil by atmospheric pressure changes enhances the movement of soil moisture upward. It has been shown that evaporation of a water droplet in the free atmosphere, and to a first approximation in various soil atmospheres, is proportional to the droplet surface area 4πr2 (square centimetres), the diffusional flux of water at the droplet surface, and the transfer of heat as the droplet evaporates. The equation for the rate of shrinkage of a water droplet due to evaporation is

      where dr/dt is the rate of change in the radius of the water droplet (centimetres per second), D is the diffusion coefficient of water vapour in air (cubic centimetres per second), ρvo is the equilibrium vapour concentration at the droplet surface, Sp is the degree of undersaturation of water vapour in the environment, r is the radius of the droplet (centimetres), ρL is the density of liquid water (grams per cubic centimetre), and X is a dimensionless parameter depending on D, ρvo, temperature, the heat of evaporation of water vapour, the coefficient of thermal conductivity of air, and the spherical coordinate system necessary to define processes occurring to a spherical water droplet. Water droplets shrink—dr/dt < 0, evaporate—when the water vapour concentration in the environment (atmosphere or soil atmosphere) is less than the saturation water vapour concentration at the droplet surface. They grow—dr/dt > 0, condense—when the converse is true in the free atmosphere. The term dr/dt has negative values for evaporation and positive ones for condensation. Use of this equation shows, as an example, that it would take 23 minutes for a water droplet to shrink (evaporate) in size from 50 to five micrometres in air at 10° C and a water vapour undersaturation of 1 percent.

      Besides simple evaporation of water from soils, water is also returned to the atmosphere by transpiration in plants. Plants draw water from soil moisture through their vast network of root hairs and rootlets. This water is carried upward through the plant trunk and branches into the leaves, where it is discharged as water vapour. The term evapotranspiration is used in climatic and hydrologic studies to include the combined water loss from the Earth's surface resulting from evaporation and transpiration. The maximum possible evapotranspiration is termed potential evapotranspiration and is governed by the available heat energy. It is taken as equal to evaporation from a large water surface and is generally much less than actual evapotranspiration. Actual evapotranspiration is never greater than precipitation except on irrigated land because of percolation of water into groundwater bodies and surface runoff.

      The soil-moisture zone gains water by precipitation and infiltration and loses water by evapotranspiration, overland flow, and percolation of water downward due to gravity into the groundwater zone. The contact between the groundwater zone (phreatic zone) and the overlying unsaturated zone ( vadose zone) is called the groundwater table (water table). The water balance equation for change of soil-moisture storage in a soil is given as

      where S is storage, P is precipitation, E is evaporation, and R is surface runoff plus percolation rate into the groundwater zone; all terms are in units of length per unit of time (e.g., millimetres per day, centimetres per month). In humid, midlatitude climates where a strong contrast between winter and summer temperatures exists, there is an annual cycle of the water content of soils. The annual cycle of moisture in soil in Ohio, U.S., demonstrates the processes controlling soil moisture. Of special importance is the fact that the soils are saturated in this temperate climate in the spring, and the evaporation rate is low because of the low input of radiant energy from the Sun. By contrast, in the summer, evaporation increases because of increasing solar radiation, and with the growth of plants so does transpiration. Soil moisture is reduced to very low levels at this time of year.

Groundwaters and river runoff
      This section focuses on term R in equation (9—>) representing groundwater and river runoff losses from the soil-moisture zone. Water percolates from the soil-moisture zone through the unsaturated (vadose) zone to the water table. Flow through the unsaturated zone is complicated. After a rainfall, water may form nearly a continuous phase in pores in this zone, but with drying the last amount of water is held in wedges at points of contact of solid grains and as thin films on solid surfaces. The flow paths of water become more tortuous, and the water-conducting properties decrease rapidly. Structured soils and fractured rock in the vadose zone may act as conduits for fluids to reach the water table. Because of the complex geometry of water contained in the unsaturated zone, the properties of water are expressed by means of empirical relationships. Darcy's law, derived in 1856 from experimentation by the French engineer Henri Darcy, permits quantification of water flow through porous media. The law states that the rate of flow Q of a fluid through a porous layer of medium (e.g., a sand bed) is directly proportional to the area A of the layer and to the difference Δh between the fluid heads at the inlet and outlet faces of the layer and inversely proportional to the thickness L of the layer, or, expressed analytically,

      where K is a constant characteristic of the medium. The term K for a porous rock medium is the volume of fluid of unit viscosity passing through a unit cross section of the rock in unit time under the action of a unit pressure gradient and is called permeability. The permeability of a rock is dependent on the geometric properties of the rock, such as porosity, shape and size distribution of constituent rock grains, and degree of cementation of the rock. Permeabilities of rocks vary greatly. Unconsolidated sands (sand) may have permeabilities measured in hundreds of darcys, whereas consolidated sands that will transmit reasonable amounts of fluid have permeabilities of 0.01 to one darcy. A rough idea of the meaning of one darcy of permeability (which equals 9.869 × 10−12 square metre) can be obtained by imagining a cube of sand one foot on a side. If the sand has a permeability of one darcy, approximately one barrel of water per day will pass through the one-foot cube with a one-pound pressure head. The general equation of Darcy can be modified to express flow in both the unsaturated zone and the saturated groundwater zone.

      Groundwater is constantly in motion. When a lake or stream intersects the groundwater table, groundwater communicates directly with these bodies of water. If the groundwater table is higher than the stream or lake level, a pressure head will develop such that the groundwater flows into the water body; conversely, if the groundwater table is lower than the river or lake level, the pressure gradient induces flow into the groundwater. Most groundwater ultimately reaches the channels (streambed) of surface streams and rivers and flows to the sea. On the average, groundwater contributes to total river runoff about 30 percent of its water on a global basis.

      Water runoff from the land surface is that part of precipitation which eventually appears in perennial or intermittent surface streams. Streamflow-generation mechanisms have been studied for several decades, and there is now considerable knowledge regarding rainfall-runoff processes and their controls. This understanding is the result of both careful observations from field experiments and the heuristic simulations of hypothetical realities with rigorous mathematical models. The discharge measured at the downstream end of a channel reach is supplied by channel inflow at the upstream end of the reach and by the lateral inflows that enter the channel from the hillslope along the reach. The lateral inflows may arrive at the stream in one of three forms: (1) groundwater flow, (2) subsurface storm flow, or (3) overland flow.

