Lake

Lake
/layk/, n.
Simon, 1866-1945, U.S. engineer and naval architect.

* * *

I
Relatively large body of slow-moving or standing water that occupies an inland basin.

Lakes are most abundant in high northern latitudes and in mountain regions, particularly those that were covered by glaciers in recent geologic times. The primary sources of lake water are melting ice and snow, springs, rivers, runoff from the land surface, and direct precipitation. In the upper part of lakes there is a good supply of light, heat, oxygen, and nutrients, well distributed by currents and turbulence. As a result, a large number of diverse aquatic organisms can be found there. The most abundant forms are plankton (chiefly diatoms), algae, and flagellates. In the lower levels and in the sediments, the main forms of life are bacteria. See also limnology.
II
(as used in expressions)
Lake Ijssel
Albert Lake
Alexandrina Lake
Assal Lake
Athabasca Lake
Atitlán Lake
Avernus Lake
Baikal Lake
Lake Baykal
Balaton Lake
Balkhash Lake
Lake Balqash
Bangweulu Lake
Biwa Lake
Bras d'Or Lake
Chad Lake
Champlain Lake
Chapala Lake
Chiem Lake
Chilwa Lake
Chuzenji Lake
Como Lake
Constance Lake
Dian Lake
Edward Lake
Ellesmere Lake
Er Lake
Erie Lake
Eyasi Lake
Eyre Lake
Garda Lake
Gatún Lake
Geneva Lake
George Lake
Hammar Lake
Huron Lake
Ilmen Lake
Iseo Lake
Itasca Lake
Izabal Lake
Khanka Lake
Kivu Lake
Ladoga Lake
Lanao Lake
Llanquihue Lake
Lugano Lake
Magadi Lake
Maggiore Lake
Malawi Lake
Lake Nyasa
Managua Lake
Manitoba Lake
Mapam Lake
Maracaibo Lake
Mead Lake
Memphremagog Lake
Michigan Lake
Mjosa Lake
Moeris Lake
Mweru Lake
Nahuel Huapí Lake
Naivasha Lake
Näsi Lake
Lake Näsijärvi
Nasser Lake
Lake Nubia
Nemi Lake
Neuchâtel Lake
Ngami Lake
Nicaragua Lake
Nipigon Lake
Nipissing Lake
Okeechobee Lake
Onega Lake
Ontario Lake
Ozarks Lake of the
Päijänne Lake
Peipsi Lake
Petén Itzá Lake
Pielinen Lake
Pontchartrain Lake
Prespa Lake
Rudolf Lake
Lake Turkana
Rukwa Lake
Saimaa Lake
Saint Clair Lake
Scutari Lake
Sevan Lake
Siljan Lake
Simcoe Lake
Superior Lake
Lake Bombon
Tahoe Lake
Tai Lake
Tana Lake
Tanganyika Lake
Taupo Lake
Texcoco Lake
Tiberias Lake
Titicaca Lake
Todos los Santos Lake
Towada Lake
Trasimeno Lake
Urmia Lake
Valencia Lake
Van Lake
Victoria Lake
Volta Lake
Winnipeg Lake
Winnipegosis Lake
Winnipesaukee Lake
Woods Lake of the
Yojoa Lake
Ysyk Lake
Lake Issyk Kul
Zürich Lake
Chargoggagoggmanchauggauggagoggchaubunagungamaugg Lake
Lake Webster
Lucerne Lake
Vierwaldstättersee Lake of the Four Forest Cantons
Great Lakes trout

* * *

▪ physical feature
Introduction
 any relatively large body of slowly moving or standing water that occupies an inland basin of appreciable size. Definitions that precisely distinguish lakes, ponds, swamps, and even rivers and other bodies of nonoceanic water are not well established. It may be said, however, that rivers and streams are relatively fast moving; marshes and swamps contain relatively large quantities of grasses, trees, or shrubs; and ponds are relatively small in comparison to lakes. Geologically defined, lakes are temporary bodies of water. For a list of the major natural lakes of the world, see below (lake).

      This article treats lake basins and sedimentation; the physical and chemical properties of lake waters; lake currents, waves, and tides; and the hydrologic balance of lakes. For information on related systems, see the article river. The place of lakes within the hydrologic cycle is further dealt with in hydrosphere, as are certain aspects of lake sedimentation and water chemistry. See inland water ecosystem (boreal forest) for information on lacustrine life-forms.

General considerations

Occurrence
      Within the global hydrologic cycle, freshwater lakes play a very small quantitative role, constituting only about 0.009 percent of all free water, which amounts to less than 0.4 percent of all continental fresh water. Saline lakes and inland seas contain another 0.0075 percent of all free water. Freshwater lakes, however, contain well over 98 percent of the important surface waters available for use. Apart from that contained in saline bodies, most other continental waters are tied up in glaciers and ice sheets and the remainder is in groundwater.

      Four-fifths of the 125,000 cubic kilometres (30,000 cubic miles) of lake waters occur in a small number of lakes, perhaps 40 in all. Among the largest are Lake Baikal (Baikal, Lake), in Central Asia, containing about 23,000 cubic kilometres (5,500 cubic miles) of water; Lake Tanganyika (Tanganyika, Lake) (19,000 cubic kilometres [4,600 cubic miles]), in eastern Africa; and Lake Superior (Superior, Lake) (12,000 cubic kilometres [2,900 cubic miles]), one of the Great Lakes of North America. The Great Lakes contain a total of about 25,000 cubic kilometres (6,000 cubic miles) of water and, together with other North American lakes larger than 10 cubic kilometres (2 cubic miles), constitute about one-fourth of the world's lake waters. The Caspian Sea, though not considered a lake by some hydrologists, is the world's largest inland sea. Located in Central Asia, the Caspian Sea has an area of about 386,000 square kilometres (149,000 square miles).

      Although lakes are to be found throughout the world, the continents of North America, Africa, and Asia contain about 70 percent of the total lake water, the other continents being less generously endowed. A fourth of the total volume of lake water is spread throughout the world in uncounted numbers of small lakes. Anyone who has flown over much of the Canadian (Canada) plains area cannot help but be struck by the seemingly endless skein of lakes and ponds covering the landscape below. Though the total volume of water involved is comparatively small, the surface area of lake water is substantial. The total surface area of all Canadian lakes has been estimated to exceed the total surface area of the province of Alberta. The U.S. state of Alaska has more than 3,000,000 lakes with surface areas greater than 8 hectares (20 acres).

      The larger, deeper lakes are a significant factor within the cycle of water—from rain to surface water, ice, soil moisture, or groundwater and thence to water vapour. These lakes receive the drainage from vast tracts of land, store it, pass it on seaward, or lose it to the atmosphere by evaporation (vaporization). On a local basis, even the smaller lakes play an important hydrologic role. The relatively high ratio of exposed surface area to the total water volume of these lakes accentuates their effectiveness as evaporators. In some cases the efficiency of lakes in losing water to the atmosphere is locally undesirable, because of public and industrial requirements for lake water. A striking example of this condition is the Aral Sea, located in Central Asia. Although it is still one of the world's largest bodies of inland water, in the second half of the 20th century its area was reduced by two-fifths and its mean surface level had dropped by more than 12 metres (40 feet), as a result primarily of the diversion of the Syr Darya and Amu Darya rivers for irrigating adjoining fields. In some basins (e.g., the Chad basin in Africa), lakes are the terrestrial end point of the hydrologic cycle. With no outflow downstream toward the oceans, these closed lakes swell or recede according to the balance of local hydrologic conditions.

Uses and abuses of lakes
      In today's industrial societies, requirements for water—much of which is derived from lakes—include its use for dilution and removal of municipal and industrial wastes, for cooling purposes, for irrigation, for power generation, and for local recreation and aesthetic displays. Obviously, these requirements vary considerably among regions, climates, and countries.

      In another vein, it is convenient to use water to dilute liquid and some solid wastes to concentrations that are not intolerable to the elements of society that must be exposed to the effluent or wish to use it. The degree of dilution that may be acceptable varies from situation to situation and is often in dispute. In some cases, dilution is used purely to facilitate transport of the wastes to purification facilities (sewage system). The water may then be made available for reuse.

      Lake water is also used extensively for cooling purposes. Although this water may not be affected chemically, its change in thermal quality may be detrimental to the environment into which it is disposed, either directly, by affecting fish health or functions, or indirectly, by causing excessive plant production and ultimate deoxygenation due to biological decay. Both fossil- and nuclear-fuelled power plants are major users of cooling water. Steel mills and various chemical plants also require large quantities.

Economy and ecology
      Concern with thermal pollution of surface waters is concentrated principally on rivers and small lakes. With power requirements in modern societies increasing by about 7 percent per year, however, some apprehension has been expressed about the future thermal loading of even the largest lakes. It was predicted that thermal inputs to each of the North American Great Lakes will increase by nearly 11 times during the last three decades of the 20th century. In terms of energy to be disposed in this fashion, the numbers are staggeringly large. These lakes have such large volumes, however, and such large surface areas (from which much of the heat goes into the atmosphere) that there is some question about the nature and magnitude of the actual effects.

      The economic importance of waterways as communication links is enormous. In the earliest times, when travel by many societies was substantially by water, travel routes became established that resulted in relationships between cultural factors and surface hydrology networks. Today, river and lake systems serve as communication links and play an important role in shipping because of the large cargo capacities of merchant vessels and the still fairly uncongested condition of inland waterways. Oceanic shipping lanes play the major role, but river and lake systems, which link inland ports with the oceans, have been key factors in the rates of economic growth of many large inland ports.

      Commercial fisheries and other food industries reap great harvests from the major lakes of the world. The quality of the fish catch (commercial fishing) has steadily decreased, however, as a result of pollution in many lakes, with the more desirable species becoming less plentiful and the less desirable species gradually dominating the total. Other commercial harvests from lakes include waterfowl, fur-bearing mammals, and some plant material, such as rice.

      Each of the uses described has associated with it the means for abuse of the very characteristics of lakes that make them desirable. Wise management of natural resources has never been humankind's forte. Municipalities and industries have polluted lakes chemically and thermally, the shipping that plies large inland water bodies leaves oil and other refuse in its wake, water used for irrigation often contains chemical residues from fertilizers and biocides when it is returned to lakes, and the populace that so desperately demands clean bodies of water for its recreation often ignores basic sanitary and antipollution practices, to the ultimate detriment of the waters enjoyed.

