hydrologic sciences

hydrologic sciences

Introduction

      the fields of study concerned with the waters of the Earth. Included are the sciences of hydrology, oceanography, limnology, and glaciology.

      In its widest sense hydrology encompasses the study of the occurrence, the movement, and the physical and chemical characteristics of water in all its forms within the Earth's hydrosphere. In practice hydrologists usually restrict their studies to waters (water) close to the land surface of the Earth. Water in the atmosphere is usually studied as part of meteorology. Water in the oceans and seas is studied within the science of oceanography; water in lakes and inland seas within limnology; and ice on the land surface within glaciology. Clearly there is some overlap between these major scientific disciplines; both hydrologists and meteorologists, for example, have contributed to the study of water movement in the lower boundary layers of the atmosphere. All are linked by the fundamental concept of the hydrologic cycle, according to which the waters of the sea are evaporated (vaporization), are subsequently condensed within the atmosphere, fall to the Earth as precipitation, and finally flow in the rivers back to the sea.

      Water is the most abundant substance on Earth and is the principal constituent of all living things. Water in the atmosphere plays a major role in maintaining a habitable environment for human life. The occurrence of surface waters has played a significant role in the rise and decline of the major civilizations in world history. In many societies the importance of water to humankind is reflected in the legal and political structures. At the present time rising populations and improving living standards are placing increasing pressures on available water resources. There is, in general, no shortage of water on the Earth's land surface, but the areas of surplus water are often located far from major centres of population. Moreover, in many cases these centres prove to be sources of water pollution. Thus, the availability and quality of water are becoming an ever-increasing constraint on human activities, notwithstanding the great technological advances that have been made in the control of surface waters.

Study of the waters close to the land surface
       hydrology deals with that part of the hydrologic cycle from the arrival of water at the land surface as precipitation to its eventual loss from the land either by evaporation or transpiration back to the atmosphere or by surface and subsurface flow to the sea. It is thus primarily concerned with waters close to the land surface. It includes various component disciplines of a more specialized nature. Hydraulics is concerned with the mechanics and dynamics of water in its liquid state. Hydrography is the description and mapping of the bodies of water of the Earth's surface (including the oceans), with a particular concern for navigation charts. Hydrometry involves measurements of surface water, particularly precipitation and streamflow. Hydrometeorology focuses on water in the lower boundary layer of the atmosphere. Groundwater hydrology and hydrogeology have to do with subsurface water in the saturated (saturation) zone, while soil water physics involves the study of subsurface water in the unsaturated zone. Engineering hydrology is concerned with the design of man-made structures that control the flow and use of water.

      Underlying all the hydrologic sciences is the concept of water balance, an expression of the hydrologic cycle for an area of the land surface in terms of conservation of mass. In a simple form the water balance may be expressed as

      S = PQEG,

      where S is the change of water storage in the area over a given time period, P is the precipitation input during that time period, Q is the stream discharge from the area, E is the total of evaporation and transpiration to the atmosphere from the area, and G is the subsurface outflow. Most hydrologic studies are concerned with evaluating one or more terms of the water balance equation. Because of the difficulties in quantifying the movement of water across the boundaries of an area under study, the water balance equation is most easily applied to an area draining to a particular measurement point on a stream channel. This area is called a catchment (or sometimes a watershed in the United States). The line separating adjacent catchments is known as a topographical divide, or simply a divide. The following sections describe the study of the different elements of the catchment water balance and the way in which they affect the response of catchments over time under different climatic regimes.

Evaluation of the catchment water balance
      Precipitation results from the condensation of water from the atmosphere as air is cooled to the dew point, the temperature at which the air becomes saturated with respect to water vapour. The cooling process is usually initiated by uplift of the air, which may result from a number of causes, including convection, orographic effects over mountain ranges, or frontal effects at the boundaries of air masses of different characteristics. Condensation within the atmosphere requires the presence of condensation nuclei to initiate droplet formation. Some of the condensate may be carried considerable distances as cloud before being released as rain or snow, depending on the local temperatures. Some precipitation in the form of dew or fog results from condensation at or near the land surface. In some areas, such as the coastal northwest of the United States, dew and fog drip can contribute significantly to the water balance. The formation of hail requires a sequence of condensation and freezing episodes, resulting from successive periods of uplift. Hailstones usually show a pattern of concentric rings of ice as a result.

      Direct measurements of precipitation are made by a variety of gauges, all of which consist of some form of funnel that directs the infalling water to some storage container. Storage gauges simply store the incident precipitation, and the accumulated water is usually measured on a daily, weekly, or monthly basis. Recording gauges allow rates of precipitation to be determined.

      Rainfall (rain) volumes are usually converted to units of depth—volume per unit area. Measurements obtained from different types of rain gauges are not directly comparable because of varying exposure, wind, and splash effects. The most accurate type of gauge is the ground-level gauge, in which the orifice of the gauge is placed level with the ground surface and surrounded by an antisplash grid. Rain gauge catches decrease as the orifice is raised above the ground, particularly in areas subject to high winds. In areas of significant snowfall, however, it may be necessary to raise the rain gauge so that its orifice is clear of the snow surface. Various shields for the gauge orifice have been tried in an effort to offset wind effects. Wind effects are greater for snow than for rain and for small drops or light rainfall than for large drops.

      An impression of the spatial distribution of precipitation intensity can be achieved through indirect measurements of precipitation, in particular radar scattering. The relationship between rainfall intensity and measured radar signals depends on various factors, including the type of precipitation and the distribution of drop size. Radar measurements are often used in conjunction with rain gauges to allow on-line calibration in converting the radar signal to precipitation amounts. The radar measurements are, however, at a much larger spatial scale. Resolution of five to 10 square kilometres is common for operational systems. Even so, this provides a much better picture of the spatial patterns of precipitation over large catchment areas than has been previously possible. The use of satellite remote sensing to determine rainfall volumes is still in its early stages, but the technique appears likely to prove useful for estimating amounts of precipitation in remote areas.

      The measurement of inputs of snow to the catchment water balance is also a difficult problem. The most basic technique involves the snow course, a series of stakes to measure snow depths. Snowfalls can, however, vary greatly in density, depending primarily on the temperature history of snow formation. Accumulated snow changes its density over time prior to melting. Snow density can be measured by weighing a sample of known volume taken in a standard metal cylinder. Other techniques for measuring snowfall include the use of snow pillows, which record the changing weight of snow lying above them, or the use of rain gauges fitted with heating elements, which melt the snow as it falls. These techniques are subject to wind effects, both during a storm event and between events because of redistribution of snow by the wind.