      Groundwater flow provides the base flow component of streams that sustains their flow between storms. The “flashy” response of streamflow to individual precipitation events may be ascribed to either subsurface storm flow or overland flow. Subsurface storm flow can be a dominant streamflow-generation mechanism only when the impeding subsoil horizon laterally diverts infiltrating water downslope. Under intense rainfall events during which the surface soil layer becomes saturated to some depth, water is able to migrate through “preferred pathways” rapidly enough to deliver contributions to the stream during the peak runoff period. The conditions for subsurface storm flow are quite restrictive. The mechanism is most likely to be operative on steep, humid, forested hillslopes with very permeable surface soils.

      Overland flow is generated at a point on a hillslope only after surface ponding takes place. Ponding cannot occur until the surface soil layers become saturated. It is now widely recognized that surface saturation can occur because of two quite distinct mechanisms—namely, Horton overland flow and Dunne overland flow.

      The former classic mechanism is for a precipitation rate that exceeds the saturated hydraulic conductivity of the surface soil. A moisture content versus depth profile during such a rainfall event will show moisture contents that increase at the surface as a function of time. At some point in time the surface becomes saturated, and an inverted zone of saturation begins to propagate downward into the soil. It is at this time that the infiltration rate drops below the rainfall rate and overland flow is generated. The time is called the ponding time. The necessary conditions for the generation of overland flow by the Horton mechanism are (1) a rainfall rate greater than the saturated hydraulic conductivity of the soil and (2) a rainfall duration longer than the required ponding time for a given initial moisture profile. Horton overland flow is generated from partial areas of the hillslope where surface hydraulic conductivities are lowest.

      In Dunne overland flow, the precipitation rate is less than the saturated hydraulic conductivity, and the initial water table is shallow or there is a shallow impeding layer. Surface saturation occurs because of a rising water table; ponding and overland flow occur at a time when no further soil-moisture storage is available. The Dunne mechanism is more common to near-channel areas. Dunne overland flow is generated from partial areas of the hillslope where water tables are shallowest. Both Horton and Dunne mechanisms result in variable source areas that expand and contract through wet and dry periods.

      Total river discharge and the chemistry of the discharge vary from continent to continent; some continents are wetter and some drier than the world average, but the deviations are not extreme. The runoff per unit area from Asia and Europe is almost exactly equal to the world average; it is a little lower in Africa and North America; and it is considerably higher in South America. Antarctica is frozen and Australia is arid, and so they contribute little runoff. Also, since their areas are relatively small, they do not affect the global runoff average significantly. The waters draining the continents have quite different chemistries; those from Europe are very rich in calcium and bicarbonates, whereas those from Africa and South America are not. North American and Asian rivers are somewhat intermediate in their concentrations of these dissolved constituents. Such differences in composition reflect a variety of factors, including runoff, temperature, and relief, but certainly the bulk composition of the continental rocks in contact with these waters and their underground sources play a major role. The surface rocks of Europe are rich in carbonates and those of South America are not; the latter are dominated by sediments rich in silicate minerals.

      The chemistry of groundwater and river runoff is being modified by human activities on a global scale. The natural dissolved riverine input of major constituents to the oceans already has been increased by more than 10 percent because of human activities. In the case of sodium, chlorine, and sulfate, the increases are as high as 30 percent. In the United States alone, total water utilization is equivalent to one-third of total runoff, with about 2 percent of the water used coming from underground wells. In the southwestern region of the country, water supplies have been tapped heavily and in some areas have been exhausted with no hope of replacement. This extensive utilization of fresh waters in the United States and throughout the globe make them particularly susceptible to pollution. Leachates from fertilizers, herbicides, and pesticides are found in some freshwater bodies; toxic and inorganic or organic chemicals are present; radioactive elements have been detected; and some surface-water bodies have had their salinities increased dramatically, rendering them useless for human consumption. It is therefore imperative that nations closely monitor the utilization of freshwater systems and promote their conservation.

Origin and evolution of the hydrosphere
      It is not very likely that the total amount of water at the Earth's surface has changed significantly over geologic time. Based on the ages of meteorites, the Earth is thought to be 4.6 billion years old. The oldest rocks known date 3.8 billion years in age, and these rocks, though altered by post-depositional processes, show signs of having been deposited in an environment containing water. There is no direct evidence for water for the period between 4.6 and 3.8 billion years ago. Thus, ideas concerning the early history of the hydrosphere are closely linked to theories about the origin of the Earth.

      The Earth is thought to have accreted from a cloud of ionized particles around the Sun. This gaseous matter condensed into small particles that coalesced to form a protoplanet, which in turn grew by the gravitational attraction of more particulates. Some of these particles had compositions similar to that of carbonaceous chondrite meteorites, which may contain up to 20 percent water. Heating of this initially cool, unsorted conglomerate by the decay of radioactive elements and the conversion of kinetic and potential energy to heat resulted in the development of the Earth's liquid iron core and the gross internal zonation of the planet (i.e., differentiation into core, mantle, and crust). It has been concluded that the Earth's core formed over a period of about 500 million years. It is likely that core formation resulted in the escape of an original primitive atmosphere and its replacement by one derived from the loss of volatile substances from the planetary interior.