Problems and effects
      Among the major problems affecting the optimum utilization and conservation of lake waters are eutrophication (aging processes), chemical and biological poisoning, and decreases in water volumes. In the former case, discussed in more detail later, the enrichment of lakes with various nutrients supports biological productivity to an extent in which the ultimate death and decay of biological material places an excessive demand on the oxygen content, resulting in oxygen depletion in the worst cases. Phosphates and nitrates are two of the types of nutrients that are most important in this connection, particularly since they are often introduced in critical quantities in waste effluents from human sources. Other examples of chemical pollution of lakes include the introduction of DDT and other pesticides and heavy metals such as mercury. Bacteriological contamination of lake waters resulting in levels that constitute a hazard to health is another common result of disregard for the environment.

      Water-quantity problems are complex, being related to natural vagaries of supply and levels of consumptive utilization of water. In the latter case, the percentage of water returned to the source after utilization varies with the use. The largest losses are due to actual water diversions and processes that result in evaporative losses. The use of large quantities of lake water for cooling purposes by industry and utilities, for example, may raise lake temperatures near the effluents sufficiently to cause increased evaporation. The use of certain types of cooling towers results in even larger losses. Some of the water evaporated will stay within the lake basin, but some will be lost from it.

      Another example of this type of loss is connected with the possible application of weather-modification (weather modification) techniques to alleviate the heavy lake-effect snowfalls experienced along the lee shores of large lakes in intermediate latitudes. Redistribution of precipitation always raises the possibility of redistribution of water among various basins.

      Lake-effect snowfall is just one example of the influence of lakes on local climate. The ability of large bodies of water to store heat during heating periods and to lose it more gradually than the adjacent landmasses during cooling periods results in a modifying influence on the climate. Because of this propensity, a lake cools air passing over it in summer and warms air passing over it in winter. Consequently, the predominantly downwind side of a lake is more influenced by the ameliorating effects of a lake.

      In most instances, moisture is also passed to the atmosphere. In summer, lake cooling serves to stabilize the air mass, but winter heating tends to decrease stability. The moisture-laden, unstable winter flows off lakes produce so-called snowbelts, which affect downwind cities. The snowbelts are usually of limited extent, often within about a kilometre of the lake shore.

Robert K. Lane Ed.

Lake basins

Classification of basins
      The name given to the study of lakes is limnology. Limnologists have used several criteria for the development of systems for classifying lakes and lake basins (drainage basin) but have resorted particularly to the mechanisms that have produced lake basins (tectonic basins and rift valleys). These have been summarized and examined in A Treatise on Limnology, by the American limnologist G.E. Hutchinson, which includes treatment of tectonism, volcanism, landslides, glaciation, solution, river action, wind action, coastline building, organic accumulation, animal activity, meteoritic impact, and human activity.

Basins formed by tectonism, volcanism, and landslides
      Tectonism (diastrophism)—or movements of the Earth's crust—have been responsible for the formation of very large basins (hydrosphere). During the late Miocene (the Miocene Epoch includes the time interval from 26,000,000 to 7,000,000 years ago), broad, gentle earth movements resulted in the isolation of a vast inland sea across southern Asia and southeastern Europe. Through most of the Tertiary Period (from 65,000,000 to 2,500,000 years ago), sub-basins developed that gradually were characterized by a great range of salinities. Resumption of communication with the oceans occurred later, and there is evidence of considerable variation in water levels. The present remnants of these inland bodies of water include the Caspian Sea and the Aral Sea, along with numerous smaller lakes. The Black Sea, which was also once part of this large inland basin, is now in direct communication with the oceans.

      In some cases, elevated land areas may already contain depressions that eventually form lake basins. Lake Okeechobee, Florida, is cited as being such a basin, formed by uplift of the ocean floor.

      Tectonic uplift may interfere with natural land-drainage patterns in such a way as to produce lake basins. The Great Basin of South Australia, some of the lakes in Central Africa (e.g., Lakes Kioga and Kwania), and to some extent Lake Champlain, in the northeastern United States, are examples of this mechanism. Land subsidence, due to earthquake activity, also has resulted in the development of depressions in which lakes have evolved. Many such cases have been reported within the past 300 years.

      The damming of valleys as a result of various tectonic phenomena has resulted in the formation of a few lake basins, but faulting, in its great variety of forms, has been responsible for the formation of many important lake basins. Abert Lake, in Oregon, lies in the depression formed by a tilted fault block against the higher block. Indeed, many lakes in the western United States are located in depressions formed through faulting, including Lake Tahoe, in the Sierra Nevada, California. Great Salt Lake, Utah, and other nearby salt lakes are remnants of a large Pleistocene lake, Lake Bonneville (Bonneville, Lake), which was formed at least partly by faulting activity (the Pleistocene Epoch includes the interval from about 2,500,000 to about 10,000 years ago).

      In other parts of the world, faulting has also played an important role in basin formation. Lake Baikal (Baikal, Lake) and Lake Tanganyika, the two deepest lakes in the world, occupy basins formed by complexes of grabens (downdropped faulted blocks). These lakes are among the oldest of modern lakes, as are other graben lakes, particularly those within the East African Rift System, which extends through the East African lake system and includes the Red Sea (see further tectonic landforms: Rift valleys (tectonic basins and rift valleys)).

      Basins formed from volcanic activity are also greatly varied in type. The emanation of volcanic material from beneath the surface can be explosive, or it can issue in a gentle and regular manner. This range of activity and the variation of types of material which may be involved produce many different types of basins.

      One broad category includes those occupying the actual volcanic craters or their remnants. Crater lakes may occupy completely unmodified cinder cones, but these are rare. Craters caused by explosions or by the collapse of the roofs of underground magma (molten silica) chambers and those caused by explosion of new volcanic sources and that are built of nonvolcanic material are other examples. The latter are termed maars, following the local name for such forms in Germany. They are found, however, in several locations, including Iceland, Italy, and New Zealand. The maars of the volcanic district of Eifel, West Germany, are among the best known of these formations.

      The collapse of magma chambers and the development of very large surface craters called calderas (caldera) is an important source of lake basins. Crater Lake, Oregon, is a typical example, exhibiting characteristically great depth and a high encircling rim. Some caldera basins evolved with gently sloping sides, however, due to the deposition of material from a series of explosions and a gentler collapse of the structure. Secondary cones may develop within calderas, as shown by Wizard Island, in Crater Lake. The largest caldera in the world, which contains Lake Toba (Toba, Lake), in Sumatra, was formed through a combination of volcanic action and tectonic activity. Lake Toba's basin is contained in a strike-slip fault belt along the entire length of the Barisan Mountains of Sumatra. A vast, initial eruption of lava under gas pressure collapsed the magma reservoir, forming a depression that filled with water, producing the lake. Renewed volcanic activity subsequently led to the formation of an island in the centre, but a second collapse later cut it in two. Additional tectonic activity has further modified the lake's configuration.

      Lake basins may also arise from the action of lava flows that emanate from volcanic fissures or craters. Lake Mývatyn (Mývatn), in Iceland, was formed in a basin arising from the collapse of the interior part of a large lava flow. Other basins have formed as the result of volcanic damming. This usually happens where a lava flow interrupts the existing drainage pattern.

      Lake basins also may form following the blockage of a drainage depression by landslides. These may be temporary in nature because of the eroding action of the lake on the damming material. Lake Sārez in the Pamirs is stable, being dammed by a rockslide.

Basins formed by glaciation
      The basin-forming mechanism responsible for the most abundant production of lakes, particularly in the Northern Hemisphere, is glaciation. The Pleistocene glaciers, which seem to have affected every continent, were especially effective in North America, Europe, and Asia. The retreat of ice sheets produced basins through mechanical action and through the damming effect of their ice masses at their boundaries.

      In some cases, lakes actually exist in basins made of ice. In other cases, water masses may form within ice masses. Such occurrences are rare and are not very stable. Damming by ice masses is a more common phenomenon but is also likely to be relatively temporary. Glacial moraine (heterogeneous sedimentary deposits at glacier margins) is also responsible for the occurrence of dammed lake basins. The Finger Lakes of New York State are dammed by an endmoraine.

      Ice sheets moving over relatively level surfaces have produced large numbers of small lake basins through scouring in many areas. This type of glacial rock basin contains what are known as ice-scour lakes and is represented in North America by basins in parts of the high Sierras and in west central Canada (e.g., near Great Slave Lake). Tens of thousands of these lakes are found in the ice-scoured regions of the world. Many of them are interconnected with short streams and may contain narrow inlets. Characteristically, they may be dotted with numerous islands and sprawling bays. Many are comparatively shallow. Where they are particularly abundant, they may cover up to 75 percent of the total surface, as in Quetico canoe country of Minnesota.

      Glacier scouring associated with the freezing and thawing of névé (firn) (granular snow adjacent to glacier ice) at the head of a glaciated valley may produce a deepened circular basin termed a cirque. These are found in widely scattered mountain locations. The action of glaciers in valleys can produce a similar type of basin, often occurring in series and resembling a valley staircase. Ice movement from valleys through narrow openings has produced another type of rock basin, known as glint lake basins, particularly in Scandinavian regions.

       piedmont and fjord (i.e., a river valley that has been “drowned” by a rise of sea level) lakes are found in basins formed by glacial action in long mountain valleys. Excellent examples are found in Norway, the English Lake District, the European Alps, and the Andes. In North America, several regions contain this type of lake basin. In British Columbia, many good examples exist, the largest of which are the Okanagan and Kootenay systems. These are long, narrow lakes of substantial depth. In northwestern Canada, some of the largest lakes, including Lake Athabaska (Athabasca, Lake), Great Slave Lake, and Great Bear Lake, are of this type, although they are not found in the same type of mountainous terrain. These lakes, as well as the North American Great Lakes, resulted from the movements of large ice sheets that deepened existing valleys.

