glacier

glacier
glaciered, adj.
/glay"sheuhr/, n.
an extended mass of ice formed from snow falling and accumulating over the years and moving very slowly, either descending from high mountains, as in valley glaciers, or moving outward from centers of accumulation, as in continental glaciers.
[1735-45; < dial. F, deriv. of OF glace ice < LL glacia (for L glacies)]

* * *

I
Large mass of perennial ice that forms on land through the recrystallization of snow and that moves forward under its own weight.

The term ice sheet is commonly applied to a glacier that occupies an extensive tract of relatively level land and that flows from the centre outward. Glaciers occur where snowfall in winter exceeds melting in summer, conditions that prevail only in high mountain areas and polar regions. Glaciers occupy about 11% of the Earth's land surface but hold roughly three-fourths of its fresh water; 99% of glacier ice lies in Antarctica and Greenland.
II
(as used in expressions)

* * *

Introduction
 any large mass of perennial ice that originates on land by the recrystallization of snow or other forms of solid precipitation and that shows evidence of past or present flow.

      Exact limits for the terms large, perennial, and flow cannot be set. Except in size, a small snow patch that persists for more than one season is hydrologically indistinguishable from a true glacier. One international group has recommended that all persisting snow and ice masses larger than 0.1 square kilometre (about 0.04 square mile) be counted as glaciers.

General observations

Main types of glaciers
      Glaciers are classifiable in three main groups: (1) glaciers that extend in continuous sheets, moving outward in all directions, are called ice sheets if they are the size of Antarctica or Greenland and ice caps if they are smaller; (2) glaciers confined within a path that directs the ice movement are called mountain glaciers; and (3) glaciers that spread out on level ground or on the ocean at the foot of glaciated regions are called piedmont glaciers or ice shelves, respectively. Glaciers in the third group are not independent and are treated here in terms of their sources: ice shelves with ice sheets, piedmont glaciers with mountain glaciers. A complex of mountain glaciers burying much of a mountain range is called an ice field.

Distribution of glaciers
      A most interesting aspect of recent geological time (some 30 million years ago to the present) has been the recurrent expansion and contraction of the world's ice cover. These glacial fluctuations influenced geological, climatological, and biological environments and affected the evolution and development of early humans. Almost all of Canada, the northern third of the United States, much of Europe, all of Scandinavia, and large parts of northern Siberia were engulfed by ice during the major glacial stages. At times during the Pleistocene Epoch (from 1.8 million to 11,500 years ago), glacial ice covered 30 percent of the world's land area; at other times the ice cover may have shrunk to less than its present extent. It may not be improper, then, to state that the world is still in an ice age. Because the term glacial generally implies ice-age events or Pleistocene time, in this discussion “glacier” is used as an adjective whenever reference is to ice of the present day.

      Glacier ice today stores about three-fourths of all the fresh water in the world. Glacier ice covers about 11 percent of the world's land area and would cause a world sea-level rise of about 90 metres (300 feet) if all existing ice melted. Glaciers occur in all parts of the world and at almost all latitudes. In Ecuador, Kenya, Uganda, and Irian Jaya (New Guinea), glaciers even occur at or near the Equator, albeit at high altitudes.

Glaciers and climate
      The cause of the fluctuation of the world's glacier cover is still not completely understood. Periodic changes in the heat received from the Sun (solar energy), caused by fluctuations in the Earth's orbit, are known to correlate with major fluctuations of ice sheet advance and retreat on long time scales. Large ice sheets themselves, however, contain several “instability mechanisms” that may have contributed to the larger changes in world climate. One of these mechanisms is due to the very high albedo, or reflectivity of dry snow to solar radiation. No other material of widespread distribution on the Earth even approaches the albedo of snow. Thus, as an ice sheet expands it causes an ever larger share of the Sun's radiation to be reflected back into space, less is absorbed on the Earth, and the world's climate becomes cooler. Another instability mechanism is implied by the fact that the thicker and more extensive an ice sheet is, the more snowfall it will receive in the form of orographic precipitation (precipitation resulting from the higher altitude of its surface and attendant lower temperature). A third instability mechanism has been suggested by studies of the West Antarctic Ice Sheet. Portions of an ice sheet called ice streams may periodically move rapidly outward, perhaps because of the buildup of a thick layer of wet, deformable material under the ice. Although the ultimate causes of ice ages are not known with certainty, scientists agree that the world's ice cover and climate are in a state of delicate balance.