      At an early stage the Earth thus did not have water or water vapour at its surface. Once the planet's surface had cooled sufficiently, water contained in the minerals of the accreted material and released at depth could escape to the surface and, instead of being lost to space, cooled and condensed to form the initial hydrosphere. A large, cool Earth most certainly served as a better trap for water than a small, hot body because the lower the temperature, the less likelihood for water vapour to escape, and the larger the Earth, the stronger its gravitational attraction for water vapour. Whether most of the degassing took place during core formation or shortly thereafter or whether there has been significant degassing of the Earth's interior throughout geologic time remains uncertain. It is likely that the hydrosphere attained its present volume early in the Earth's history, and since that time there have been only small losses and gains. Gains would be from continuous degassing of the Earth; the present degassing rate of juvenile water has been determined as being only 0.3 cubic kilometre per year. Water loss in the upper atmosphere is by photodissociation, the breakup of water vapour molecules into hydrogen and oxygen due to the energy of ultraviolet light. The hydrogen is lost to space and the oxygen remains behind. Only about 4.8 × 10−4 cubic kilometre of water vapour is presently destroyed each year by photodissociation. This low rate can be readily explained: the very cold temperatures of the upper atmosphere result in a cold trap at an altitude of about 15 kilometres, where most of the water vapour condenses and returns to lower altitudes, thereby escaping photodissociation. Since the early formation of the hydrosphere, the amount of water vapour in the atmosphere has been regulated by the temperature of the Earth's surface—hence its radiation balance. Higher temperatures imply higher concentrations of atmospheric water vapour, while lower temperatures suggest lower atmospheric levels.

The early hydrosphere
       Estimate of excess volatilesThe gases released from the Earth during its early history, including water vapour, have been called excess volatiles because their masses cannot be accounted for simply by rock weathering. An estimate of the excess volatiles is given in Table 6 (Estimate of excess volatiles). These volatiles are thought to have formed the early atmosphere of the Earth. At an initial crustal temperature of about 600° C, almost all of these compounds, including H2O, would have been in the atmosphere. The sequence of events that occurred as the crust cooled is difficult to reconstruct. Below 100° C all of the water would have condensed, and the acid gases would have reacted with the original igneous crustal minerals to form sediments and an initial hydrosphere that was dominated by a salty ocean. If the reaction rates are assumed to have been slow relative to cooling, an atmosphere of 600° C would have contained, together with other compounds, water vapour, carbon dioxide, and hydrogen chloride (HCl) in a ratio of 20:3:1 and cooled to the critical temperature of water (i.e., 374° C). The water therefore would have condensed into an early hot ocean. At this stage, the hydrogen chloride would have dissolved in the ocean (about one mole per litre), but most of the carbon dioxide would have remained in the atmosphere, with only about 0.5 mole per litre in the ocean water. This early acid ocean would have reacted vigorously with crustal minerals, dissolving out silica and cations and creating a residue composed principally of aluminous clay minerals that would form the sediments of the early ocean basins.

      This is one of several possible pathways for the early surface of the Earth. Whatever the actual case, after the Earth's surface had cooled to 100° C, it would have taken only a short time for the remaining acid gases to be consumed in reactions involving igneous rock minerals. The presence of cyanobacteria (e.g., blue-green algae) in the fossil record of rocks older than three billion years attests to the fact that the Earth's surface had cooled to temperatures lower than 100° C by this time, and neutralization of the original acid volatiles had taken place. It is possible, however, that, because of increased greenhouse gas concentrations (see below) in the Early Archean era (about 3.8 to 3.4 billion years ago), the Earth's surface could still have been warmer than today.

      If most of the degassing of primary volatile substances from the Earth's interior occurred early, the chloride released by the reaction of hydrochloric acid with rock minerals would be found in the oceans or in evaporite deposits, and the oceans would have a salinity and volume comparable to that of today. This conclusion is based on the assumption that there has been no drastic change in the ratios of volatiles released through geologic time. The overall generalized reaction indicative of the chemistry leading to the formation of the early oceans can be written in the form: primary igneous rock minerals + acid volatiles + H2O → sedimentary rocks + oceans + atmosphere. It should be noted from this equation that, if all the acid volatiles and H2O were released early in the history of the Earth and in the proportions found today, then the total original sedimentary rock mass-produced would be equal to that of the present, and ocean salinity and volume would be close to those of today as well. If, on the other hand, degassing were linear with time, then the sedimentary rock mass would have accumulated at a linear rate, as would have oceanic volume. The salinity of the oceans, however, would remain nearly the same if the ratios of volatiles degassed did not change with time. The most likely situation is the one presented here—namely, that major degassing occurred early in Earth's history, after which minor amounts of volatiles were released episodically or continuously for the remainder of geologic time. The salt content of the oceans based on the constant proportions of volatiles released would depend primarily on the ratio of sodium chloride locked up in evaporites to that dissolved in the oceans. If all the sodium chloride in evaporites were added to the oceans today, the salinity would be approximately doubled. This value gives a sense of the maximum salinity that the oceans could have attained throughout geologic time.

      One component absent from the early Earth's surface was free oxygen; (oxygen) it would not have been a constitutent released from the cooling crust. Early production of oxygen was by the photodissociation of water in the Earth's atmosphere, a process that was triggered by the absorption of the Sun's ultraviolet radiation. The reaction is

in which represents the photon of ultraviolet light. The hydrogen produced would escape into space, while the oxygen would react with the early reduced gases by reactions such as 2H2S + 3O2 → 2SO2 + 2H2O. Oxygen production by photodissociation gave the early reduced atmosphere a start toward present-day conditions, but it was not until the appearance of photosynthetic organisms approximately three billion years ago that oxygen could accumulate in the Earth's atmosphere at a rate sufficient to give rise to today's oxygenated environment. The photosynthetic reaction leading to oxygen production is given in equation (6—>).

The transitional hydrosphere
      The nature of the rock record from the time of the first sedimentary rocks (approximately 3.8 billion years ago) to about one to two billion years ago suggests that the amount of oxygen in the Earth's atmosphere was significantly lower than it is today and that there were continuous chemical trends in the sedimentary rocks formed and, more subtly, in the composition of the hydrosphere. The chemistry of rocks shifted dramatically during this transitional period. The source rocks of sediments during this time may have been more basaltic than subsequent ones. Sedimentary debris was formed by the alteration of such source rocks in an oxygen-deficient atmosphere and accumulated primarily under anaerobic marine conditions. The chief difference between reactions involving mineral–ocean equilibria at this time and the present day was the role played by ferrous iron (i.e., reduced state of iron). The concentration of dissolved iron in modern oceans is low because of the insolubility of oxidized iron oxides. During the transition stage and earlier, oxygen-deficient environments were prevalent, and these favoured the formation of minerals containing ferrous iron from the alteration of rocks slightly more rich in basalt than those of today. Indeed, iron carbonate siderite and iron silicate greenalite, in close association with chert (chert and flint) and iron sulfide pyrite, are characteristic minerals that occur in iron formations of the middle Precambrian (about 2.4 to 1.5 billion years ago). The chert originally was deposited as amorphous silica; equilibrium between amorphous silica, siderite, and greenalite at 25° C and a total pressure of one atmosphere requires a carbon dioxide pressure of about 10−2.5 atmosphere, or 10 times the present-day value.