      The Wisconsin (Wisconsin Glacial Stage) (latest stage of Pleistocene glaciation) ice sheet was responsible for shaping the present Great Lakes system, which drains mainly eastward to the Atlantic through the St. Lawrence River, during its retreat. The principal stages in the history of these lakes have received much study, and several stages of retreat and advance of the ice sheet have been identified. Behind the lobes of the ice sheet, ice lakes developed that drained according to the modifications of preexisting valleys for glacial action. As the mass of ice retreated far to the north, glacial rebound (uplift of the Earth's crust in response to removal of the loading by ice) caused a general tilting of the land surface; the new lake basins also contributed to the subsequent changes through their own erosional action.

      The material comprising glacial moraines or glacial outwash may provide dams that confine postglacial waters. The Finger Lakes, in New York State, constitute one interesting group of this type. These lakes were formed through glacial scouring of existing valleys, which were blocked at both the northern and southern ends by morainic deposits.

      A variety of basin types have been formed in the different types of glacial drift deposits, including basins in morainic material, kettle lakes, channels formed by water movement in tunnels beneath the ice masses, and lake basins formed by thawing in permafrost. An interesting example of glacial action is the formation of giant's kettles (kettle); these are glacial potholes in the form of deep cylindrical holes. Their origin is still uncertain. Sand, gravel, or boulders are sometimes found at their bottom. The kettles vary from a few centimetres to a metre or more in diameter. Good examples are found in the Alps, Germany, Norway, and in the United States.

Basins formed by fluvial and marine processes
      Fluvial action in several forms can produce lake basins; the most important processes include waterfall action, damming by sediment deposition from a tributary (fluviatile dams), sediment deposition in river deltas, damming by tidal transport of sediments upstream, changes in the configuration of river channels (e.g., oxbow lakes and levee lakes), and solution of subsurface rocks by groundwater.

      This last mechanism has produced the well-known formations in the karst region, in Yugoslavia, which include subterranean and surface cavities and basins in limestone. The term karstic phenomena is applied to similar cases in many parts of the world (see further cave and karst landscape (coastal landforms)). Solution lakes in Florida (e.g., Deep Lake) are also of this origin, as are Lünersee and Seewlisee, in the Alps. Other rock types susceptible to solution basin formation include gypsum and halite. Mansfeldersee in Saxony was formed in this manner.

      In some coastal areas, longshore marine currents may deposit sufficient sediment to block river outflows. This damming action may be of varying intensity, and it may also occur in lake regions, where such current action causes sediment deposition that leads to the formation of multiple lakes. Accumulation of organic plant material can also result in structures that produce lake basins; Silver Lake, Nova Scotia, evolved from damming by plant material. Structural formations of coral are another potential cause of damming.

Basins formed by wind action, animal activity, and meteorites
      Wind action may lead to dam or dune construction or erosion and thus can play a role in lake-basin formation. The latter case has been demonstrated in North America; a number of basins in Texas and northward, on the plains east from the Rocky Mountains, are thought to have originated from wind erosion—at least in part. Moses Lake in Washington state was formed by windblown sand that dammed the basin.

      Mammals have constructed lake-forming dams; the American beaver is highly skilled at this, and its activities in this connection have established it as a symbol of industriousness. Man has also been busy in this regard and is fully capable of producing lakes that would rival the largest of the more natural variety. Plans once proposed for the damming of the Yukon River in Alaska would, if carried through, result in the formation of a lake larger than Lake Erie in surface area. Other human activities, such as quarrying and mining, also have produced cavities suitable for lake formation.

      The last major mechanism of basin formation is that due to meteoritic impact. Meteorite craters (meteorite crater) are best preserved in arid climates and are often dry for this reason. A few lakes are known in craters, however, including Ungava Lake, in Quebec. In many other cases, it has not been possible to definitely confirm that basins that have the appearance of meteorite craters have, indeed, been produced by meteorite impact. Controversial ones include the bay lakes of southeast North America.

Topography of basins
      Lakes meet with both the atmosphere and the underlying material of their terrestrial basins and interact with each. The topography and configuration of the lake bottom and the nature of the bottom materials vary considerably. They are of sufficient importance to most lake processes to warrant recognition as basic lake characteristics.

      The surface area of a lake can easily be determined by cartographic techniques, but lake-volume determinations require knowledge of lake depths. Throughout the world, lakes important enough to warrant study have been sounded, and many nations have completed comprehensive programs to determine the bathymetry of large numbers of lakes. Lake sounding involves traversing a lake to collect either point or continuous measurements of depth until an accurate survey is made. Modern sounding devices measure the time taken for emitted sound to return after reflection from the bottom, relying on a knowledge of the speed of sound in water. The more sophisticated of these also provide for detection of the depths of stratification in sedimentary materials on the lake bottom. The employment of laser devices from aircraft is a recent development that is based on the transmission of light beams with wavelengths that will penetrate water.

      For more practical purposes, lake morphology is a stable characteristic. Shore erosion, sediment deposition and transfer, and other processes, however, including dredging by man, may significantly alter a lake's bottom topography and thus affect navigation, currents, and ecological factors, such as fish spawning grounds.

Sediments and sedimentation
      Lake sediments are comprised mainly of clastic material (sediment of clay, silt, and sand sizes), organic debris, chemical precipitates, or combinations of these. The relative abundance of each depends upon the nature of the local drainage basin, the climate, and the relative age of a lake. The sediments of a lake in a glaciated basin, for example, will first receive coarse clastics, then finer clastics, chemical precipitates, and then increasingly large amounts of biological material, including peats and sedges.

      Geologists can deduce much about a lake's history and the history of the lake basin and climate from the sedimentary records on its bottom. A sediment core contains such clues as ripple marks caused by current or wave action, carbonaceous layers, and alternations of strata that include cold- and warm-water species of fossils, pollen, and traces of chemicals of human derivation. These data provide the basis for extensive documentation of lake history (paleolimnology). Some well-known historical events, such as major volcanic eruptions, the clearing of North American forests by early settlers as revealed by pollen concentrations, the first extensive use of certain heavy metals by industry, and nuclear explosions, provide reference points in the sediment record.

      Many of the materials that are detrimental to the ecology of a lake—e.g., excessive quantities of nutrients, heavy metals, pesticides, oil, and certain bacteria—are deposited in lake sediments by chemical precipitation or the settling of particulate matter. These materials are potentially available for regeneration into the lake water and must be considered in any planning for measures to abate lake pollution. Within the uppermost lake sediments, large volumes of interstitial water are often present. This water may have high concentrations of nutrients and other constituents and enhance the exchange potential with the lake proper.

Clastic sediments
      Waters draining into a lake carry with them much of the suspended sediment that is transported by rivers and streams from the local drainage basin. Current and wave action along the shoreline is responsible for additional erosion and sediment deposition, and some material may be introduced as a result of wind action. Rivers and streams transport material of many different sizes, the largest being rolled along the riverbed (the bed load). When river water enters a lake, its speed diminishes rapidly, bed-load transport ceases, and the suspended load begins to settle to the bottom, the largest sizes first. Lake outlets carry with them only those materials that are too small to have settled out from the inflows or those that have been introduced adjacent to the outflow. Because dynamic processes that keep materials suspended are generally more active near the shore, lake sediments are usually sorted by size; the rocks, pebbles, and coarse sands occur near shore, whereas the finer sands, silts, and muds are, in most cases, found offshore.

      Clastic material over most of a lake basin consists principally of silts (silt) and clays (clay), especially away from shores and river mouths, where larger material is deposited. Clays exist in a variety of colours, black clays containing large concentrations of organic matter or sulfides and whiter clays usually containing high calcium carbonate concentrations. Other colours, including reds and greens, are known to reflect particular chemical and biological influences.

      Organic sediments are derived from plant and animal matter: förna is recognizable plant and animal remains, äfja finely divided remains in colloidal suspension, and gyttja is a deposit formed from äfja that has been oxidized. Rapid accumulation of organic matter in still lakes is not uncommon; in the English Lake District, 5 metres (15 feet) of lake sediment of organic origin accumulated over a period of about 8,000 years. Pollen analysis has been used to accurately decipher climatic conditions of the lake in the past.

      Varved deposits (varved deposit) are the product of an annual cycle of sedimentation; seasonal changes are responsible for the information. Varves are a common feature in many areas and especially so where the land has received meltwaters from ice sheets and glaciers. The deposits consist of alternating layers of fine and coarse sediments.

      Coarse clastic materials seldom are larger than boulders (25 centimetres [10 inches]), and the type of material in sizes larger than silt and clay frequently reveals its source. Materials along lakeshores can in most cases be traced back to a particular eroded source within the local drainage basin, and the distribution of this material provides evidence of the predominant current or wave patterns in the lake.

      Volcanic ash is deposited downwind from its source. Ash from volcanic activity during the Pleistocene Epoch can often be dated and used as a stratigraphic marker. Lakes throughout the northwestern United States contain some of the best examples (the Mazama ash), and one deposit in the central United States, called the Pearlette ash deposit, occurs in beds as thick as 3 metres (10 feet).

Chemical precipitates
      The major chemical precipitates in lake systems are calcium, sodium, and magnesium carbonates and dolomite, gypsum, halite, and sulfate salts. Calcium carbonate is deposited as either calcite or aragonite when a lake becomes saturated with calcium and bicarbonate ions. Photosynthesis can also generate precipitation of calcium carbonate, when plant material takes up carbon dioxide and bicarbonate and raises the pH above about 9 (the pH is a measure of the acidity or alkalinity of water; acid waters have a pH of less than 7, and the pH of alkaline waters range from 7 to 14).

      Dolomite deposition occurs in very alkaline lakes when calcium carbonate and magnesium carbonate combine. Recent dolomites have been found in Lake Balqash (Balkhash, Lake), in Kazakhstan. In many saline lakes, gypsum deposition has occurred; Lake Eyre (Eyre, Lake), Australia, is estimated to contain more than 4,000,000,000 tons of gypsum. For gypsum to be deposited, sulfate, calcium, and hydrogen sulfide must be present in particular concentrations. hydrogen sulfide occurs in deoxygenated portions of lakes, usually following the depletion of oxygen resulting from decomposition of biological material. Bottom-dwelling organisms are usually absent.

      Lakes that contain high concentrations of sodium sulfate are called bitter lakes, and those containing sodium carbonate are called alkali lakes. Soda Lake, California, is estimated to contain nearly 1,000,000 tons of anhydrous sulfate. Magnesium salts of these types are also quite common and can be found in the same sediments as the sodium salts. Other salts of importance occurring in lake sediments include borates, nitrates, and potash. Small quantities of borax are found in various lakes throughout the world. Lakes with high alkalinity levels, such as Mono Lake in California, can still support some forms of life.