      Only the largest ice masses directly influence global climate, but all ice sheets and glaciers respond to changes in local climate—particularly changes in air temperature or precipitation. The fluctuations of these glaciers in the past can be inferred by features they have left on the landscape. By studying these features, researchers can infer earlier climatic fluctuations.

Formation and characteristics of glacier ice

Transformation of snow to ice
      Glacier ice is an aggregate of irregularly shaped, interlocking single crystals that range in size from a few millimetres to several tens of centimetres. Many processes are involved in the transformation of snowpacks to glacier ice, and they proceed at a rate that depends on wetness and temperature. Snow crystals in the atmosphere are tiny hexagonal plates, needles, stars, or other intricate shapes. In a deposited snowpack these intricate shapes are usually unstable, and molecules tend to evaporate off the sharp (high curvature) points of crystals and be condensed into hollows in the ice grains. This causes a general rounding of the tiny ice grains so that they fit more closely together. In addition, the wind may break off the points of the intricate crystals and thus pack them more tightly. Thus, the density of the snowpack generally increases with time from an initial low value of 50–250 kilograms per cubic metre (3–15 pounds per cubic foot). The process of evaporation and condensation may continue: touching grains may develop necks of ice that connect them (sintering) and that grow at the expense of other parts of the ice grain, or individual small grains may rotate to fit more tightly together. These processes proceed more rapidly at temperatures near the melting point and more slowly at colder temperatures, but they all result in a net densification of the snowpack. On the other hand, if a strong temperature gradient is present, water molecules may migrate from grain to grain, producing an array of intricate crystal shapes (known as depth hoar) of lowered density. If liquid water is present, the rate of change is many times more rapid because of the melting of ice from grain extremities with refreezing elsewhere, the compacting force of surface tension, refreezing after pressure melting (regulation), and the freezing of water between grains.

      This densification of the snow proceeds more slowly after reaching a density of 500–600 kilograms per cubic metre, and many of the processes mentioned above become less and less effective. Recrystallization under stress caused by the weight of the overlying snow becomes predominant, and grains change in size and shape in order to minimize the stress on them. This change usually means that large or favourably oriented grains grow at the expense of others. Stresses due to glacier flow may cause further recrystallization. These processes thus cause an increase in the density of the mass and in the size of the average grain.

      When the density of the aggregate reaches about 830 to 840 kilograms per cubic metre, the air spaces between grains are sealed off, and the material becomes impermeable to fluids. The time it takes for pores to be closed off is of critical importance for extracting climate-history information from ice cores. With time and the application of stress, the density rises further by the compression of air bubbles, and at great depths the air is absorbed into the ice crystal lattices, and the ice becomes clear. Only rarely in mountain glaciers does the density exceed 900 kilograms per cubic metre, but at great depths in ice sheets the density may approach that of pure ice (917 kilograms per cubic metre at 0° C and atmospheric pressure).

      Snow that has survived one melting season is called firn (or névé); its density usually is greater than 500 kilograms per cubic metre in temperate regions but can be as low as 300 kilograms per cubic metre in polar regions. The permeability change at a density of about 840 kilograms per cubic metre marks the transition from firn to glacier ice. The transformation may take only three or four years and less than 10 metres of burial in the warm and wet environment of Washington state in North America, but high on the plateau of Antarctica the same process takes several thousand years and burial to depths of about 150 metres.

Mass balance
 Glaciers are nourished mainly by snowfall, and they primarily waste away by melting and runoff or by the breaking off of icebergs (calving). In order for a glacier to remain at a constant size, there must be a balance between income (accumulation) and outgo (ablation). If this mass balance is positive (more gain than loss), the glacier will grow; if it is negative, the glacier will shrink.