      The oceans of this transitional period can be thought of as a solution that resulted from an acid leach of basaltic rocks, and, because the neutralization of the volatile acid gases was not restricted primarily to land areas as it is today, much of this alteration may have occurred by submarine processes. Anaerobic depositional environments with internal carbon dioxide pressures of about 10−2.5 atmosphere prevailed, and the oxygen-deficient atmosphere itself may have had a carbon dioxide pressure close to 10−2.5 atmosphere. If so, the pH of early ocean water was lower than that of modern seawater and the calcium concentration was higher; moreover, the early ocean water was probably saturated with respect to amorphous silica—roughly 120 ppm.

      To simulate what might have occurred, it is helpful to imagine emptying the Pacific basin, throwing in great masses of broken basaltic material, filling it with hydrogen chloride dissolved in water so that the acid becomes neutralized, and then carbonating the solution by bubbling carbon dioxide through it. Oxygen would not be permitted into the system. The hydrochloric acid would leach the rocks, resulting in the release and precipitation of silica and the production of a chloride ocean containing sodium, potassium, calcium, magnesium, aluminum, iron, and reduced sulfur species in the proportions present in the rocks. As complete neutralization was approached, the aluminum could begin to precipitate as hydroxides and then combine with precipitated silica to form cation-deficient aluminosilicates. As the neutralization process reached its end, the aluminosilicates would combine with more silica and with cations to form such minerals as chlorite, and ferrous iron would combine with silica and sulfur to produce greenalite and pyrite. In the final solution, chlorine would be balanced by sodium and calcium in roughly equal proportions, with subordinate amounts of potassium and magnesium; aluminum would be quantitatively removed, and silicon would be at saturation with amorphous silica. If this solution were then carbonated, calcium would be removed as calcium carbonate, and the chlorine balance would be maintained by abstraction of more sodium from the primary rock. The sediments formed in this system would contain chiefly silica, ferrous iron silicates, chloritic minerals, calcium carbonate, calcium-magnesium carbonates, and small amounts of pyrite.

      If the hydrogen chloride added were in excess of the carbon dioxide, the resultant oceans would have a high content of calcium chloride (CaCl2), but with a pH still near neutrality. If the carbon dioxide added were in excess of the chlorine, calcium would be precipitated as carbonate until it reached a level roughly that of present-day ocean waters—namely, a few hundred parts per million.

      If this newly created ocean were left undisturbed for several hundred million years, its waters would evaporate and be transported onto the continents (in the form of precipitation); streams would transport their loads into it. The sediments produced in this ocean would be uplifted and incorporated into the continents. The influence of the continental debris would gradually be felt and the pH might change somewhat. Iron would be oxidized out of the ferrous silicates to yield iron oxides, but the composition of the water would not vary substantially.

      The primary minerals of igneous rocks are all mildly basic compounds. When these minerals react in excess with acids such as hydrogen chloride and carbon dioxide, they produce neutral or mildly alkaline solutions as well as a set of altered aluminosilicate and carbonate reaction products. It is improbable that seawater has changed through time from a solution approximately in equilibrium with these reaction products—i.e., with clay minerals and carbonates.

The modern hydrosphere
      It is likely that the hydrosphere achieved its modern chemical characteristics about 1.5 to two billion years ago. The chemical and mineralogical compositions and the relative proportions of sedimentary rocks of this age differ little from their counterparts of the Paleozoic era (from 540 to 245 million years ago). Calcium sulfate deposits of late Precambrian age (about 1.5 billion to 540 million years ago) attest to the fact that the acid sulfur gases had been neutralized to sulfate by this time. Chemically precipitated ferric oxides in late Precambrian sedimentary rocks indicate available free oxygen, whatever its percentage. The chemistry and mineralogy of middle and late Precambrian shales are similar to those of Paleozoic shales. The carbon isotopic signature of carbonate rocks has been remarkably constant for more than three billion years, indicating exceptional stability in size and fluxes related to organic carbon. The sulfur isotopic signature of sulfur phases in rocks strongly suggests that the sulfur cycle involving heterotrophic bacterial reduction of sulfate was in operation 2.7 billion years ago. It therefore appears that continuous cycling of sediments similar to those of today has occurred for 1.5 to two billion years and that these sediments have controlled hydrospheric, and particularly oceanic, composition.

      It was once thought that the saltiness of the modern oceans simply represents the storage of salts derived from rock weathering and transported to the oceans by fluvial processes. With increasing knowledge of the age of the Earth, however, it was soon realized that, at the present rate of delivery of salts to the ocean or even at much reduced rates, the total salt content and the mass of individual salts in the oceans could be attained in geologically short time intervals compared to the planet's age. The total mass of salt in the oceans can be accounted for at today's rates of stream delivery in about 12 million years. The mass of dissolved silica in ocean water can be doubled in just 20,000 years by the addition of stream-derived silica; to double the sodium content would take 70 million years. It then became apparent that the oceans were not merely an accumulator of salts; rather, as water evaporated from the oceans, together with some salt, the salts introduced must be removed in the form of minerals deposited in sediments. Accordingly, the concept of the oceans as a chemical system changed from that of a simple accumulator to that of a steady-state system in which rates of inflow of materials equal rates of outflow. The steady-state concept permits influx to vary with time, but the inflow would be matched by nearly simultaneous and equal variation of efflux.

      In recent years, this steady-state conceptual view of the oceans has undergone some modification. In particular, it has been found necessary to treat components of ocean water in terms of all their influxes and effluxes and to be more cognizant of the time scale of application of the steady-state concept. Indeed, the recent increase in the carbon dioxide concentration of the atmosphere due to the burning of fossil fuels may induce a change in the pH and dissolved inorganic carbon concentrations of surface ocean water on a time scale measured in hundreds of years. If fossil-fuel burning were to cease, return to the original state of seawater composition could take thousands of years. Ocean water is not in steady state with respect to carbon on these time scales, but on a longer geologic time scale it certainly could be. Even on this longer time scale, however, oceanic composition has varied because of natural changes in the carbon dioxide level of the atmosphere and because of other factors.