      Summary statistics on precipitation are usually produced on the basis of daily, monthly, and annual amounts falling at a given location or over a catchment area. The frequency at which a rainfall of a certain volume occurs within a certain period is also important to hydrologic analysis. The assessment of this frequency, or the recurrence interval of the rainfall from the sample of available data, is a statistical problem generally involving the assumption of a particular probability (probability and statistics) distribution to represent the characteristics of rainfalls. Such analyses must assume that this distribution is not changing over time, even though it has been shown that in some areas of the world climatic change may cause rainfall statistics to vary. It has long been speculated that rainfalls may exhibit cyclic patterns over long periods of time, and considerable effort has been expended in searching for such cycles. In some areas the annual seasonal cycle is of paramount importance, but demonstrations of longer periodicities have not proved of general applicability.

      Patterns of rainfall intensity and duration are of great importance to the hydrologist in predicting catchment discharges and water availability and in dealing with floods, droughts, land drainage, and soil erosion. Rainfalls vary both within and between rainstorms, sometimes dramatically, depending on the type and scale of the storm and its velocity of movement. Within a storm, the average intensity tends to decrease with an increase in the storm area.

      On a larger scale, seasonal variations in rainfall vary with climate. Humid temperate areas tend to have rainfalls that are fairly evenly distributed throughout the year; Mediterranean areas have a winter peak with low summer rainfalls; savanna areas have a double peak in rainfall; and equatorial areas again have a relatively even distribution of rainfall over the course of the year. Average annual rainfalls also vary considerably. The minimum recorded long-term average is 0.76 millimetre at Arica, Chile; the maximum 11,897.36 millimetres at Tutunendo, Colom. The maximum recorded rainfall intensities are 38 millimetres in one minute (Barot, Guadeloupe, 1970); 1,870 millimetres in a single day (Cilaos, Réunion, 1952); and 26,461 millimetres in one year (Cherrapunji, India, 1861).

      When precipitation reaches the surface in vegetated areas, a certain percentage of it is retained on or intercepted by the vegetation. Rainfall that is not intercepted is referred to as throughfall. Water that reaches the ground via the trunks and stems of the vegetation is called stemflow. The interception storage capacities of the vegetation vary with the type and structure of the vegetation and with meteorologic factors. Measurements have shown that up to eight millimetres of rainfall can be intercepted by some vegetation canopies. The intercepted water is evaporated back into the atmosphere at rates determined by the prevailing meteorologic conditions and the nature of the vegetation. In humid temperate areas evaporation of intercepted water can be an important component of the water balance. Forest areas have been shown to have greater interception losses than adjacent grassland areas. This is due to the greater aerodynamic roughness of the forest canopy, resulting in a much more efficient transfer of water vapour away from the surface.

      When water from a rainstorm or a period of snowmelt reaches the ground, some or all of it will infiltrate the soil. The rate of infiltration depends on the intensity of the input, the initial moisture condition of the surface soil layer, and the hydraulic characteristics of the soil. Small-scale effects such as the presence of a surface seal of low permeability (due to the rearrangement of surface soil particles by rain splash) or the presence of large channels and cracks in the surface soil may be important in controlling infiltration rates. Water in excess of the infiltration capacity of the soil will flow overland as surface runoff once the minor undulations in the surface (the depression storage) have been filled. Such runoff occurs most frequently on bare soils and in areas subject to high rainfall intensities. In many environments rainfall intensities rarely exceed the infiltration capacities of vegetated soil surfaces. The occurrence of surface runoff is then more likely to be generated by rainfall on completely saturated soil.

      Rates of evapotranspiration of water back to the atmosphere depend on the nature of the surface, the availability of water, and the “evaporative demand” of the atmosphere (i.e., the rate at which water vapour can be transported away from the surface under the prevailing meteorologic conditions). Estimation of evapotranspiration rates is important in determining expected rates of stream discharge and in controlling irrigation (irrigation and drainage) schemes. The concept of potential evapotranspiration—the possible rate of loss without any limits imposed by the supply of water—has been an important one in the development of hydrology. Most direct measurements of rates of potential evapotranspiration are made using standard evapotranspiration pans with an open water surface. Such measurements serve as a useful standard for comparative purposes, but measured rates may be very different from appropriate potential rates for the surrounding surfaces because of the different thermal and roughness characteristics of the vegetation. In fact, the measured pan rate may be affected by the nature of the surrounding surface due to the influence of evapotranspiration on the humidity of the lower atmosphere.

      A distinction also must be drawn between potential rates of evapotranspiration and actual rates. Actual rates may be higher than pan rates for a well-watered, rough vegetation canopy. With a limited water supply available from moisture in the soil, actual rates will fall below potential rates, gradually declining as the moisture supply is depleted. Plants can have some effect on rates of evapotranspiration under dry conditions through physiological controls on their stomata—small openings in the leaf surfaces that are the primary point of transfer of water vapour to the atmosphere. The degree of control varies with plant species.

      The only reliable way of measuring actual evapotranspiration is to use large containers (sometimes on the order of several metres across) called lysimeters, evaluate the different components of the water balance precisely, and calculate the evapotranspiration by subtraction. A similar technique is often employed at the catchment scale, although the measurement of the other components of the water balance is then necessarily less precise.

Soil moisture
      The soil provides a major reservoir for water within a catchment. Soil moisture levels rise when there is sufficient rainfall to exceed losses to evapotranspiration and drainage to streams and groundwater. They are depleted during the summer when evapotranspiration rates are high. Levels of soil moisture are important for plant and crop growth, soil erosion, and slope stability. The moisture status of the soil is expressed in terms of a volumetric moisture content and the capillary potential of the water held in the soil pores. As the soil becomes wet, the water is held in larger pores, and the capillary potential increases.

      Capillary potential may be measured by using a tensiometer consisting of a water-filled porous cup connected to a manometer or pressure transducer. Soil moisture content is often measured gravimetrically by drying a soil sample under controlled conditions, though there are now available moisture meters based on the scattering of neutrons or absorption of gamma rays from a radioactive source.

      The rate at which water flows through soil is dependent on the gradient of hydraulic potential (the sum of capillary potential and elevation) and the physical properties of the soil expressed in terms of a parameter called hydraulic conductivity, which varies with soil moisture in a nonlinear way. Measured sample values of hydraulic conductivity have been shown to vary rapidly in space, making the use of measured point values for predictive purposes at larger scales subject to some uncertainty.