      The gradual increase of sediment thickness through time may threaten the very existence of a lake. When a lake becomes shallow enough to support the growth of bottom-attached plants, these may accelerate the extinction of a lake. In several European countries, steps are being taken to restore lakes threatened by choking plant growth. Lake Hornborgasjön, Sweden, long prized as a national wildlife refuge, became the subject of an investigation in 1967. Lake Trummen, also in Sweden, was treated by dredging its upper sediments. In Switzerland, Lake Wiler (Wilersee) was treated by the removal of water just above the sediments during stagnation periods.

Lake waters

Chemical composition
      Although the chemical composition (water) of lakes varies considerably throughout the world, owing to the varying chemistry of the erosion products of different lake basins, in most cases the principal constituents are quite similar. Human influences also have contributed substantially to the chemical composition of lakes, and, although industrial effluents vary somewhat from lake to lake, many of the chemical effects of human activities are similar throughout the world. Another source in the chemical balance of lakes is the dissolved and suspended material contained in precipitation. Again, human activities have been in large part responsible for steadily increasing concentrations of this input.

Salinity, nutrients (nutrient), and oxygen
      Salinity is the total concentration of the ions present in lake water and is usually computed from the sodium, potassium, magnesium, calcium, carbonate, silicate, and halide concentrations. Several important bodies of inland waters, often called inland seas, have very high salinities. Great Salt Lake, in Utah, has a salinity of about 200,000 milligrams per litre, as compared with Lake Superior's value of about 75 and an estimated mean for all rivers of 100 to 150. These ions are steadily introduced to lakes from rivers and rainwater, where they concentrate because of the evaporative loss of relatively pure water.

      Where inflowing rivers erode igneous rocks, lake salinity values are relatively low, but, where soluble salts are available for erosion, salinities are relatively high. In general, it has been found that, of the cations (positively charged ions), calcium concentrations are highest, followed by magnesium, sodium, and potassium, in that order. Among the major anions (negatively charged ions), carbonate is generally the most abundant, followed by sulfate and chloride.

      Other inorganic ions, though present in smaller concentrations, are of great importance. In particular, the nutrients (especially phosphate, nitrate, and silicate), heavy metals (e.g., mercury, manganese, copper, lead), and polychlorinated hydrocarbons (DDT, for example) have attracted recent interest because of their role in ecological problems. Although sources of nutrients and mercury exist that are not directly related to human activities, budget studies and studies of the historical records available in sediment cores clearly reveal the great impact of human disposal of these constituents in lakes. Rainfall (precipitation) and dry fallout are small but significant chemical inputs to lakes. The release of gases and particulate matter into the atmosphere from factories and similar sources has increased dramatically in recent years, with consequent alterations in the chemistry of rainwater. It has been estimated, for example, that 16,000 tons of nitrogen, about 8 percent of the total from all sources, is introduced annually to Lake Erie from atmospheric action.

      The substance of most interest in lakes is oxygen; once introduced to the lake water, its concentration is subject to factors within the water. Biological production (photosynthesis) releases oxygen into the water, while biological decay consumes it. Various chemical reactions within the lake system also affect the concentration of dissolved oxygen. The main source is the passage of oxygen through the air-water interface, which is affected principally by the lake temperatures; at low temperatures the partial pressure of dissolved oxygen in water is reduced. Consequently, during cold seasons, especially when vertical mixing is greatly enhanced because of a lack of thermal structure and increased wind stirring, lakes are replenished with oxygen. In the warmer seasons, although surface waters may remain more or less saturated and even supersaturated, the concentrations are lower. Beneath the surface, oxygen consumption through biological decay may cause serious depletion. Oxygen depletion also occurs near the bottom because of processes at the mud-water interface, many of which are still inadequately explained.

      In winter months, a rapid formation of ice or the establishment of strong winter thermal stratification may significantly inhibit the replenishment of oxygen. Where ice cover lasts for long periods, a loss of oxygen at the mud-water interface may have repercussions for the whole lake, particularly if density currents cause significant vertical transport.

      In tropical regions, where the winter replenishment mechanism (turnover) is absent, there is great reliance on the occasional occurrence of cold spells or on significant nighttime cooling to promote oxygen replenishment. Deep lakes in these regions are often anoxic (lacking in oxygen) in the deeper portions.

      At any particular time, lake waters or waters entering a lake may have a biological or chemical potential for oxygen utilization. Measurements of this are termed BOD (biological oxygen demand) or COD (chemical oxygen demand). These concepts are used as partial indicators of the quality of waters being introduced to a lake.

      Lakes that have a vertical salinity gradient strong enough to prevent winter turnover will usually be deoxygenated at depths where the vertical diffusion of oxygen is less than the oxygen demand. Such lakes are termed meromictic.

      Another gaseous substance of great importance that is exchanged with the atmosphere at the surface is carbon dioxide. Photosynthesis requires the presence of carbon dioxide, and it is released during biological breakdown.

      Carbon dioxide is very soluble in lake water; it forms carbonic acid, which dissociates and raises the concentration of hydrogen ions (lowering the pH). The relative proportions of bicarbonate, carbonate, and free carbon dioxide depend upon the pH. At high values of pH, carbonate ions will predominate; at low values, free carbon dioxide and carbonic acid will predominate.

      Various carbonates (particularly sodium, calcium, potassium, and magnesium) are important to the carbon dioxide system. Increased pressure of carbon dioxide in the system increases the solubilities of these carbonates. In some cases, photosynthetic activity results in precipitation of certain carbonates. The entire carbon dioxide system and its behaviour at various pH values is very complex but can be interpreted from historical knowledge of lake sediments.

      In waters that are neither very acidic (pH much less than 7) nor very basic (pH much greater than 7 but less than 14), the carbon dioxide system serves as a buffer, because, within limits, a change in pH will cause a shift within the system that ultimately serves to offset the pH change. Consequently, most lakes have a pH between 6 and 8. Some volcanic lakes are extremely acid, however, with pH values below 4, and some lakes with very high pH values, such as Lake Nakuru (Nakuru, Lake), Kenya, also occur in nature.

Sulfates, nitrates, and phosphates
       sulfate usually occurs as a principal ion in lake waters. Under anaerobic conditions in which bacteria persist in the oxidation of biological material, hydrogen sulfide is produced. When anoxic conditions exist in the deep waters just above the sediments, and the water is acidic enough to precipitate the iron present, hydrogen sulfide occurs. The characteristic and unpleasant odour of this gas is often popularly identified with the “death” of a lake. Big Soda Lake, Nevada, is extremely rich in this substance.

       nitrogen and its various compounds form another complex system in lakes, appearing as free nitrogen in solution, organic compounds, ammonia, nitrite, and nitrate. Sources of nitrogen compounds include influents to the lake (the most important source), fixation in the lake, and precipitation. Losses are experienced mainly through effluents but also by denitrification, sediment formation, and loss to the atmosphere.

      Orthophosphate and various organic phosphates (phosphate mineral) are the most important phosphorous compounds in lakes. Phosphates and nitrates are heavily consumed in the upper portion of lakes during periods of high productivity of phytoplankton. Increased concentrations occur in deeper portions due to decay of falling biological material and regeneration from the sediments, especially during anoxic conditions or stormy periods in shallow lakes. As limiting nutrients in many lake productivity cycles, phosphates and nitrates are often identified as controllable elements in situations where abatement is necessary to control eutrophication. carbon is also a necessary constituent for production and in some cases can be the limiting component. Because carbon is less easily controlled and not often limiting, however, phosphates are most frequently named as substances to be reduced in effluents from industry and municipalities.

       silica also is present in lake waters, and, as with the other nutrients, it is introduced in influents and to some extent from the sediments. The production of diatom blooms is a major process for reducing silicate concentrations. Within this context, silica can also be regarded as a limiting nutrient.

Thermal properties
      Pure water freezes at 0° C (32° F), boils at 100° C (212° F), and has a latent heat of evaporation of 539.55 calories per gram, a latent heat of sublimation (ice) of 679 calories per gram, and a specific heat of 1.01 calories per gram, per ° C, at 0° C. The temperature of maximum density of water at atmospheric pressure occurs at 3.94° C (39.09° F). At the freezing point, ice has a lower density than water. For natural waters with high salinities, such as the oceans and inland seas, each of the values above is significantly altered. In most lakes, however, these numbers are quite representative.

      The density of water increases at pressures above one atmosphere (the pressure at sea level). Thus, pure water at 10°C (50° F) has a density of 0.9995 at one atmosphere and 1.0037 at the pressure existing at a lake depth of 1,000 metres (3,000 feet). Water raised from great depths to conditions of lower pressure experiences adiabatic cooling (without significant heat exchange with surrounding water), but there are very few lakes in which this factor can be of much significance.

Vertical mixing and overturn
      It is useful to know how the temperature of maximum density changes with depth (e.g., from 3.94° C at the surface to 3.39° C at 500 metres depth [38.10° F at 1,500 feet]). The fact that the temperature of maximum density of most lake waters is close to 4° C (39° F), whereas ice forms at temperatures close to 0° C in response to surface cooling, vertical mixing takes place. When density increases with depth, the lake is said to be stable. Unstable conditions exist when density decreases with depth. Cooling at the surface to temperatures below 4° C establishes stability based on a negative thermocline (a positive thermocline is a vertical decrease in temperature with depth), because density will increase with depth. Ice then forms at the surface, enabling liquid water to exist beneath the ice in lakes, unless they are shallow enough to freeze to the bottom.

      During the warming season, after ice has melted, heating increases the density of the surface waters, causing them to sink until stability is achieved. When surface heating proceeds above the temperature of maximum density, this process ceases, and the vertical thermal structure maintains and strengthens its stable condition, based on a positive thermocline. Turnovers tend to be seasonal.

      Mixing due to cooling or warming processes that increase the density (density current) of surface waters sufficiently to cause them to sink results in what is termed circulation, or overturn, of lake waters. Lakes that cool to below 4° C in winter experience two turnover periods, as just described, and are called dimictic lakes. Most lakes in temperate regions fall into this category. Lakes that do not cool to below 4° C undergo overturn only once per year and are called warm monomictic. Lakes that do not warm to above 4° C also experience only one overturn period per year and are called cold monomictic. There are many examples of the former, including lakes in the tropical regions and generally as far north as about 40°. The cold monomictic type, however, is less common but can be found at high latitudes and at high altitudes (in the Alps, for example).