      It appears that the best description of modern seawater composition is that of a chemical system in a dynamic quasi-steady state. Changes in composition may occur over time, but the system always seems to return to a time-averaged, steady-state composition. In other words, since 1.5 to two billion years ago, evolutionary chemical changes in the hydrosphere have been small when viewed against the magnitude of previous change.

      It should be noted that rivers supply dissolved constituents to the oceans, whereas high- and low-temperature reactions between seawater and submarine basalts and reactions in sediment pore waters may add or remove constituents from ocean water. Biological processes involved in the formation of the opaline silica skeletons of diatoms and radiolarians and the carbonate skeletons of planktonic foraminiferans and coccolithophorids chiefly remove calcium and silica from seawater. Exchange reactions between river-borne clays entering seawater are particularly significant for sodium and calcium ions. Most of the carbon imbalance in ocean water represents carbon released to the ocean–atmosphere system during precipitation of carbonate minerals—i.e.,

      In the case of iron, it has been documented that “dissolved” iron carried by rivers is rapidly precipitated as hydroxides in the mixing zone with seawater and that the reduced dissolved iron released from anaerobic sediments also is rapidly precipitated under the oxic conditions (i.e., those with oxygen present) prevailing in the water column. Iron is also precipitated as iron smectites, hydrated iron oxides, and nontronite (iron-rich montmorillonite) in the deep sea. It is thus likely that iron is removed by these processes.

sodium imbalance
      The imbalance in sodium is large; 45 percent of the river input is not accounted for in the mass balance calculations. There are, however, major uncertainties in the estimation of the pore-water flux of sodium ions. An important sink for sodium on a geologic time scale is the formation of evaporites. If the amount of unbalanced sodium is expressed in terms of halite deposition, it would correspond to 1.6 × 1014 grams of sodium chloride per year as compared with a potential total depositional rate of 3.3 × 1014 grams annually. There are no important sodium chloride deposits forming today; thus, one possibility is that sodium is accumulating in the oceans. If so, in 6 × 106 years at an accumulation rate of 63 × 1012 grams of sodium annually, the average salinity of the oceans would increase less than one part per thousand. The chlorine balance for the oceans, however, indicates that it is likely that the major problem in the imbalance for sodium lies in the flux estimates for sediment pore waters and perhaps submarine weathering processes.

      Modern seawater chemistry has been characteristic of roughly the past 600 million years of ocean history. Evaporite sediments provide strong evidence that the composition of seawater has not varied a great deal during this interval of geologic time. Nonetheless, it seems likely that fluctuations did occur, particularly in the concentrations of calcium, magnesium, and sulfate ions. The isotopic composition of sulfur in seawater, as recorded in evaporites, has varied dramatically during the past one billion years. Although it is difficult to relate these isotopic fluctuations to the calcium and sulfate concentrations of seawater, some scientists believe that the fluctuations do in fact imply changes in the latter. Furthermore, the major features of the sulfur isotopic curve for evaporites versus Phanerozoic time is similar to that of the strontium-87/strontium-86 ratio, and perhaps the strontium/calcium ratio, of sedimentary materials during this time interval. Such covariation is consistent with a model in which fluxes related to alteration of seafloor basalts and continental river runoff vary with time, resulting in variation in seawater composition.

Changes in the chemistry of the atmosphere-hydrosphere (atmosphere)
      The chemistry of the atmosphere has certainly changed significantly during the past one billion years of Earth history. A modification of this kind implies changes in the chemistry of the hydrosphere as well. Oxygen in the atmosphere rose substantially between two billion years ago and the beginning of the Phanerozoic eon (i.e., 540 million years ago), whereas atmospheric carbon dioxide levels probably decreased. This change led in general to a progressively more oxygenated and less acidic hydrosphere. It is likely that the development of higher land plants during the Devonian period (from 408 to 360 million years ago) resulted in an increase in atmospheric oxygen and a decrease in carbon dioxide. Air trapped within bubbles in Arctic and Antarctic ice shows that the carbon dioxide content of the atmosphere during the climax of the last ice age was about 180 parts per million by volume (ppmv), and atmospheric CO2 levels reached approximately 280 ppmv during the last great interglacial of 120,000 years ago, long before modern society initiated its extensive fossil-fuel burning and deforestation activities (see below Buildup of greenhouse gases (hydrosphere)). These atmospheric changes in themselves can influence the chemistry of the hydrosphere, but they also appear to be coupled with other changes in the rock–ocean–atmosphere–biota system that strongly affect hydrospheric chemistry. For example, though surface waters probably remained oxygenated during the Cretaceous and Devonian periods of Earth history, there is evidence that intermediate and deep ocean waters were more anoxic (oxygen depleted) than today. These “anoxic events” are characterized by dramatic changes in the Earth's ocean–atmosphere system, including changes in the rates of the cyclic transfer of elements at the Earth's surface and in atmospheric composition. The probable impact of a bolide (either an asteroid or comet) and increased volcanism at the end of the Cretaceous (about 66.4 million years ago), though still the subject of hot debate, certainly could have caused short-term changes in the chemistry of the Earth's atmosphere-hydrosphere. Scientists speculate that such an event could have had any of several results, including (1) changes in the Earth's radiant-energy balance because of vast amounts of particulate and gaseous input into the atmosphere and subsequent cooling, (2) acid rain stemming from the input into the atmosphere of nitrogen gases generated by the bolide impact, (3) increased trace metal fluxes to the hydrosphere brought on by the destruction of the bolide and increased volcanism, and perhaps (4) increased “anoxia” of the hydrosphere owing to the death of land and marine organisms.

      Whatever the case, evidence of change is present in the rock record albeit the composition of the modern hydrosphere has not varied greatly for the past one billion years. Moreover, as will be seen in the following section, humankind is modifying not only the chemistry of local and regional water bodies but also that of the entire global atmosphere-hydrosphere by increasing the rates of input of natural substances and by introducing new synthetic substances to the environment.