      Water also moves in soil because of differences in temperature and chemical concentrations of solutes in soil water. The latter, which can be expressed as an osmotic potential, is particularly important for the movement of water into plant roots due to high solute concentrations within the root water.

      Some rocks allow little or no water to flow through; these are known as impermeable rocks, or aquicludes. Others are permeable and allow considerable storage of water and act as major sources of water supply; these are known as aquifers (aquifer). Aquifers overlain by an impermeable layer are called confined aquifers; aquifers overlain by an unsaturated, or vadose, zone of permeable materials are called unconfined aquifers. The boundary between the saturated and unsaturated zones is known as the water table. In some confined aquifers, hydraulic potentials may exceed those required to bring the water to the surface. These are artesian (artesian well) aquifers. A well drilled into such an aquifer will cause water to gush to the surface, sometimes with considerable force. Continued use of artesian water, however, will cause potentials to decline until eventually the water may have to be pumped to the surface.

      The water found in groundwater bodies is replenished by drainage through the soil, which is often a slow process. This drainage is referred to as groundwater recharge. Rates of groundwater recharge are greatest when rainfall inputs to the soil exceed evapotranspiration losses. When the water table is deep underground, the water of the aquifer may be exceedingly old, possibly resulting from a past climatic regime. A good example is the water of the Nubian sandstone aquifer, which extends through several countries in an area that is now the Sahara desert. The water is being used extensively for water supply and irrigation purposes. Radioisotope dating techniques have shown that this water is many thousands of years old. The use of such water, which is not being recharged under the current climatic regime, is termed groundwater mining.

      In many aquifers, groundwater levels have fallen drastically in recent times. Such depletion increases pumping costs, causes wells and rivers to dry up, and, where a coastal aquifer is in hydraulic contact with seawater, can cause the intrusion of saline water. Attempts have been made to augment recharge by the use of waste waters and the ponding of excess river flows. A scheme to pump winter river flows into the Chalk aquifer that underlies London has reversed the downward trend of the water table.

      Water-table levels in an aquifer are measured by using observation wells. Successive measurements of water levels over time may be plotted as a well hydrograph. The hydraulic characteristics of a particular aquifer around a well can be determined by the response of the water table to controlled pumping. Many aquifers exhibit two types of water storage: primary porosity consisting of the smaller pores and secondary porosity or fractures within the rock mass. The latter may make up only a small proportion of the total pore space but may dominate the flow characteristics of the aquifer. They are of particular importance to the movement of pollutants through the groundwater.

runoff and stream discharge
      Runoff is the downward movement of surface water under gravity in channels ranging from small rills to large rivers. Channel flows of this sort can be perennial, flowing all the time, or they can be ephemeral, flowing intermittently after periods of rainfall or snowmelt. Such surface waters provide the majority of the water utilized by humans. Some rivers, such as the Colorado River in the western United States, are used so intensively that often no water reaches the sea. Others flowing through hot, dry areas, as, for example, the Lower Nile, became smaller downstream as they lose water to evaporation and groundwater storage.

      Stream discharge is normally expressed in units of volume per unit time (e.g., cubic metres per second), although this is sometimes converted to an equivalent depth over the upstream catchment area. There are a number of techniques for measuring stream discharge. Measurements of velocities using current meters or ultrasonic sounding can be multiplied by the cross-sectional area of flow. Dilution of a tracer can also be used to estimate the total discharge. Weirs of different types are frequently employed at discharge measurement sites. These are constructed so as to give a unique relationship between upstream water level and stream discharge. Water levels can then be measured continuously, usually with a float recorder, to construct a record of discharge over time—namely, a stream hydrograph. Analysis of the hydrographic response to catchment inputs can reveal much about the nature of the catchment and the hydrologic processes within it.

      Stream discharge data are presented in terms of daily, monthly, and annual flow volumes, though for some purposes (e.g., flood routing) shorter time periods may be appropriate. The frequency characteristics of peak discharges and low flows are also of importance to water resource planning. These are analyzed using some assumed probability distribution in a way similar to rainfalls. A time recording of annual maximum flood is usually used in flood-frequency analysis. For design purposes the hydrologist may be asked to estimate the flood with a recurrence interval of 50 or 100 years or longer. There are few discharge records that are longer than 50 years, so such estimates are almost always based on inadequate data.

      Knowledge of the discharge characteristics of catchments is essential to water supply planning and management, flood forecasting and routing, and floodplain regulation. Discharges vary over short lengths of time during storm periods, seasonally with the seasonal changes in evapotranspiration losses, and over longer periods of time as the rainfall regime changes from year to year. Discharge characteristics also vary with climate. In some places discharge represents only a minor component of the catchment water balance, the losses being dominated by evapotranspiration.

      The discharge hydrograph that results from a rainstorm represents the integrated effects of the surface and subsurface flow processes in the catchment. Traditionally, hydrologists have considered the bulk of a storm hydrograph to consist of storm rainfall that has reached the stream primarily by surface routes. Recent work using naturally occurring isotope tracers such as deuterium has shown, however, that in many humid temperate areas the bulk of the storm hydrograph consists of pre-event water. This water has been stored within the catchment between storms and displaced by the rainfall during the storm. This suggests that subsurface flow processes may play a more important role in the storm response of catchments than has previously been thought possible.

Modeling catchment hydrology
      The availability of high-speed computers has resulted in a widespread use of computer models in the analysis and prediction of hydrologic variables for research as well as for practical design and management purposes. These models vary greatly in type and complexity, from simple computer implementations of methods previously based on manual calculations to attempts to solve the nonlinear partial differential equations describing surface and subsurface flow processes that require much computation. All have their practical limitations.

      The simpler models treat the catchment as a single “lumped” (or undifferentiated) unit. It is clearly not possible to describe hydrologic processes in detail in such a model, and most processes are represented as empirical functional relationships between inputs and outputs. Some “lumped” models do not refer to the internal hydrologic processes of the catchment at all but use systems analysis techniques to relate inputs to outputs. The parameters of such computer models are calibrated by fitting the model to simulate a known discharge record. It is consequently very difficult to interpret parameters derived in this manner in a physically meaningful way or to extend the use of the model to sites where there are no discharge records. Parameter values for ungauged sites can sometimes be estimated from empirical relationships between catchment characteristics and parameter values derived from fitting a model at a number of gauged sites. The uncertainties in such a procedure, however, are high.