      All the types described that circulate at least once throughout are called holomictic. It is possible, however, for lakes to be stable despite the thermal processes that normally induce overturn owing to the existence of a positive salinity gradient with depth (chemocline). This type is called meromictic, and, in those cases where stability is permanent in at least part of the lake, the deep waters do not experience overturn and consequently are deoxygenated. Three principal origins of meromixis have been recognized. Ectogenic meromixis results from either the intrusion of seawater into a lake, as in the case of flooding from an unusually high sea level (e.g., Hemmelsdorfersee, in Germany), or the introduction of fresh water through land drainage and precipitation to a saline lake (e.g., Soda Lake, Nevada). Crenogenic meromixis is due to the introduction of saline water by springs, and biogenic meromixis is due to the uptake of salts from the lake sediments. North American examples include Lake Mary, Wisconsin, and Sodon Lake, Michigan.

      A strong vertical salinity gradient that exists in the upper portion of a lake will affect the thermal structure by inhibiting the downward mixing of heat. In holomictic lakes, however, the downward mixing of heat due primarily to wind action usually compresses or concentrates the thermocline until it essentially separates an upper layer (epilimnion) from a lower layer (hypolimnion), each possessing weak or nonexistent vertical thermal gradients. The thermocline normally begins to grow at the beginning of the warming season. As summer passes and autumn commences, it intensifies and deepens. The onset of the cooling sees the beginning of the decay of the thermocline from above, although it usually continues to deepen until it is completely destroyed. The process just described is commonly found in lakes in temperate regions and is a seasonal phenomenon. During any period of strong warming, one or more shallower thermoclines may be observed to develop and move downward to the seasonal thermocline.

The heat budget of lakes
      The heat budget of a lake includes several major factors: net incoming solar radiation, net exchange of long-wave radiation emitted by the lake surface and the atmosphere, transfer of sensible heat at the surface interface, and latent-heat processes. Those processes that are usually of much smaller importance include net inflow and outflow of heat advected by streamflow, precipitation, and groundwater flow, conduction from terrestrial heat flow, and dissipation of kinetic energy. In some cases, however, river inflow may be of more importance, such as where flow is from a nearby glacier or where the volume inflow is a significant fraction of the lake volume. Within a large lake the heat-budget considerations for a particular location must also take into account the local advection of heat within the lake by currents.

      Incoming solar energy varies seasonally and with the latitude and is greatly influenced by cloud cover. The fraction that is reflected away from the lake surface depends upon the solar angle, the turbidity of the atmosphere, and the wave state, or surface roughness. In middle latitudes this ranges from about 6 percent in summer months to about 14 percent in winter.

      The amount of radiation emitted by the lake surface is proportional to the fourth power of the surface temperature, whereas the radiation emitted by the clouds and atmosphere overlying the lake depends primarily upon the amount and height of the clouds and the temperature and moisture content of the atmosphere near the lake surface.

      The fluxes of sensible heat and moisture at the lake surface are of great importance yet are still poorly understood. They depend upon the vertical gradients of temperature and vapour pressure above the water, respectively, and upon the factors that influence the transfer processes, such as wind and atmospheric stability. The transfer of sensible heat may be either into or out of the lake surface, usually on a seasonal basis but also sometimes on a diurnal basis. It is also possible but less likely for condensation to occur on a lake surface.

      Heat flow through the bottom of lakes is normally of small significance, but exceptions exist. In a very deep lake where low rates of heating are important, such as Lake Baikal, Russia, the results may be detectable. In some ice-covered lakes where other sources of heating are small, heat flux through sediments also has been shown to be significant.

      The dissipation of wind energy that has been transferred to water movements is quite insignificant, as is the effect of heat transfer due to chemical and photosynthetic processes.

      In latitudes and altitudes where ice is a factor, the latent heats of fusion and of evaporation of ice must also be considered within any heat-budget considerations. Heat-balance studies have been performed for lakes that are always ice covered. Solar radiation is often an important factor where ice thickness and consistency permits penetration. The heat balance of the ice is often difficult to assess, as long-wave radiation and evaporation factors are not easily measured and are very important. The exchange of sensible heat may not be large during summer months in these cases but is likely to be significant in the colder months. Several lakes that are ice covered have been shown to be meromictic; two examples are Lake Tuborg, Ellesmere Island and Lake Bonney, Antarctica.

      Heat-balance measurements or estimates have been made for many lakes throughout the world. Results show that the difference between the highest and lowest heat content for each lake varies from around 5,000 calories per square centimetre for high and low latitudes to around 45,000 calories for some midlatitude lakes.

      The relative importance of each of the major terms of the heat budget is shown by data for two North American lakes: Lake Ontario (Ontario, Lake), a large, deep, middle-latitude lake; and Lake Hefner, a relatively small, shallow lake in Oklahoma. The energy unit frequently employed is the langley (one gram calorie per square centimetre), and the figures given are approximate monthly means of langleys per day. Net solar radiation input to Lake Ontario varies from 80 to 600 (Lake Hefner varies from 200 to 600), midwinter to midsummer. Net losses due to long-wave radiation from Lake Ontario are nearly 100 throughout the year (Lake Hefner varies from 100 to 200). Evaporation losses for Lake Ontario vary from 250 in midwinter to slightly negative values in early summer (Lake Hefner varies from 450 in late summer to 150 in spring). Conduction of heat from the surface of Lake Ontario varies from 250 in winter to about minus 100 in summer (Lake Hefner varies considerably from 80 to -80 for the same time interval).

      Heat added to a lake at the surface is usually mixed mechanically downward as a result of wind action. This process keeps the upper portion of a lake relatively uniform thermally. Consequently, a thermal gradient (thermocline) becomes established between the upper mixed layer (epilimnion) and the deep portion of the lake (hypolimnion). In shallow lakes or shallow portions of large lakes, the thermocline will eventually intercept the lake bottom so that no hypolimnion exists. Normally, as the heating season progresses, the thermocline intensifies and deepens. Secondary thermoclines may develop in the epilimnion, and these will migrate downward to the main seasonal thermocline. On very warm, still days, a thin surface layer may store heat before a mixing episode transfers heat downward. When the cooling season commences, the mixing that tends to destroy the thermocline is enhanced by vertical convection. If the cooling continues until the entire thermocline is eliminated, the lake becomes essentially isothermal and no longer exhibits the characteristics of a two-layered system.

      When a lake is stratified, the most important process for downward transfer of heat to the hypolimnion is through eddy conduction. The coefficient of eddy conductivity is determined empirically and varies substantially from lake to lake. Mixing processes are generally more active in coastal areas, so that isotherms can be expected to slope downward toward shore. In large, relatively unprotected lakes, wind stress at the surface causes convergence or divergence or both of shallow waters along coastlines. Isotherms will slant upward toward the shore, and hypolimnion water may even become exposed at the surface. These occurrences are of great importance with regard to the distribution of heat within stratified lakes.

      Heat introduced to lakes in large quantities, as a waste product of cooling processes in power-generating plants and other industrial concerns, is presently viewed with some concern as a pollutant, especially in small lakes. If the heat is injected at the surface it will spread initially according to the momentum of the influent and the speed and direction of ambient surface currents. When the initial momentum is sufficiently dissipated, the heat will spread mainly as a consequence of turbulent mixing processes. Throughout these events, substantial losses of heat to the atmosphere may occur, so that the full effects of the thermal input are not borne solely by the lake. Temperature values at the surface, adjacent to the influent-heat source, may be raised to a very high level—as much as several centigrade degrees. Under certain conditions fish-activity tolerances may be exceeded, and undesirable algae and plankton production may be stimulated.

      If waste heat is not released at the surface but is diffused over a large depth range or injected at depth, the large local-surface-temperature problem is avoided. Losses to the atmosphere in this case, however, are also greatly reduced, and the net heat input to the lake as a whole is much greater. Over a long period, this may prove to be more detrimental to the general ecology than near-surface injection.

Lake hydraulics

Currents
      The principal forces acting to initiate water movements in lakes are those due to hydraulic gradients, wind stress, and factors that cause horizontal or vertical density gradients. Lake water movement is usually classified as being turbulent.

      Hydraulic effects are frequently the result of inflows and outflows of water. These may be substantial and continuous or weak and sporadic; in terms of the ratio of the volume of the inflow or outflow to the lake volume, the latter is the most frequently observed situation.

      The stress of wind moving over the lake surface causes a transport of water within the lake, as well as the movement of energy downwind through the mechanism of surface waves. The wind is therefore one of the most important external forces on a lake. It can be relatively consistent in speed and direction, or it can be highly variable in either or both.

Pressure gradients
      Water movements can occur as a result of internal pressure gradients and from density gradients caused by variations in temperature, sediment concentration, or the concentration of dissolved substances. Surface water in lakes can become denser than underlying water either by cooling or heating, because the temperature of maximum density for pure lake water is about 4° C (39° F). Water entering a lake from rivers with a high concentration of dissolved substances will sink to a lake level of similar density. These movements are both horizontal and vertical, but the net effect is downward, if not vertical, motion.

      Horizontal pressure gradients can result from many different processes that act to produce density gradients. One example is the situation of solar heating in a shallow nearshore region, where the heat is committed to the warming of a relatively small volume of water. This produces a water of lower density than the near-surface water of an adjacent deep region, where the heat is spread throughout a greater volume. Consequently, the pressure gradient force will act to move the warmer water offshore and to replace it from below with cooler water.

      Lake currents are the result of complex interactions of forces, but in many cases a small number of particular forces dominate. In the case of horizontal flow in the absence of horizontal pressure gradients, assuming no friction, water set in motion will curve to the right in the Northern Hemisphere because the Earth rotates from west to east. This effect is called the Coriolis force, and it will continue to influence water motion until there is a balance with the centrifugal force. This movement causes free-floating markers to move in an elliptical manner with a period that depends upon the latitude. In Lake Ontario, for example, it is about 17 hours. Where a dominating pressure gradient exists, the balance of the pressure-gradient force with the Coriolis force results in the so-called geostrophic flow, at right angles to the pressure gradient, with low pressure on the left (Northern Hemisphere). These conditions are most nearly realized only in very large lakes and in the oceans.