Impact of human activities on the hydrosphere
      The activities of modern society are having a severe impact on the hydrologic cycle. The dynamic steady state is being disturbed by the discharge of toxic chemicals, radioactive substances, and other industrial wastes and by the seepage of mineral fertilizers, herbicides, and pesticides into surface and subsurface aquatic systems. Inadvertent and deliberate discharge of petroleum, improper sewage disposal, and thermal pollution also are seriously affecting the quality of the hydrosphere.

      The present discussion focuses on three major problems—eutrophication, acid rain, and the buildup of the so-called greenhouse gases. Each exemplifies human interference in the hydrologic cycle and its far-reaching effects.

      Historically, aquatic systems have been classified as oligotrophic or eutrophic. Oligotrophic waters are poorly fed by the nutrients nitrogen and phosphorus and have low concentrations of these constituents. There is thus low production of organic matter by photosynthesis (equation [6—>]) in such waters. By contrast, eutrophic waters are well supplied with nutrients and generally have high concentrations of nitrogen and phosphorus and, correspondingly, large concentrations of plankton owing to high biological productivity. The waters of such aquatic systems are usually murky, and lakes and coastal marine systems may be oxygen-depleted at depth. The process of eutrophication is defined as high biological productivity resulting from increased input of nutrients or organic matter into aquatic systems. For lakes, this increased biological productivity usually leads to decreased lake volume because of the accumulation of organic detritus. Natural eutrophication occurs as aquatic systems fill in with organic matter; it is distinct from cultural eutrophication, which is caused by human intervention. The latter is characteristic of aquatic systems that have been artificially enriched by excess nutrients and organic matter from sewage, agriculture, and industry. Naturally eutrophic lakes may produce 75–250 grams of carbon per square metre per year, whereas those lakes experiencing eutrophication because of human activities can support 75–750 grams per square metre per year. Commonly, culturally eutrophic aquatic systems may exhibit extremely low oxygen concentrations in bottom waters. This is particularly true of stratified systems, as, for instance, lakes during summer where concentrations of molecular oxygen may reach levels of less than about one milligram per litre—a threshold for various biological and chemical processes.

      Aquatic systems may change from oligotrophic to eutrophic, or the rate of eutrophication of a natural eutrophic system may be accelerated by the addition of nutrients and organic matter due to human activities. The process of cultural eutrophication, however, can be reversed if the excess nutrient and organic matter supply is shut off.

      Not only do freshwater aquatic systems undergo cultural eutrophication, but coastal marine systems also may be affected by this process. On a global scale, the input by rivers of organic matter to the oceans today is twice the input in prehuman times, and the flux of nitrogen, together with that of phosphorus, has more than doubled. This excess loading of carbon, nitrogen, and phosphorus is leading to cultural eutrophication of marine systems. In several polluted eastern U.S. estuaries and in some estuaries of western Europe (e.g., the Scheldt of Belgium and The Netherlands), all of the dissolved silica brought into the estuarine waters by rivers is removed by phytoplankton growth (primarily diatoms) resulting from excess fluxes of nutrients and organic matter. In the North Sea, there is now a deficiency of silica and an excess of nitrogen and phosphorus, which in turn has led to a decrease in diatom productivity and an increase in cyanobacteria productivity—a biotic change brought about by cultural eutrophication.

      The emission of sulfur dioxide and nitrogen oxides to the atmosphere by human activities—primarily fossil-fuel burning—has led to the acidification of rain and freshwater aquatic systems. Acid rain is a worldwide problem and has been well documented for the eastern United States and the countries of western Europe.

      Acid rain is defined as precipitation with a pH of less than 5.7 that results from reactions involving gases other than carbon dioxide. The overall reactions that produce such precipitation are those of equations (1—>), (2—>), and (3—>) and

      A figure showing the average pH = −log aH+ (aH+ is activity of the hydrogen ion) of precipitation over the eastern United States for the period October 1979 through September 1980 revealed that low pH values are a result of equilibration of rainwater with the atmospheric acid gases of carbon, nitrogen, and sulfur. Equilibration only with atmospheric carbon dioxide would give a pH of 5.7. The significantly lower values are a result of reactions with nitrogen- and sulfur-bearing gaseous atmospheric components derived primarily from fossil-fuel burning sources. Nitrate and sulfate concentrations in precipitation over the eastern United States are strongly correlated with pH—the lower the pH of rain, the higher the concentrations of nitrate and sulfate. Such low pH values and increased nitrate and sulfate concentrations also are found in the rains of western Europe and other industrialized regions of the world.

 Table 8—> provides an illustrative example of events occurring globally—namely, the processes—that remove anthropogenic emissions of sulfur dioxide and nitrogen oxides from the atmosphere of the eastern United States. Wet and dry deposition also removes the hydrogen ion produced in the rain by the oxidation and hydrolysis of these acid gases. This excess hydrogen ion can bring about the acidification of freshwater aquatic systems, particularly those with little buffer capacity (e.g., lakes situated in crystalline rock terrains). Furthermore, the lower pH values of rainwater, and consequently of soil water, can lead to increased mobilization of aluminum. Acidification of freshwater lakes in the eastern United States and increased aluminum concentrations in their waters are thought to be responsible for major changes in the ecosystems of the lakes. In particular, many lakes of this region lack substantial fish populations today, even though they supported large numbers of fish in the early 1900s. Acid rain also may be among the factors responsible for damage to the major forests of the eastern United States and western Europe.

Buildup of greenhouse (greenhouse effect) gases
      One problem brought about by human action that is definitely affecting the hydrosphere globally is that of the greenhouse gases (so called because of their heat-trapping “greenhouse” properties) emitted to the atmosphere. Of the greenhouse gases released by anthropogenic activities, carbon dioxide has received much attention. It has been shown from the measurements of carbon dioxide in air bubbles trapped in ice and from the continuous measurement of carbon dioxide concentrations in air samples collected at Mauna Loa, Hawaii, since 1958 that the present atmospheric concentration of nearly 350 ppmv is 25 percent higher than its late-1700s value. Much of this increase is due to carbon dioxide released to the atmosphere from the burning of coal, oil, gas, and wood and from the slash-and-burn activities that accompany deforestation practices (as, for example, those adopted in the Amazon River basin). The component of the hydrosphere most greatly affected by this emission of carbon (carbon cycle) dioxide is the ocean.