      The more complex computer models attempt to analyze the internal processes of the hydrologic system in greater detail, taking into account the spatial nature of the catchment, its topography, soils, vegetation, and geology. These are “distributed” models, usually formulated in terms of flow equations for each hydrologic process considered to be important. Some processes such as channel flows and groundwater flows can be described in a reasonably satisfactory way. In the case of other processes, as, for example, flow through the soil and evapotranspiration, hydrologists cannot be so sure of their descriptions. Distributed models tend to have many parameters. In principle, many of these parameters can be measured in the field or can be estimated from the physical characteristics of the catchment. In practice, such models have proved difficult to apply and have not been shown to provide more accurate results than simpler models in spite of their theoretical rigour.

      Most models for hydrologic forecasting in practical use today are deterministic; that is to say, given a sequence of inputs to the model, the outputs are uniquely determined. In a probabilistic description of catchment hydrology, the effects of uncertainty in the model inputs, parameters, or descriptive equations must be reflected in a degree of uncertainty in the outputs. Such a model is known as a stochastic model.

Water quality
      Natural water quality is a dilute solution of elements dissolved from the Earth's crust or washed from the atmosphere. Its ionic concentration varies from less than 100 milligrams per litre in snow, rain, hail, and some mountain lakes and streams to as high as 400,000 milligrams per litre in the saline lakes of internal drainage systems or old groundwaters associated with marine sediments.

      Water quality is influenced by natural factors and by human activities, both of which are the subject of much hydrologic study. The natural quality of water varies from place to place with climate and geology, with stream discharge, and with the season of the year. After precipitation reaches the ground, water percolates through organic material such as roots and leaf litter, dissolves minerals from the soil and rock through which it flows, and reacts with living things from microscopic organisms to humans. Water quality also is modified by temperature, soil bacteria, evaporation, and other environmental factors.

      Pollution is the degradation of water quality by human activities. Pollution of surface and subsurface waters arises from many causes, but it is having increasingly serious effects on hydrologic systems. In some areas the precipitation inputs to the system are already highly polluted, primarily by acids resulting from the combustion of fossil fuels in power generation and automobiles.

      Other serious causes of pollution have been the dumping of industrial wastes and the discharge of untreated sewage into watercourses. Salt spread on roads in winter has resulted in the contamination of subsurface drinking water supplies in certain areas, as, for example, in Long Island, New York. Excess water resulting from deforestation or irrigation return flows that leach salts from soils in semiarid areas are major sources of pollution in the western United States and Western Australia.

Study of lakes
       limnology is concerned with both natural and man-made lakes, their physical characteristics, ecology, chemical characteristics, internal energy fluxes, and exchanges with the environment. It often includes the ecology and biogeochemistry of flowing freshwaters. The study of former lakes is known as paleolimnology. It involves inferring the history of a former lake basin on the basis of the evidence contained in the sediments of the lake bed.

      Lakes may be formed as a result of tectonic activity, glacial activity, volcanism, and by solution of the underlying rock. Man-made lakes or reservoirs may result from the building of a dam within a natural catchment area or as a complete artificial impoundment. In the former case the reservoir may be filled by natural flow from upstream; in the latter the supply of water must be piped or pumped from a surface or subsurface source. The use of reservoir water for water supply, river regulation, or hydroelectric power generation may cause rapid changes in water levels that would not normally occur in a natural lake. In addition, water is usually drawn from a reservoir at some depth, resulting in a shorter residence time relative to an equivalent natural lake.

The history of lakes
      A newly formed lake generally contains few nutrients and can sustain only a small amount of biomass. It is described as oligotrophic. Natural processes will supply nutrients to a lake in solution in river water and rainwater, in the fallout of dust from the atmosphere, and in association with the sediments washed into the lake. The lake will gradually become eutrophic, with relatively poor water quality and high biological production. Infilling by sediments means that the lake will gradually become shallower and eventually disappear. Natural rates of eutrophication are normally relatively slow. Human activities, however, can greatly accelerate the process by the addition of excessive nutrients in wastewater and the residues of agricultural fertilizers. The result may be excessive biomass production, as evidenced by phytoplankton “blooms” and rapid growth of macrophytes such as Eichhornia.

The physical characteristics of lakes
      The most important physical characteristic of the majority of lakes is their pattern of temperatures, in particular the changes of temperature with depth. The vertical profile of temperature may be measured using arrays of temperature probes deployed either from a boat or from a stationary platform. Remote-sensing techniques are being used increasingly to observe patterns of temperature in space and, in particular, to identify the thermal plumes associated with thermal pollution.

      In summer the water of many lakes becomes stratified into a warmer upper layer, called the epilimnion, and a cooler lower layer, called the hypolimnion. The stratification plays a major role in the movement of nutrients and dissolved oxygen and has an important control effect on lake ecology. Between the layers there usually exists a zone of very rapid temperature change known as the thermocline. When the lake begins to cool at the end of summer, the cooler surface water tends to sink because it has greater density. Eventually this results in an overturn of the stratification and a mixing of the layers. Temperature change with depth is generally much smaller in winter. Some lakes, called dimictic lakes, can also exhibit a spring overturn following the melting of ice cover, since water has a maximum density at 4° C.

      A second important characteristic of lakes is the way that the availability of light changes with depth. Light decreases exponentially (as described by Beer's law) depending on the turbidity of the water. At the compensation depth the light available for photosynthetic (photosynthesis) production is just matched by the energy lost in respiration. Above this depth is the euphotic zone, but below it in the aphotic zone phytoplankton—the lowest level in the ecological system of a lake—cannot survive unless the organisms are capable of vertical migration.

      Patterns of sediment deposition in lakes depend on the rates of supply in inflowing waters and on subsurface currents and topography. Repetitive sounding of the lake bed may be used to investigate patterns of sedimentation. Remote sensing of the turbidity of the surface waters also has been used to infer rates of sedimentation, as in the artificial Lake Nasser in Egypt. In some parts of the world where erosion rates are high, the operational life of reservoirs may be reduced dramatically by infilling with sediment.

Water and energy fluxes in lakes
      The water balance of a lake may be evaluated by considering an extended form of the catchment water balance equation outlined above with additional terms for any natural or artificial inflows. An energy balance equation may be defined in a similar way, including terms for the exchange of long-wave and shortwave radiation with the Sun and atmosphere and for the transport of sensible and latent heat associated with convection and evaporation. Heat also is gained and lost with any inflows and discharges from the lake. The energy balance equation controls the thermal regime of the lake and consequently has an important effect on the ecology of the lake.