      In those small lakes where hydraulic effects dominate, steady flow conditions may be achieved through balance with friction. This situation is commonly encountered in rivers, and relationships exist between mean current speed and the slope and mean depth of the river or narrow lake. These are called gradient currents and occur following situations where the wind or atmosphere pressure gradient causes a tilting of the lake surface (denivellation). In cases where the Coriolis force is a significant factor, the flow down a lake will tend to move toward the right (in the Northern Hemisphere). The development of a deeper countercurrent to the left will occur to compensate for the piling up of water on the right side.

      Horizontal pressure gradients will be important in lakes where there are significant inflows of water with markedly different density from ambient lake density or where significant differential surface heating occurs.

Wind stress
      Currents resulting from wind stress are the most common in lakes. Considerable research is still underway into the mechanism of transfer of wind momentum to water momentum. The stress on the lake is proportional to some power of wind speed, usually taken to be 2, although it evidently varies with wind speed, wave conditions, and atmospheric stability. In large, deep lakes, away from the boundaries, where wind-stress effects may be balanced by Coriolis-force effects, theory suggests that the surface current will move in a direction 45° to the right of the wind and that deeper currents are progressively weaker and directed farther to the right. The depth at which flow is opposite to the wind direction is effectively the depth below which there is no influence from the wind. This depth, designated D, can theoretically occur at about 100 metres (300 feet) in large, deep, midlatitude lakes. Observations show varying degrees of fidelity to theory because of complications from coastal effects and thermal stratification.

      In coastal regions, if water depth is a significant fraction of or greater than D, winds blowing parallel to the shore will transport water either onshore or offshore. In the latter case, where the coast is to the left of the wind flow (Northern Hemisphere), the water driven offshore is replaced by cooler, deeper water (upwelling).

Internal waves and Langmuir circulation
      Under stratified conditions a strong thermocline will essentially separate a lake into two layers. Shearing forces that develop between these layers cause a motion, termed internal waves, that may serve to directly dissipate a substantial portion of a lake's kinetic energy and act as a coupling between motion in the epilimnion and hypolimnion. A great range of periodicities is observed in the oscillations of the thermoclines, particularly in large lakes. Internal seiches, which are responsible for relatively long-period internal waves, are discussed later.

      A small-scale circulation phenomenon that has aroused considerable attention on lakes is Langmuir circulation. On windy days, parallel “streaks” can be observed to develop on the water surface and exhibit continuity for some distance. These streaks may be caused by convergence zones where surface froth and debris collect. Langmuir circulation thus appears to be a relatively organized mixing mechanism wherein sinking occurs at the streaks and upwelling occurs between the streaks. Under favourable circumstances, this appears to be a key process for mixing heat downward in lakes.

Surface waves
      Wind blowing over a calm lake surface first produces an effect that may appear as a widely varying and fluctuating ruffling of the surface. The first wave motion to develop is relatively regular, consisting of small, uniformly developed waves called capillary waves. These are quite transient, dissipating rapidly if the wind dies away or developing to the more commonly observed and more persistent gravity waves.

      Energy will be continually fed to the waves by the frictional drag of the air moving over the water and by the direct force of the wind on the upwind face of the waves. The latter effect occurs only while the waves move more slowly than the wind. Pressure differences at the air–water interface also contribute energy to surface waves. Energy losses occur due mainly to turbulence in the water and, to a smaller extent, to the effects of viscosity.

      Waves will continue to grow as long as there is a net addition of energy to them. Their height will increase as a function of wind speed and duration and the distance over which it blows ( fetch). Most lakes are so small that fetch considerations are unimportant. Studies in larger lakes, however, have shown that the height of the highest waves are related to the fetch. In these lakes, waves as high as several metres are common, although waves of about seven metres (23 feet) are the highest to be expected. Wave heights in a given portion of a lake may vary considerably, due to interactions that suppress some waves and amplify others. As waves develop, their lengths increase, even after their height has stopped increasing. The phenomenon of swell, commonly observed in the oceans, is not truly realized, even in the largest lakes.

      Waves travel in the same direction as the wind that generated them and at right angles to their crests. If they meet a solid object rather than a sloping beach, much of their energy will be reflected. If they enter shallow water obliquely, they are refracted. Wave speed, for waves longer than four times the depth of the water, is approximately equal to the square root of the product of the depth and the gravitational acceleration. For waves in relatively deep water, the wave speed is proportional to the square root of the wavelength.

      As wave height increases, the sharpening of the wave crest may result in instability and a breaking off of the crest, a process hastened by the wind. This results in the familiar whitecaps. Waves that run ashore break up in surf. The wave height first decreases slightly, then increases, and the speed decreases, and eventually the wave form disappears as it crumbles into breakers. These can be plunging forms, in which the top curls right over the forward face, or of the spilling type, in which the crest spills down the forward face. A particular wave may break several times before reaching shore.

Seiches (seiche)
Cause and characteristics
      If a denivellation, or tilting of a lake's surface, occurs as a result of a persistent wind stress or atmospheric pressure gradient, the cessation of the external forcing mechanism will result in a flow of water to restore the lake level. The flow would be periodic and uniform with depth, except for the damping effects of the lake-bottom friction and internal turbulence. Because of this, each successive tilt of the lake surface in the opposite direction occurs at a level slightly less than the previous one. The oscillation proceeds, moving the water back and forth until damping levels the water or until wind and pressure effect another tilt. This process is seiching; the lake oscillation is a seiche. The basic seiche has a single node, but harmonics of the oscillation occur, with several nodes being possible.

      The period of the uninodal seiche can be estimated from a formula that equates it to twice the length in the direction of the tilt, divided by the square root of the product of the mean lake depth and the gravitational acceleration.

      Seiches have been noted, recorded, and studied for hundreds of years. Lake Geneva (Geneva, Lake), Switzerland, was one of the first lakes to be studied in connection with seiching; it has an observed uninodal period of about 74 minutes and a binodal period of about 35 minutes. The observed uninodal periods of Loch Treig and Loch Earn, Scotland; Lago di Garda, Italy (Garda, Lake); Lake Vetter (Vätter, Lake), Sweden; and Lake Erie (Erie, Lake), North America, are approximately nine, 14.5, 43, 179, and 880 minutes, respectively.

      Long, relatively narrow lakes that are exposed to a predominance of wind flow along their major axes are most likely to exhibit so-called longitudinal seiches. Transverse seiching can occur across the narrower dimension of a lake; that observed in Lake Geneva, for example, has a period of about 10 minutes.

      The height of the denivellation depends upon the strength and duration of the forcing mechanism, as well as on the lake size and dimensions. In small lakes, level changes of a few centimetres are common, whereas, in the Great Lakes, intense storms can produce changes as great as two metres (seven feet). If the disturbance causing the tilting moves across the lake at close to the speed of the shallow-water wave speed, a profound amplification can occur, with possible disastrous consequences.

      True tides that result from the gravitational effects of the Moon and Sun are rarely measurable in lakes, but small values of tidal components occasionally have been discerned.

Internal seiches
      Internal seiching results from thermal stratification. The layers separated by the thermoclines oscillate relative to one another. Observed uninodal periods for Loch Earn, Lake Geneva, Lake Baikal, and Lake Cayuga (New York) are approximately 16, 96, 900 (binodal), and 65 hours, respectively.

      Because hypolimnion water is very different from epilimnion water with regard to both thermal and biological characteristics, the massive movements of water and the turbulent exchanges that can occur during internal seiching are very important. Substantial portions of the bottom of shallow lakes can experience periodic alternation of exposure to hypolimnetic and epilimnetic water, and hypolimnetic water can be periodically exposed to the surface.

Effects of wave and current action
Shore erosion and coastal (coast) features
      In a lake's early stages of existence, its shore is most susceptible to changes from wave and current action. As these changes occur, there is a tendency over time to an equilibrium condition—a balance between form and processes that depends upon the nature of the materials present (e.g., the size of sand and gravel present). The effectiveness of waves in the erosion process depends in part upon the depth and slope of the lake bottom. Where the shore consists of a sheer cliff adjacent to deep water, wave energy will be reflected away without much erosional effect. The refraction of waves in zones of irregular coastline tends to concentrate wave energy at some locations and dilute it in others. Thus, features extended out into the lake will receive more wave energy, and the tendency is to smooth out an irregular coastline. Other net effects of shore erosion are an increase in the surface area of a lake and a reduction in its mean depth.

      As erosion takes place, the distribution of erosion products results in transport of finer material offshore. The resulting terrace is called the beach in its above-water manifestation and the littoral shelf where it is below water. Landward, beyond the beach, a wave-cut cliff is usually found. The steeper slope that often separates the littoral shelf from the benthos (bottom) zone in the central part of the lake is called the step-off by some limnologists.

      Water movement directed at an angle to the coastline will result in the generation of currents along the shore. Erosion products will then be transported down the coast and may be deposited in locations where transport energy is dissipated due to movement around a bend or past an obstruction. A buildup of such material is called a spit. If a bay becomes completely enclosed in this way, the spit is called a bar.

      Water in very shallow lakes that are subjected to strong winds may be piled against the lee shore to such an extent that countercurrents will develop from along the lee shore around each side of the lake. The cutting effects of these currents are known as end-current erosion and may characteristically alter the shape of a lake frequently subjected to winds from a particular direction.

Bottom morphology
      The bottom morphology of a lake can be greatly influenced by deposition of sediment carried by inflowing rivers and streams. Although this process can be modified by wave and current action, most lakes are sufficiently quiet to permit the formation of substantial deltas (delta). In very old lake basins the relief may become so extensively decreased due to the great buildup of deltaic deposits and the long-term effects of river widening, that deposition on the outer portions of a delta will fail to balance the effects of wave erosion. A delta, in these circumstances, will begin to shrink in size (see further river: Deltas (river)).