      Figure 5A shows the manner in which carbon is cycled in the global environment on a long-term geologic basis. Before human activities had substantially affected the carbon dioxide cycle, there was a net flux of carbon dioxide from the oceans through the atmosphere to the land, where the gas was used in the net production of organic matter and the chemical weathering of minerals in continental rocks. Because of fossil-fuel burning and land-use practices, the net transfer from the ocean to the land has been reversed, and the ocean has now become an important sink of carbon dioxide (Figure 5B). The oceans are currently gaining 2,340 million tons of carbon per year. The net chemical reaction of adding carbon dioxide to the ocean (provided there is no reaction with carbonate solids) is

      and a lowering of the pH of surface seawater. Such a pH effect has not been observed but conceivably could occur if carbon dioxide continues to be released to the atmosphere by human activities.

      Based on greenhouse climate models and other considerations, it is possible that atmospheric carbon dioxide concentrations may double their late-1700s level by the years 2030–2050 and, along with those of other greenhouse gases (e.g., methane and nitrous oxide), give rise to a global mean surface temperature increase of 1.5° to 5° C. This projected temperature increase would be two to three times greater at the poles than at the equator and greater in the Arctic than in the Antarctic. At present there is no worldwide program to decrease greenhouse gas emissions, except for that affecting chlorofluorocarbon (Freon) releases; thus, it is conceivable that atmospheric carbon dioxide concentrations in the late 21st and early 22nd centuries might reach levels greater than twice their 1700s value. Whatever the case, the effect of the potential rise in surface temperature would be to speed up the hydrologic cycle and probably the rate of chemical weathering of continental rocks. Increases of 4 to 7 percent in the global mean evaporation and precipitation rates might occur for a doubling of the carbon dioxide level and a few degrees rise in global mean temperature. The effect on the water balance would be regional in nature, with some places becoming wetter and others drier. In general, there would be a trend toward greater and longer periods of summer dryness induced by lower soil moisture content and higher evaporation rates in the mid-latitudes of the Northern Hemisphere. In the arid western regions of the United States, which depend on irrigation for growing plants, severe water shortages could occur. By contrast, precipitation and runoff might increase, except in summer, at latitudes beyond 60° N because of a greater poleward transport of moisture. In summer, in a zone centred around 60° N, greater dryness might occur due to an earlier end of snowmelt.

      Global warming could further affect the hydrologic cycle by the melting of ice and snow in the Greenland and Antarctic ice caps and in mountain glaciers, resulting in the transfer of water to the oceans. This process, together with thermal expansion of the oceans because of global warming, could lead to a slow rise in sea level of about 0.7 metre over the next century. If the West Antarctic ice sheet were to disintegrate, a much larger and more rapid rise in sea level of 5–6 metres could occur over the next several hundred years. The melting of all glacial ice would raise the sea level about 56 metres. It is also possible that a global warming could result in a reduction in the areal extent and thickness of sea ice in the Arctic and circum-Antarctic regions. Complete melting of the Arctic sea ice might occur, causing a northward shift in storm tracks and a reduction in Northern Hemispheric precipitation during the spring and fall. Furthermore, a worldwide reduction in sea ice might lead to increased evaporation from the ocean and increased low-altitude cloudiness, which would reflect solar radiation and cause cooling.

      The potential changes in the hydrologic cycle induced by a global warming resulting from anthropogenic emissions of greenhouse gases do not seem great. Yet, their consequences could be severe for ecosystems and human populations, especially since the latter are so sensitive to and dependent on such changes. A global rise in sea level of one metre, for example, would almost completely inundate the coastal areas of Bangladesh. Island nations and continental beaches and cities would be endangered. Agricultural lands could be displaced, just as patterns of arid, semiarid, and wet lands might become modified. It is essential that society plan for such potential changes so that, if they do occur, appropriate adjustments can be made to accommodate them.