      An important role in controlling the distribution of temperature in a lake is played by currents due either to the action of the wind blowing across the surface of the lake or to the effect of the inflows and outflows, especially where, for example, a lake receives the cooling water from a power-generation plant. In large lakes the effect of the Earth's rotation has an important effect on the flow of water within the lake. The action of the wind can also result in the formation of waves and, when surface water is blown toward a shore, in an accumulation of water that causes a rise in water level called wind setup. In Lake Erie in North America, increases in water level of more than one metre have been observed following severe storms. After a storm the water raised in this way causes a seiche (an oscillatory wave of long period) to travel across the lake and back. Seiches are distinctive features of such long, narrow lakes as Lake Zürich, when the wind blows along the axis of the lake. Internal seiche waves can occur in stratified lakes with layers of different density.

The water quality of lakes
      The biological health of a lake is crucially dependent on its chemical characteristics. Limnologists and hydrobiologists are attentive to the dissolved oxygen content of the water because it is a primary indicator of water quality. Well-oxygenated water is considered to be of good quality. Low dissolved oxygen content results in anaerobic fermentation, which releases such gases as toxic hydrogen sulfide into the water, with a drastic effect on biological processes.

      Another major concern of limnologists and hydrobiologists is the cycling of basic nutrients within a lake system, particularly carbon, nitrogen, phosphorus, and sulfate. An excess of the latter in runoff waters entering a lake may result in high concentrations of hydrogen ions in the water. Such acid (low values of pH) waters are harmful to the lake biology. In particular, aluminium compounds are soluble in water at low pH and may cause fish to die because of the response induced in their gills.

Study of the oceans and seas
       oceanography is concerned with all aspects of the Earth's oceans and seas. Physical oceanography is the study of the properties of seawater, including the formation of sea ice, the movement of seawater (e.g., waves, currents, and tides), and the interactions between the so-called World Ocean and the atmosphere. Chemical oceanography is the study of the composition of seawater and of the physical, biological, and chemical processes that govern changes in composition in time and space. Marine geology deals with the geologic evolution of the ocean basins, while biological oceanography or marine ecology focuses on the plant and animal life of the sea.

The origin of the ocean basins
      About 71 percent of the Earth's surface is covered by seawater, with the proportion of sea to land being greater in the Southern Hemisphere (four to one) than in the Northern Hemisphere (1.5 to one). Current theories of plate tectonics explain the development of the present ocean basins in terms of a splitting of a large continental landmass brought about by convective circulation within the Earth's mantle. This circulation causes material to rise from the mantle, resulting in the formation of new lithosphere at the mid-oceanic ridges. Continued ascent of material in these areas has resulted in the movement of older material away from the ridges, a process known as seafloor spreading (seafloor spreading hypothesis) (see also above). Clear evidence of the broad symmetry of these movements has been produced by studies of the residual magnetism of the rocks of the seafloor, which exhibit successive zones of magnetic reversals parallel to the mid-oceanic ridges. Near the continental masses some of the oceanic material sinks into the mantle in zones of subduction, which are associated with oceanic troughs, deep-focus earthquakes, and volcanic activity.

The physical properties of seawater
      The physical properties of seawater depend on the chemical constituents dissolved in it. The spatial variability of seawater composition is only partially known, since many areas of the oceans have not been fully sampled. It has been shown that while the salinity of seawater varies from place to place, the relative proportions of the major constituents remain fairly constant. Chlorine accounts for about 55 percent of dissolved solids, sodium 30.6 percent, sulfate 7.7 percent, magnesium 3.7 percent, and potassium 1.1 percent. The average salinity of seawater is about 35 grams of dissolved salts per kilogram of seawater. Higher values occur in areas where evaporation (vaporization) rates are high, such as the Red Sea (41 grams per kilogram), and in areas where relatively pure water freezes out as sea ice, thereby increasing the salinity of the water below. Variations in salinity may be measured indirectly by measurements of the electrical conductivity of seawater, which also yield an accurate estimate of density. With the electrical conductivity method, it is easy to obtain in situ estimates of density and salinity. Using temperature sensors it is also possible to obtain in situ measurements of seawater temperature, the other very important physical characteristic of seawater. It is primarily variations in temperature and density that drive the circulation of water in the oceans.

      The density of seawater in high latitudes tends to gradually increase with depth. In tropical areas there is usually a well-mixed layer of uniform density close to the surface above the pycnocline, a layer in which density increases extremely rapidly. Below the pycnocline density continues to increase but much more slowly. The pycnocline is a very stable layer that acts as a barrier to the transfer of water and energy between the surface and subsurface layers.

      Seawater has a higher specific heat (about 0.95) than land (0.5 on average), so that changes in the temperature of the oceans are much slower than on land. In fact, the oceans represent a vast store of heat that exercises an important, but not fully understood, control on the variability of climate regimes observed on land. The temperature distribution at the surface of the open seas tends to follow lines of latitude, with warmer waters in tropical regions. Close to the landmasses, the lines of equal temperature (isotherms) become deflected, indicating warm and cold coastal currents. Below the surface there is generally a well-mixed layer of fairly uniform temperature about 50 to 200 metres thick below which lies the thermocline, which may extend to depths of 1,000 metres. Below the thermocline, temperature decreases slowly with depth. A typical temperature profile in equatorial regions would be 20° C down to 200 metres, 8° C at 500 metres, 5° C at 1,000 metres, and 2° C at 4,000 metres.

      Plotting the temperature and salinity of a sample of seawater on a graph with linear axes (a T–S diagram) can be a powerful research tool. A mass of fully mixed water having a homogeneous distribution of temperature and salinity would plot as a single point on a T–S diagram. For actual water masses it is common to find that points plotted for samples taken from different depths plot as a curve (sometimes complex) on the diagram. Such curves provide an indication of the mixing between different water masses that is taking place in the profile. T–S curves also are useful for uncovering errors in data.

The circulation of the oceans
      One major cause of the circulation of waters in the oceans is the difference in the energy budget between the tropics and the poles (the thermohaline circulation). While it is now thought that differences in solar heating have a relatively minor direct effect on ocean circulation, the formation of sea ice and the loss of heat from the oceans at the poles causes a movement of colder, denser water toward the Equator at depth. The major surface currents of the oceans are driven by the surface shear stresses (shear stress) imposed by the wind. These motions are influenced by the topography of the ocean basins and the Coriolis effect due to the Earth's rotation, so that in the Northern Hemisphere the moving water becomes deflected toward the right, while in the Southern Hemisphere toward the left. This results in major clockwise circulations in the North Atlantic (including the Gulf Stream) and the North Pacific, with counterclockwise circulations in the South Atlantic, South Pacific, and Indian oceans. Within the tropics there tends to be a pattern of westward-flowing currents in both the Northern and Southern hemispheres, with an eastward-flowing countercurrent close to the Equator itself. There may also be an eastward-flowing undercurrent at depth. It is certain that there is still much to be learned about the details of ocean circulation, particularly at depth.