      It is very important to understand lake processes that affect the basin morphology and to be able to predict their trends and their impact on human activities. Increasingly, man is imposing his ability to change natural events in lakes, and he has often encountered problems by not anticipating a lake's reaction to his projects. The actual creation of a lake by damming a river is a major undertaking of this type. One fairly recent example is Lake Diefenbaker, in Saskatchewan. In this region of prairie farmland, the banks of the new lake are extremely vulnerable to erosion, and planners have had to contend with the consequences of bank cutting and infilling of the basin. There are many examples of lesser engineering undertakings that have had to face the consequences of a lake's reaction. The building of jetties or breakwaters, for example, may interfere with natural circulation features. In some cases this has resulted in the reduction of flow past a harbour and increase in flow past a previously stable shoreline, with the result that the harbour has filled in or been blocked by sediment deposition, while the stable shoreline has become badly eroded.

The hydrologic balance of the lakes

The water budget
      The role of lakes within the global hydrologic cycle has been described earlier. Lakes depend for their very existence upon a balance between their many sources of water and the losses that they experience. This so-called water budget of lakes is important enough to have warranted considerable study throughout the world, with each lake or lake system possessing its own hydrologic idiosyncrasies. Aside from being of scientific interest, water-budget studies serve to reveal the dependence of each lake on particular hydrologic factors, thus enabling better management practices. These may include restrictions on water utilization during drought conditions, dike construction and evacuations prior to flooding, control of water levels to ensure efficient power production, and major decisions associated with diversions of watercourses in order to enhance water-quantity- and water-quality-management activities.

      While people may accommodate to predicted imbalances in the hydrologic budget, it is usually difficult to influence the basic natural factors that cause the imbalances. Precipitation and evaporation, for the most part, are uncontrollable, although some advances have been made in evaporation suppression from small lakes through the use of monomolecular surface films. Groundwater flow is not controllable, except where highly restricted flow can be tapped. Rivers and streams, however, can be subjected to regulation by well-established practices through the use of dams, storage reservoirs, and diversions. It is mainly through these controls that efforts are made to make the most efficient usage of water as a resource.

      When engineers take steps to alter elements of a basin's water budget, careful consideration must be given to the consequences of the hydrology and ecology of the entire watershed. Dredging operations for the purpose of harbour clearance or improvements to a navigable channel, for example, may increase the outflow from an upstream lake, increase shore erosion, or regenerate undesirable sedimentary constituents into the lake or river water. The damming of a river or a lake outlet to increase local water storage may also result in undesirable effects, such as an increased evaporation from the larger surface area, the restriction of fish movement, or changes in the thermal climate of the downstream flow. Diversions and dam-site construction may also result in flooding of important bird-breeding areas or a lowering of other lakes in the system, resulting in undesirable consequences.

Water input
      The usual major input of water to a lake derives from streams and rivers, precipitation, and groundwater. In some cases inflow may come directly from glacier melt. The relative importance of each of the major sources varies from lake to lake.

      Stream and river flow are usually seasonally variable, depending upon precipitation cycles and snowmelt. At low altitudes some rivers exhibit a peak during a high precipitation period in winter and then a second peak associated with a subsequent spring snowmelt that feeds the nearby high-altitude tributaries. In regions where precipitation can occur in great quantities at high rates, streams swell quickly and water is delivered in relatively large volumes to downstream lakes.

      A great deal of work has been done to improve the ability to measure and record streamflow. Consequently, it is usually the most accurately known of the inflow terms in the water budget. Most frequently, the height of the river level (stage) correlates well with the water discharge. In other cases, direct river-flow measurements are taken periodically with flow meters.

       precipitation reaching a lake's surface directly may be the major input; this is true of Lake Victoria, in eastern Africa. In other cases, where the lake basin is large with well-developed drainage to a deep lake of small surface area, precipitation may be a small component. Precipitation that falls elsewhere in the lake basin may reach the lake through either surface or groundwater flow, or it may be lost due to evapotranspiration.

      Measurements or estimates of precipitation for a basin are difficult to achieve. Even where elaborate networks of rain gauges exist or where these are supplemented by meteorological radar installations, total basin-precipitation data are still considered to be poor. Measurements of direct precipitation over lakes are exceedingly rare; this situation is especially serious in the case of a large lake for which nearby land data are not necessarily representative of conditions over the lake. Each climatic region throughout the world has its typical precipitation pattern, and the lakes within the regions are affected accordingly.

       groundwater reaches lakes either through general seepage or through fissures (springs). Groundwater is taken to be water in that zone of saturation that has as its surface the water table. The depth of the water table (aquifer) can be determined by digging a well into the saturated zone and noting the level of water—unless the water is under pressure, in which case it will rise in the well to a level above the water table. Clearly, it is possible for a lake level to coincide with the water table. In fact, unless impermeable material intervenes, the water table will drop to, rise to, or lie level with a lake surface. Groundwater that is lost from the saturated zone to a lake is termed groundwater discharge. Groundwater introduced to the saturated zone from a lake is termed recharge. The rate at which groundwater is exchanged between a lake and the saturated zone depends mainly upon the level of the water table and the pressure conditions within the saturated zone.

      In permeable materials the zone above the water table is called the zone of aeration, and water within it is called soil moisture. Soil moisture is classified into three types: hygroscopic water adsorbed on the surface of soil particles, water held by surface tension in capillary spaces in the soil and moving in response to capillary forces, and water that drains through the soil under gravitational influence. The latter will most significantly contribute to groundwater recharge and to the water balance of a lake. The second category will generally be subject to loss due to transpiration by plants.

Water output
      Lakes that have no outlets, either above or below surface, are termed closed lakes, whereas those from which water is lost through surface or groundwater flows are called open lakes. Closed lakes, therefore, lose water only through evaporation. In these cases, the loss of water that is less saline than the source water results in an increasing lake salinity.

      Evaporation (vaporization) results from a vertical gradient of vapour pressure over the water surface. Next to the water surface, saturation conditions exist that are a function of the temperature at the interface. The vapour pressure in the air above the surface is calculated from the temperature of the air and the wet-bulb temperature. The rate at which evaporation occurs also depends upon the factors that affect the removal of the saturated air above the surface, such as wind speed and thermal convection.

      Studies of evaporation must surely constitute a sizable proportion of all hydrological and oceanographic work. The principal categories of evaporation studies are water budget, energy budget, bulk aerodynamic techniques, and direct measurements of vapour flux (see further hydrosphere).

      The so-called aerodynamic technique is based upon Dalton's formula, which correlates evaporation with the product of the vapour pressure gradient and the wind speed. Studies during the past 20 years have produced a host of variations of this equation, determined empirically using independent measurements of evaporation. One of the most often used of these was developed in a study of Lake Hefner, and even this work has been subsequently modified to suit other climates and conditions. Few workers are satisfied with the present state of the art in the use of the aerodynamic equations. Nevertheless, once an equation of this type is satisfactorily developed for a particular lake, having been checked with independent methods, it is attractive because it usually employs data that can be routinely observed.

      The direct measurement of vapour fluxes is an extremely intricate proposition, as motions over a water surface are usually turbulent, and instruments capable of measuring rapidly changing vertical motions and humidities are required. Not the least of the difficulties is the likelihood that the kind of turbulence over large bodies far from land is significantly different from that over land. Recent advances in theoretical developments and instrumentation continue to encourage this type of study. In turn, successes in this field offer the opportunity for the refinement of empirical techniques more practically suited for general lake investigators.

      In many lake studies, data from evaporation pans have been used to determine lake evaporation. Pans have even been developed for flotation on lakes. Pans cannot truly simulate lakes, however, as they constitute a different type of system (they are not exposed to the atmosphere in the same way, they exchange heat through their sides, and they do not store heat in the same way as lakes).

      Some examples of evaporation estimates include annual totals of between 60 and 90 centimetres (two and three feet) for Lake Ontario (Ontario, Lake) (using different techniques and for different years); about 75 centimetres (2.5 feet) for Lake Mendota, Wisconsin; over 210 centimetres (seven feet) for Lake Mead (Mead, Lake), Arizona and Nevada; about 140 centimetres (4.5 feet) for Lake Hefner; about 660 millimetres (26 inches) for the IJsselmeer, in The Netherlands; and about 109 millimetres (4.25 inches) for Lake Baikal.

      Water output from a lake in the form of surface-water outflow generally depends upon the lake level and the capacity of the effluent channel. Although lakes often have many surface inflows or at least several incoming streams or rivers, they generally have but one surface effluent.

Water-level fluctuations
      The net water balance for a particular lake will vary according to the periodic and nonperiodic variations of the inputs and outputs and is reflected in the fluctuations of the lake level. Because the prime influencing factors are meteorological, the periodicity of seasonal events are often seen in water-level records.

      Lake-level rises generally coincide with or closely follow seasons of high precipitation, and falls of level generally coincide with seasons of high evaporation. Complications are introduced by a variety of factors, however. The storage of heavy winter precipitation as snowpack is one example. The release of this water during the spring thaw may also be hampered by the presence of river ice, resulting in late-spring or summer peaks. In large drainage basins the full effects of heavy precipitation may not be immediately realized in the lake-water balance because of the time required for basin drainage. Where glacier melt is a major input to a lake, the changes in level respond to seasonal heating as well as seasonal precipitation.

      Although artificial controls, in the form of diversions, river dredging, and dams, affect the levels of the Great Lakes, the latter provide good examples of seasonal variations because of the lengthy record of levels available. The rivers draining to these large lakes are relatively stable; that is, the ratio of maximum to minimum flow is about 2 or 3 to 1, compared to 30 to 1 for the Mississippi River and 35 to 1 for the Columbia River. A 67-year average of lake levels by month shows that high water occurs, on the average, in September for Lake Superior (Superior, Lake) and in June for Lake Ontario. Lows occur in March and December–January, respectively. The mean range in seasonal levels, for this period, is about 30 centimetres (1 foot) for Lake Superior and about 45 centimetres (1.5 feet) for Lake Ontario. The pattern varies considerably from year to year, however, and periods of exceptional precipitation and drought are shown in the records. These events ultimately affect the downstream lakes, but, because of their relatively small discharge volumes, it takes 3.5 years for 60 percent of the full effect of a supply change to Lake Huron–Michigan to appear in the outflow from Lake Ontario.