Additional Reading
General introductory discusssions on the distribution of water on and around the Earth and the role of water in supporting life are found in Cynthia A. Hunt and Robert M. Garrels, Water: The Web of Life (1972); C.L. Mantell and A.M. Mantell, Our Fragile Water Planet: An Introduction to the Earth Sciences (1976); Elizabeth Kay Berner and Robert A. Berner, The Global Water Cycle: Geochemistry and Environment (1987); H.M. Raghunath, Ground Water, 2nd ed. (1987); Eberhard Czaya, Rivers of the World (1981; originally published in German, 1981); Mary J. Burgis and Pat Morris, The Natural History of Lakes (1987); and Neil Wells, The Atmosphere and Ocean: A Physical Introduction (1986).Biogeochemical properties of the hydrosphere are discussed in Ronald J. Gibbs, “Mechanisms Controlling World Water Chemistry,” Science 170(3962):1088–1090 (1970); R. Wollast and Fred T. Mackenzie, “Global Cycle of Silica,” in S.R. Aston (ed.), Silicon Geochemistry and Biogeochemistry (1983), pp.39–76; P. Buat-Menard, “Particle Geochemistry in the Atmosphere and Oceans,” in Peter S. Liss and W. George N. Slinn (eds.), Air-Sea Exchange of Gases and Particles (1983), pp. 455–532; Robert A. Berner, A.C. Lasaga, and Robert M. Garrels, “The Carbonate-Silicate Geochemical Cycle and Its Effect on Atmospheric Carbon Dioxide Over the Past 100 Million Years,” American Journal of Science 283(7):641–683 (1983); Lawrence A. Hardie and Hans P. Eugster, “The Evolution of Closed-Basin Brines,” Mineralogical Society of America Special Paper 3:273–290 (1970); James I. Drever, The Geochemistry of Natural Waters (1982); A. Lerman, Geochemical Processes: Water and Sediment Environments (1979, reprinted 1988); G. Evelyn Hutchinson, A Treatise on Limnology, vol. 1 (1975); Werner Stumm (ed.), Chemical Processes in Lakes (1985); and Georg Matthess, The Properties of Groundwater (1982; originally published in German, 1973).Analyses of the processes of the hydrologic cycle and its utilization are presented in Robert C. Averett and Diane M. Mcknight (eds.), Chemical Quality of Water and the Hydrologic Cycle (1987); R.A. Freeze, “A Stochastic-Conceptual Analysis of Rainfall-Runoff Processes on a Hillslope,” Water Resources Research 16(2):391–408 (1980); T. Dunne, “Field Studies of Hillslope Flow Processes,” in M.J. Kirkby (ed.), Hillslope Hydrology (1978), pp. 227–293; Ray K. Linsley and Joseph B. Franzini, Water-Resources Engineering, 3rd ed. (1979); Mark J. Hammer and Kenneth A. Mackichan, Hydrology and Quality of Water Resources (1981); and Alvin S. Goodman, Principles of Water Resources Planning (1984).For the evolution of the hydrosphere, see James C.G. Walker, Evolution of the Atmosphere (1977); J. Veizer,“The Evolving Exogenic Cycle,” in C. Bryan Gregor et al. (eds.), Chemical Cycles in the Evolution of the Earth (1988), pp. 175–220; Heinrich D. Holland, The Chemical Evolution of the Atmosphere and Oceans (1984); and Robert M. Garrels and Fred T. Mackenzie, Evolution of Sedimentary Rocks (1971).The impact of human activities on the hydrosphere is studied in Robert M. Garrels, Fred T. Mackenzie, and Cynthia A. Hunt, Chemical Cycles and the Global Environment: Assessing Human Influences (1975); Arthur N. Strahler and Alan H. Strahler, Environmental Geoscience: Interaction Between Natural Systems and Man (1973); W.D. Bischoff, V.l. Paterson, and Fred T. Mackenzie, “Geochemical Mass Balance for Sulfur- and Nitrogen-Bearing Acid Components: Eastern United States,” in Owen P. Bricker (ed.), Geological Aspects of Acid Deposition (1984), pp. 1–21; S.C. Chapra, “Simulation of Recent and Projected Total Phosphorus Trends in Lake Ontario,” Journal of Great Lakes Research 6(2):101–112 (1980); Brian Henderson-Sellers and H.R. Markland, Decaying Lakes: The Origins and Control of Cultural Eutrophication (1987); Thomas D. Brock, A Eutrophic Lake: Lake Mendota, Wisconsin (1985); G. Dennis Cooke et al., Lake and Reservoir Restoration (1986); Sven Olof Ryding and Walter Rast (eds.), The Control of Eutrophication of Lakes and Reservoirs (1989); Louis Thibodeaux, Chemodynamics, Environmental Movement of Chemicals in Air, Water, and Soil (1979); Stanley E. Manahan, Environmental Chemistry, 4th ed. (1984); Howard S. Peavy, Donald R. Rowe, and George Tchobanoglous, Environmental Engineering (1985); James L. Regens and Robert W. Rycroft, The Acid Rain Controversy (1988); and Daniel D. Chiras, Environmental Science: A Framework for Decision Making, 2nd ed. (1988).Fred T. Mackenzie

* * *


Universalium. 2010.

Игры ⚽ Поможем сделать НИР

Look at other dictionaries:

  • hydrosphère — [ idrɔsfɛr ] n. f. • 1897; de hydro et sphère ♦ Géogr. L élément liquide de la Terre (eau liquide, glaces et neiges, vapeur d eau). ● hydrosphère nom féminin Totalité des eaux de la planète, comprenant aussi bien les océans, les mers, les lacs,… …   Encyclopédie Universelle

  • Hydrosphere — Hydrosphère L hydrosphère est la partie de la planète occupée par l eau liquide (océans, mers, lacs, fleuves, nappes phréatiques) et solide (calottes polaires, glaciers, banquise). Cette sphère s étend depuis environ 8 km d altitude (la cime des… …   Wikipédia en Français

  • hydrosphere — [hī΄drōsfir΄, hī΄drəsfir΄] n. [ HYDRO + SPHERE] all the water on the surface of the earth, including oceans, lakes, glaciers, etc.: water vapor, clouds, etc. may be considered part of the atmosphere or the hydrosphere …   English World dictionary

  • Hydrosphere — Hy dro*sphere, n. [Pref. hydro , 1 + sphere.] 1. (Meteor.) The aqueous vapor of the entire atmosphere. [Webster 1913 Suppl.] 2. (Phys. Geog.) The aqueous envelope of the earth, including the ocean, all lakes, streams, and underground waters, and… …   The Collaborative International Dictionary of English

  • hydrosphere — 1887, from HYDRO (Cf. hydro ) + SPHERE (Cf. sphere) …   Etymology dictionary

  • hydrosphere — ► NOUN ▪ the seas, lakes, and other waters of the earth s surface, considered collectively …   English terms dictionary

  • Hydrosphère — L hydrosphère est la partie de la planète occupée par l eau liquide (océans, mers, lacs, fleuves, nappes phréatiques, nuages) , solide (calottes polaires, glaciers, banquise) et gazeuse (vapeur d eau). Cette sphère s étend depuis environ 8 km d… …   Wikipédia en Français

  • Hydrosphere — A hydrosphere (from Greek ύδωρ hydor , water + σφαίρα sphaira , sphere ) in physical geography describes the collective mass of water found on, under, and over the surface of a planet.Other hydrospheresA thick hydrosphere is thought to exist… …   Wikipedia

  • hydrosphere — hidrosfera statusas T sritis chemija apibrėžtis Žemės rutulio vandens telkinių visuma. atitikmenys: angl. hydrosphere rus. гидросфера …   Chemijos terminų aiškinamasis žodynas

  • hydrosphere — hidrosfera statusas T sritis ekologija ir aplinkotyra apibrėžtis Žemę gaubiantis nevientisas vandens sluoksnis, Žemės vandenų visuma (vandenynai, jūros, atmosferos, sausumos vandenys). Vanduo dengia 70,8 % Žemės rutulio paviršiaus. atitikmenys:… …   Ekologijos terminų aiškinamasis žodynas

Share the article and excerpts

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