      There are two fundamental approaches to measuring ocean currents: (ocean current) the Lagrangian and Eulerian methods. In the Lagrangian method individual parcels of water are tracked using floats or buoys. Satellite-tracked buoys equipped with radio transmitters are now commonly used in the study of surface currents. Currents at depth may be studied with Swallow floats, which are adjusted to be neutrally buoyant at a certain density of seawater. Tracer techniques, such as those involving the use of dyes and discharges of pollutants, may also be employed to track flowing water at least in coastal areas. The greatest number of measurements of surface currents by the Lagrangian method, however, have come from the records of the drift of ships contained in navigation logs.

      The Eulerian method consists of measuring the velocity of flow past a fixed point (a moored ship, anchored line, or structure) with a current meter, of which there are a number of different types. Flow velocities may be measured as a function of both depth and time at any site.

      An indirect method for estimating current velocities is the geostrophic (geostrophic motion) method. It is based on the fact that the movement of water masses away from the sea surface and any solid boundary can be assumed to be frictionless and unaccelerated. Under such conditions the pressure gradient and the effects of gravity and Coriolis forces should balance exactly. The expected rates and directions of flow can then be computed theoretically. It has been shown that the geostrophic currents are good approximations of actual currents.

      The hydrodynamics of ocean currents can be described by the dynamic equations of fluid flow. The advent of high-speed digital computers has made it possible to obtain approximate numerical solutions to these equations for many problems of practical interest, including the transient effects of tides. The formation and propagation of waves, together with their refraction in shallow coastal waters, also can be computed numerically.

Biogeochemical cycles (biogeochemical cycle) in the oceans
      The ocean is a great store of chemicals that receives inputs from rivers and the atmosphere and, on average, loses equal amounts to sedimentary deposits on the ocean floor. Biological processes play a large part in processing the chemicals received and in maintaining the remarkable consistency in the composition of seawater. Fortunately this consistency does not extend to all the elements found in seawater. Concentrations of some of the minor, or trace, elements can be used to infer the mixing, biological, and sedimentation processes that occur. Throughout the oceans the major variations in composition are in the upper layers, where the greatest biological activity is found.

      The use of a number of different radioisotopes in dating sediments and calculating rates of sedimentation and mixing within the oceans have been important in studying the biogeochemical cycles of the oceans. A particularly interesting use of radiometric dating was in investigating the formation of the manganese nodules that occur on certain segments of the seabed and in the underlying sediments. These nodules consist primarily of manganese and iron oxides, even though concentrations of these elements in seawater are very low. Dating techniques have shown that the growth rates of the nodules are on the order of three millimetres per 1,000 years, or 1,000 times less than the accumulation rate of the sediments on which they lie.

Remote sensing of the oceans
      One of the fundamental problems faced by oceanographers is the sheer size of the oceans and the consequent need to rely on special surface vessels and submersibles for direct measurements. It can be very costly to operate either type of vessel on long deep-sea expeditions. Moreover, observations from such craft can provide only a partial picture of oceanic phenomena and processes in terms of both space and time. Consequently, there has been considerable interest in taking advantage of remote-sensing techniques in oceanography, particularly those that use satellites (satellite observatory). Remote sensing allows measurements to be made of vast areas of ocean repeated at intervals in time.

      The first satellite devoted to oceanographic observations was Seasat, which orbited the Earth for three months in 1978. Its polar orbit made it possible to provide coverage of 95 percent of the major oceans every 36 hours. Seasat carried radiometers for observations at visible, infrared, and microwave wavelengths, along with radar scatterometers, imaging radar, and an altimeter. This array of instruments yielded much data, including an estimation of sea-surface temperatures, net radiation inputs to the sea surface, wave heights, and wind speeds close to the sea surface. In addition, patterns of near-surface sediment movement and other information were derived from an analysis of the satellite images. For further information about remote-sensing techniques used in oceanographic research, see undersea exploration: Acoustic and satellite sensing (undersea exploration).

Study of ice on the Earth's land surface
       glaciology deals with the physical and chemical characteristics of ice on the landmasses; the formation and distribution of glaciers and ice caps; the dynamics of the movement of glacier ice; and interactions of ice accumulation with climate, both in the present and in the past. Glacier ice covers only about 10 percent of the Earth's land surface at the present time, but it was up to three times as extensive during the Pleistocene Ice Age.

The accumulation of ice
      Glacier ice forms from the accumulation of snow over long periods of time in areas where the annual snowfall is greater than the rate of melting during summer. This accumulated snow gradually turns into crystalline ice as it becomes buried under further snowfalls. The process can be accelerated by successive melting and freezing cycles. The crystalline ice incorporates some of the air of the original snow as bubbles, which only disappear at depths exceeding about 1,000 metres. Successive annual layers in the ice often can be distinguished by differences in crystalline form, by layers of accumulated dust particles that mark each summer melt season, or by seasonal differences in chemical characteristics such as oxygen isotope ratios. The layers become thinner with depth as the density of the ice increases.

      Oxygen isotope ratios indicate the temperature at which the snow making up the ice was formed. Seasonal variations in isotope ratios not only allow annual layers to be distinguished but also can be used to determine the residence times of melt waters within an ice mass. Long-term variations in isotope ratios can be employed to ascertain temperature variations related to climatic change. An ice core of 1,390 metres taken at Camp Century in Greenland has been used in this way to indicate temperatures during the past 120,000 years, and it shows clearly that the last glacial period extended from 65,000 to about 10,000 years ago. These results have been corroborated by measurements of additional cores from Greenland and Antarctica. In spite of the fact that temperatures may remain below freezing throughout the year, ice accumulation over much of Antarctica is very slow, since precipitation rates are low (they are equivalent to those in many desert areas).

      On any glacier there is a long-term equilibrium between accumulation and ablation (losses due to melt runoff and other processes). Continued accumulation eventually causes ice to move downhill, where melt rates are higher. The elevation at which accumulation balances losses changes seasonally as well as over longer periods. In many areas of the world, the annual meltwaters are a crucial part of the water resources utilized by man. In the past it was very difficult to predict amounts of spring melt runoff because of the difficulties in assessing snow accumulation in mountainous terrain. Remote-sensing techniques now allow accumulation over much larger areas to be estimated, and they also offer the possibility of updating those estimates during the melt season.