      The seasonal changes in a lake's level may be superimposed on longer-term trends, which in some cases dominate. Several of the large lakes of the world have water-level records that illustrate long-term periods of relative abundance of water and drought. In Central Africa, Lakes Victoria (Victoria, Lake), Albert (Albert, Lake), Tanganyika (Tanganyika, Lake), and Nyasa (Nyasa, Lake) exhibit substantial long-term features, some of which are consistent, suggesting that a common climatological factor is responsible. Nevertheless, others of these features are not consistent within the lakes and have not been adequately explained.

      The principal climatological factors that would most affect long-term lake-level variations have not been recorded for long periods at many locations. Regular precipitation observations were not made before about 1850. Some useful evidence is found in such natural records as tree rings and peat-bog stratigraphy.

      On a worldwide basis, there is evidence of a period of low levels in the middle 19th century and near the end of the first quarter of the 20th century. Lake George, in Australia, the Caspian Sea, several lakes in western North America, and Pangong Lake, in Tibet, are examples that have exhibited these features.

Lake extinction
      The life history of a lake may take place over just a few days, in the case of one formed by a beaver dam, or, for the largest lakes, it may cover geologic time periods. A lake may come to its end physically through loss of its water or through infilling by sediments and other materials. Reference has previously been made to the chemical-biological death of a lake, which is not necessarily the end of it as a physical entity but may in fact be its termination as a desirable body of water.

      Geologic processes involving the uplift and subsequent erosion of mountains and the advance and retreat of glaciers establish lake basins and then proceed to destroy them through infilling. Lake basins may also lose their water through drought or through changes in the drainage pattern that result in depletion of water inflows or enhancement of outflows.

      The chemical-biological changes within a lake's history offer a fine example of ecological succession. In the early stages a lake contains little organic material and has a poorly developed littoral zone. Particularly in temperate zones, such conditions favour a plentiful oxygen content, and the lake is said to be oligotrophic. As erosion progresses and as lake enrichment and organic content increase, the lake may become sufficiently productive to place an excessive demand upon the oxygen content. When periods of oxygen depletion occur, a lake is said to be eutrophic. An intermediate stage in this course of events is called mesotrophy. In the case of oligotrophy the vertical oxygen distribution is essentially uniform, or orthograde. Under eutrophic conditions, oxygen values decrease with depth, and the vertical distribution is called clinograde.

      The limits of oligotrophic and eutrophic conditions have been set in terms of the rate at which oxygen is depleted from the hypolimnion. These limits are arbitrary but are approximately 0.03 and 0.05 milligrams per square centimetre per day as the upper limit of oligotrophy and the lower limit of eutrophy, respectively.

      As eutrophic conditions develop, bottom sediments become enriched in organic material, and bottom plants spread throughout the littoral zone. As infilling proceeds, the plant-choked littoral zone spreads lakeward. Eventually, the littoral zone becomes a marsh, and the central part of the lake diminishes to a pond. When the lake finally ceases to exist, terrestrial vegetation may flourish, even to the extent of forestation.

Robert K. Lane

Major natural lakes of the world
       Major natural lakes of the world Major natural lakes of the worldA list of major natural lakes of the world is provided in the table.

Additional Reading
Mary J. Burgis and Pat Morris, The Natural History of Lakes (1987); and Laurence Pringle et al.,, Rivers and Lakes (1985), are two treatments written for the general reader. G. Evelyn Hutchinson, A Treatise on Limnology, vol. 1, Geography, Physics, and Chemistry (1957, reissued 1975), still a classic work, contains a comprehensive treatment of the physical and chemical aspects of lakes and includes an excellent bibliography of previous work. Less fundamental but more recent summaries of scientific lake studies include Gerald A. Cole, Textbook of Limnology, 4th ed. (1994); Charles R. Goldman and Alexander J. Horne, Limnology (1983); Robert G. Wetzel, Limnology, 2nd ed. (1983); Abraham Lerman (ed.), Lakes—Chemistry, Geology, Physics (1978); and Werner Stumm (ed.), Chemical Processes in Lakes (1985). Up-to-date books on hydrology and hydrogeology are R. Allan Freeze and John A. Cherry, Groundwater (1979); Patrick A. Domenico and Franklin W. Schwartz, Physical and Chemical Hydrogeology (1990); and Rafael L. Bras, Hydrology (1990). The sedimentology of lakes is the subject of J. McManus and R.W. Duck (eds.), Geomorphology and Sedimentology of Lakes and Reservoirs (1993). Prediction of the behaviour of lakes is dealt with in B. Henderson-Sellers, Engineering Limnology (1984); and A. James (ed.), An Introduction to Water Quality Modelling, 2nd ed. (1993). The dynamics of lakes are covered by K. Hutter (ed.), Hydrodynamics of Lakes (1984); James Lighthill, Waves in Fluids (1978); and Jerome Williams and Samuel A. Elder, Fluid Physics for Oceanographers and Physicists (1989). An up-to-date book on general freshwater biology is Brian Moss, Ecology of Fresh Waters, 2nd ed. (1988). Also of interest on this subject are the studies by C.S. Reynolds, The Ecology of Freshwater Phytoplankton (1984); and by Max M. Tilzer and Colette Serruya (eds.), Large Lakes: Ecological Structure and Function (1990). The consequences of building large-scale dams are analyzed in Edward Goldsmith and Nicholas Hildyard, The Social and Environmental Effects of Large Dams (1986). Kent W. Thornton, Bruce L. Kimmel, and Forrest E. Payne (eds.), Reservoir Limnology (1990), analyzes the ecology of reservoirs in their similarities to and differences from lakes. The effects of pollution on lakes are discussed in B. Henderson-Sellers and H.R. Markland, Decaying Lakes: The Origins and Control of Cultural Eutrophication (1987); E.B. Welch and T. Lindell, Ecological Effects of Wastewater: Applied Limnology and Pollution Effects (1992); Theodora E. Colborn et al., Great Lakes, Great Legacy? (1990); and William Ashworth, The Late, Great Lakes: An Environmental History (1986). Continuing research can be found in Limnology and Oceanography (bimonthly). Robert K. Lane Ed.

      any of a class of pigments composed of organic dyes that have been rendered insoluble by interaction with a compound of a metal. The interaction may involve the precipitation of a salt in which the proportions of dye to metal are fixed, or it may be a less well defined attraction between the dye and the surfaces of particles of the inorganic compound. Some lakes are prepared by a combination of both processes. Lakes considerably extend the range of colours available in the production of paints, cosmetics, and inks for printing and lithography.

      Dyes of several chemical classes are made into lakes by techniques that vary according to the nature of the salt-forming groups in the dye molecule. Mordant dyes and acid dyes form insoluble salts with metal ions, such as those of calcium and aluminum. Basic dyes contain amino groups and form insoluble salts with inorganic metal-containing acids such as phosphotungstic or phosphomolybdic acids.

* * *


Universalium. 2010.

Игры ⚽ Поможем написать курсовую

Look at other dictionaries:

  • Lake — bezeichnet niederdeutschen ein flaches stehendes Gewässer, hd. Lacke, siehe Weiher (Gewässer) Salzlake, eine Salzlösung Lake (Lippe), Nebenfluss der Lippe bei Lippstadt Lake ist der Familienname folgender Personen: Alan Lake (1940–1984),… …   Deutsch Wikipedia

  • Lake — (lago en idioma inglés) puede referirse a: Música Lake, grupo musical de rock alemán. Lake, álbum de Lake del 1976 Geografía Reino Unido Lake, parroquia civil de la Isla de Wight. Estados Unidos Lake, pueblo del Condado de Benzie, Míchigan Lake,… …   Wikipedia Español

  • Lake — Lake, n. [AS. lac, L. lacus; akin to AS. lagu lake, sea, Icel. l[ o]gr; OIr. loch; cf. Gr. la kkos pond, tank. Cf. {Loch}, {Lough}.] A large body of water contained in a depression of the earth s surface, and supplied from the drainage of a more… …   The Collaborative International Dictionary of English

  • Lake — (l[=a]k), n. [F. laque, fr. Per. See {Lac}.] A pigment formed by combining some coloring matter, usually by precipitation, with a metallic oxide or earth, esp. with aluminium hydrate; as, madder lake; Florentine lake; yellow lake, etc. [1913… …   The Collaborative International Dictionary of English

  • lake — Ⅰ. lake [1] ► NOUN 1) a large area of water surrounded by land. 2) (the Lakes) another name for the LAKE DISTRICT(Cf. ↑Lake District). DERIVATIVES lakelet noun …   English terms dictionary

  • lake — W3S3 [leık] n [Date: 1200 1300; : Old French; Origin: lac, from Latin lacus] 1.) a large area of water surrounded by land ▪ Lake Michigan 2.) wine/milk etc lake BrE a very large amount of wine, milk etc that has been produced but is not needed or …   Dictionary of contemporary English

  • lake — [ leık ] noun count ** a large area of water surrounded by land: There were some boys swimming in the lake. a. a large amount of liquid: a wine lake …   Usage of the words and phrases in modern English

  • Lake — Lake, MS U.S. town in Mississippi Population (2000): 408 Housing Units (2000): 159 Land area (2000): 1.102019 sq. miles (2.854217 sq. km) Water area (2000): 0.000000 sq. miles (0.000000 sq. km) Total area (2000): 1.102019 sq. miles (2.854217 sq.… …   StarDict's U.S. Gazetteer Places

  • Lake, MS — U.S. town in Mississippi Population (2000): 408 Housing Units (2000): 159 Land area (2000): 1.102019 sq. miles (2.854217 sq. km) Water area (2000): 0.000000 sq. miles (0.000000 sq. km) Total area (2000): 1.102019 sq. miles (2.854217 sq. km) FIPS… …   StarDict's U.S. Gazetteer Places

  • Lake — »Salzlösung zum Konservieren von Fischen und Fleisch«: Das seit dem 14. Jh. bezeugte mnd. lake »‹Herings›salzbrühe«, das im Rahmen des Heringshandels ins Hochd. übernommen wurde, ist identisch mit mnd. lake »stehendes Wasser« (vgl. 1↑ Lache) …   Das Herkunftswörterbuch

  • lake — lake1 [lāk] n. [ME < OE lacu & OFr lac, both < L lacus, a basin, lake < IE base * laku , accumulation of water, pond, lake > LOCH, OE lagu, water, sea] 1. an inland body of usually fresh water, larger than a pool or pond, generally… …   English World dictionary

Share the article and excerpts

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