The movement of glaciers
      The mechanisms by which a large mass of ice can move under the effects of gravity have been debated since about 1750. It is now known that some of this movement is due to basal sliding but that the ice itself, a crystalline solid close to its melting point, can flow, behaving like other crystalline solids such as metals. Early measurements of flow velocities were based entirely on surveys of surface stakes, a technique still used today. During the early 19th century the Swiss geologist Louis Agassiz showed that the movement was fastest in the central part of a glacier. Rates of movement are fastest in the temperate glaciers, which have temperatures close to the melting point of ice and include about 1 percent liquid water. (This water constitutes a layer at the bottom of such an ice mass.) Velocities vary through time, quite dramatically at times. Certain glaciers (e.g., the Muldrow and Variagated glaciers in Alaska) are subject to surges of very rapid velocities at irregular periods. The causes of these catastrophic advances are still not well understood.

      Techniques for investigating the movement of ice in the field include studies of the deformation of vertical boreholes and lateral tunnels dug into the ice. The internal structure of glaciers and the Greenland and Antarctic ice caps have also been examined by means of radar sounding. This method works best in cold glaciers where the ice is below its freezing point.

      Indirect evidence of the patterns of movement is obtained from the characteristic landforms associated with glaciers, particularly scratched or striated bedrock and moraines composed of rock debris. Such forms also allow the interpretation of former patterns of movement in areas no longer covered by ice.

Practical applications

Development and management of water resources
      Water is essential to many of humankind's most basic activities—agriculture, forestry, industry, power generation, and recreation. As the hydrologic sciences provide much of the knowledge and understanding on which the development and management of available water resources are based, they are of fundamental importance.

      In 1965 the United Nations Educational, Scientific and Cultural Organization (UNESCO) initiated the International Hydrological Decade (IHD), a 10-year program that provided an important impetus to international collaboration in hydrology. Considerable progress was made in hydrology during the IHD, but much still remains to be done, both in the basic understanding of hydrologic processes and in the development and conservation of available water resources. Many developing countries remain highly susceptible to diseases related to a lack of water supplies of good quality and to the effects of drought. This has been cruelly highlighted in recent times by the severe droughts in the Sahel region of Africa in the periods 1969–74 and 1982–85 (see below).

      In the developed countries the ready availability of a supply of good quality water is expected. Yet, even in the most advanced countries, many water sources are not being used wisely. Groundwater levels in certain areas have fallen dramatically since the 1940s, leading to ever higher pumping costs. Other surface and subsurface water sources are becoming increasingly polluted by urban, agricultural, and industrial wastes in spite of increased expenditure on waste-water treatment and legislation of minimum quality standards. Humankind continues to use the oceans as a vast dumping ground for its waste products, even though little is known about the effects of such wastes on marine ecosystems. It is no exaggeration to say that the misuse of water resources will become a major source of conflict between communities, states, and nations in the years to come. Already several disputes over rights to clean water have taken on international significance.

      Since the early 1980s the acid rain problem has assumed scientific, economic, and political prominence in North America and Europe. This major environmental problem serves to illustrate the interdependence of the various hydrologic sciences with other scientific disciplines and human activities. As was noted earlier, waste gases (primarily oxides of sulfur and nitrogen) enter the atmosphere from the burning of fossil fuels by automobiles and electric power plants. These gases combine with water vapour in the atmosphere to form sulfuric and nitric acids. When rain or some other form of precipitation falls to Earth, it is highly acidic (often with a pH value of less than 4). The resultant acidification of surface and subsurface waters has been shown to have detrimental effects on the ecology of affected catchments. Areas underlain by slowly weathering bedrock, such as in Scandinavia, the Adirondack Mountains of New York, and the Canadian Shield in Quebec are particularly susceptible. Many lakes in these areas have been shown to be biologically dead. There also is evidence that the growth of trees may be affected, with consequent economic ramifications where forestry is a major activity. The areas most greatly affected may be far downwind of the source of the pollution. A number of countries have claimed that the major sources of acid rain affecting their streams and lakes lie outside their borders.

      Research has revealed that in an area susceptible to the effects of acid rain short-lived events can have a particularly damaging effect. These “acid shocks” may be due to inputs of highly acid water from a single storm or to the first snowmelt outflows in which the major part of the pollutant input accumulated over the winter is concentrated. The way in which the chemistry of the input water is modified in its flow through the catchment depends both on the nature of the soils and rocks in the catchment and on the flow paths taken through the system. These interactions are at present poorly understood. It is likely, however, that the attempt to understand the chemical processes within the different flow paths will lead to significant improvements in scientific understanding of catchment hydrology.

Concern over groundwater quantity and quality
      Groundwater problems are becoming increasingly serious in many areas of the world. Rapid increases in the rates of pumping of groundwater in many aquifers has caused a steady lowering of water table levels where extraction has exceeded rates of recharge. A notable example is the Ogallala aquifer, a sandy formation some 100 metres thick, which lies beneath the Great Plains from South Dakota to Texas. It has been estimated that as much as 60 percent of the total storage of this huge aquifer has already been extracted primarily for agricultural use. The remaining water, if it continues to be mined in this way, will become more and more expensive to extract. This situation points out the importance of understanding groundwater flow and recharge processes in complex heterogeneous formations so that safe yields of aquifers can be properly predicted.

      There are many causes of groundwater pollution; most are the accidental or incidental consequences of human activities (e.g., pollution resulting from the use of artificial fertilizers or saltwater intrusion into coastal aquifers due to excessive pumping). In some cases, however, groundwater may be contaminated because of planned human effort. Subsurface repositories of water, for example, have been considered as possible receptacles for waste products, including radioactive materials. This has resulted in both experimental and model studies of water flows in poorly permeable massive rocks that would be used to store such wastes. The effects of joints and fractures on the very slow transport of contaminants over long periods of time in such rocks is as yet uncertain but must be clarified if this form of storage is to be proved safe.

Studying the causes of droughts and other climatic patterns
      Another subject still poorly understood is the occurrence of droughts in areas of highly variable rainfall. In the early 1970s and again in the early 1980s the Sahel region of Africa suffered periods of severe drought, resulting in widespread famine and death. There have been many Sahelian droughts before, but the consequences of the recent droughts have been exacerbated by increased populations of people and grazing animals. The combination of drought and population growth results in desertification. It remains an unanswered scientific question as to whether the deterioration of the Sahel and other marginal lands is part of a long-term natural change or whether it is a result of human activities.

      Some evidence for long-range interactions in the occurrence of droughts and other climatic regimes comes from studies of the ocean currents. It is known that the oceans are a major controlling influence on climate, but the processes involved remain the subject of active research. Some clues have been revealed by studies of El Niño, a minor branch of the Pacific Equatorial Countercurrent that flows south along the coasts of Colombia and Ecuador where it meets the northward-flowing Peru Current. The cold Peru Current keeps rainfall along the coastal area of Peru very low but maintains a rich marine life, which in turn supports major bird populations and a fishing industry. In certain years El Niño becomes much stronger, forcing the Peru Current to the south. Storms rake the coast, causing flooding and erosion. The sudden change in sea temperatures causes dramatic decreases in plankton production and, consequently, in fish and bird populations. Catastrophic El Niño events occurred in 1925, 1933, 1939, 1944, 1958, and 1983. It is thought that the global changes associated with this last event included severe droughts in Australia and Central America and floods in the southwestern United States and Ecuador. Explanations of the El Niño events have invoked both local and long-range interactions in the circulation of the Pacific winds and currents. The study of such dramatic events, enhanced by remote sensing and computer modeling, is a major stimulus to understanding the general circulation of the Earth's atmosphere and oceans.

Keith J. Beven

Additional Reading
An excellent introduction is provided by Thomas Dunne and Luna B. Leopold, Water in Environmental Planning (1978); while basic concepts in engineering hydrology are covered by Ray K. Linsley, Jr., Max A. Kohler, and Joseph L.H. Paulhus, Hydrology for Engineers, 3rd ed. (1982). In more specialized areas, S. Lawrence Dingman, Fluvial Hydrology (1984), treats basic hydraulic principles; R. Allen Freeze and John A. Cherry, Groundwater (1979), deals with subsurface flow processes and water quality; and M.J. Kirkby (ed.), Hillslope Hydrology (1978), is concerned with experimental and modeling studies of catchment processes. Modeling is dealt with further in M.G. Anderson and T.P. Burt, Hydrological Forecasting (1985). (Limnology): Broad introductions are provided by Abraham Lerman (ed.), Lakes—Chemistry, Geology, Physics (1978); and Robert G. Wetzel, Limnology, 2nd ed. (1983). Some engineering aspects of man-made lakes are covered in B. Henderson-Sellers, Engineering Limnology (1984). (Oceanography): Excellent introductions are Peter K. Weyl, Oceanography: An Introduction to the Marine Environment (1970); Keith Stowe, Ocean Science, 2nd ed. (1983); George L. Pickard and William J. Emery, Descriptive Physical Oceanography: An Introduction, 4th ed. (1982); and David Tolmazin, Elements of Dynamic Oceanography (1985). The geology of the ocean basins is the subject of Francis P. Shepard, Geological Oceanography: Evolution of Coasts, Continental Margins and the Deep-Sea Floor (1977); and the chemistry of the oceans is covered in J.P. Riley and R. Chester, Introduction to Marine Chemistry (1971). (Glaciology): The classic text is Louis Lliboutry, Traité de glaciologie, 2 vol. (1964). The physical aspects of ice are covered in W.S.B. Paterson, The Physics of Glaciers, 2nd ed. (1981). Other useful texts are Samuel C. Colbeck (ed.), Dynamics of Snow and Ice Masses (1980); and D.E. Sugden and B.S. John, Glaciers and Landscape (1976, reprinted 1979).Keith J. Beven

* * *


Universalium. 2010.

Игры ⚽ Нужна курсовая?

Look at other dictionaries:

  • Earth Sciences — ▪ 2009 Introduction Geology and Geochemistry       The theme of the 33rd International Geological Congress, which was held in Norway in August 2008, was “Earth System Science: Foundation for Sustainable Development.” It was attended by nearly… …   Universalium

  • Earth and Space Sciences — ▪ 1996 Introduction GEOLOGY AND GEOCHEMISTRY       In 1995 significant developments took place in the realm of geologic mapping, which provides the foundation for the presentation and comparison of data in the Earth sciences. The most important… …   Universalium

  • hydrology — hydrologic /huy dreuh loj ik/, hydrological, adj. hydrologically, adv. hydrologist, n. /huy drol euh jee/, n. 1. the science dealing with the occurrence, circulation, distribution, and properties of the waters of the earth and its atmosphere. 2.… …   Universalium

  • BYDIN — RUSSIA (see also List of Individuals) .. 1898 Elcha/RU .. Leningrad/RU Fedor Ignatievich Bydin was born in the Smolensk Region of Russia. He graduated in 1927 from the Moscow Higher Technical School as a hydraulic engineer. Then he collaborated… …   Hydraulicians in Europe 1800-2000

  • American Meteorological Society — The American Meteorological Society (AMS) promotes the development and dissemination of information and education on the atmospheric and related oceanic and hydrologic sciences and the advancement of their professional applications. Founded in… …   Wikipedia

  • History of geophysics — The historical development of geophysics has been motivated by two factors. One of these is research curiosity of humankind related to Planet Earth and its several components, its events and its problems. The second one is economically usage of… …   Wikipedia

  • oceanography — oceanographer, n. oceanographic /oh sheuh neuh graf ik/, oceanographical, adj. oceanographically, adv. /oh sheuh nog reuh fee, oh shee euh /, n. the branch of physical geography dealing with the ocean. [1855 60; OCEAN + O + GRAPHY] * * *… …   Universalium

  • American Geophysical Union — Infobox Organization name = American Geophysical Union image border = size = caption = msize = mcaption = abbreviation = AGU motto = formation = 1919 extinction = type = Non profit status = purpose = Geophysics, and many other fields in Earth and …   Wikipedia

  • Texas A&M College of Geosciences — TAMU College Infobox name = College of Geosciences dean = Dr. Björn Kjerfve students = 738 undergrads = 480 masters = 131 phd = 127 faculty = 109 website = [http://geosciences.tamu.edu/ http://geosciences.tamu.edu/] Texas A M College of… …   Wikipedia

  • Desert Research Institute — Logo of the Desert Research Institute The Desert Research Institute (DRI) is the nonprofit research campus of the Nevada System of Higher Education (NSHE), the organization that oversees all publicly supported higher education in the U.S. state… …   Wikipedia

Share the article and excerpts

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