      Accumulation refers to all processes that contribute mass to a glacier. Snowfall is predominant, but additional contributions may be made by hoarfrost (direct condensation of ice from water vapour), rime (freezing of supercooled water droplets on striking a surface), hail, the freezing of rain or meltwater, or avalanching of snow from adjacent slopes. Ablation refers to all processes that remove mass from a glacier. In temperate regions, melting (thermal fusion) at the surface normally predominates. Melting at the base is usually very slight (1 centimetre [0.4 inch] per year or less). Calving is usually the most important process on large glaciers in polar regions and on some temperate glaciers as well. Evaporation and loss by ice avalanches are important in certain special environments; floating ice may lose mass by melting from below.

      Because the processes of accumulation, ablation, and the transformation of snow to ice proceed so differently, depending on temperature and the presence or absence of liquid water, it is customary to classify glaciers in terms of their thermal condition. A polar glacier is defined as one that is below the freezing temperature throughout its mass for the entire year; a subpolar (or polythermal) glacier contains ice below the freezing temperature, except for surface melting in the summer and a basal layer of temperate ice; and a temperate glacier is at the melting temperature throughout its mass, but surface freezing occurs in winter. A polar or subpolar glacier may be frozen to its bed (cold-based), or it may be at the melting temperature at the bed (warm-based).

      Another classification distinguishes the surface zones, or facies, on parts of a glacier. In the dry-snow zone no surface melting occurs, even in summer; in the percolation zone some surface melting may occur, but the meltwater refreezes at a shallow depth; in the soaked zone sufficient melting and refreezing take place to raise the whole winter snow layer to the melting temperature, permitting runoff; and in the superimposed-ice zone refrozen meltwater at the base of the snowpack (superimposed ice) forms a continuous layer that is exposed at the surface by the loss of overlying snow. These zones are all parts of the accumulation area, in which the mass balance is always positive. Below the superimposed-ice zone is the ablation zone, in which annual loss exceeds the gain by snowfall. The boundary between the accumulation and ablation zones is called the equilibrium line.

      The value of the surface mass balance at any point on a glacier can be measured by means of stakes, snow pits, or cores. These values at points can then be averaged over the whole glacier for a whole year. The result is the net or annual mass balance. A positive value indicates growth, a negative value a decline.

Heat or energy balance
      The mass balance and the temperature variations of a glacier are determined in part by the heat energy received from or lost to the external environment—an exchange that takes place almost entirely at the upper surface. Heat is received from short-wavelength solar radiation, long-wavelength radiation from clouds or water vapour, turbulent transfer from warm air, conduction upward from warmer lower layers, and the heat released by the condensation of dew or hoarfrost or by the freezing of liquid water. Heat is lost by outgoing long-wavelength radiation, turbulent transfer to colder air, the heat required for the evaporation, sublimation, or melting of ice, and conduction downward to lower layers.

      In temperate regions, solar radiation is normally the greatest heat source (although much of the incoming radiation is reflected from a snow surface), and most of the heat loss goes to the melting of ice. It is incorrect to think of snow or ice melt as directly related to air temperature; it is the wind structure, the turbulent eddies near the surface, that determines most of the heat transfer from the atmosphere. In polar regions, heat is gained primarily from incoming solar radiation and lost by outgoing long-wavelength radiation, but heat conduction from lower layers and the turbulent transfer of heat to or from the air also are involved.

Glacier flow
 In the accumulation area the mass balance is positive year after year. Here the glacier would become thicker and thicker were it not for the compensating flow of ice away from the area (see video—>). This flow supplies mass to the ablation zone, compensating for the continual loss of ice there.

      Glacier flow is a simple consequence of the weight and creep properties of ice. Subjected to a shear stress over time, ice will undergo creep, or plastic deformation. The rate of plastic deformation under constant shear stress is initially high but tapers off to a steady value. If this steady value, the shear-strain rate, is plotted against the stress for many different values of applied stress, a curved graph will result. The curve illustrates what is known as the flow law or constitutive law of ice: the rate of shear strain is approximately proportional to the cube of the shear stress. Often called the Glen flow law by glaciologists, this constitutive law is the basis for all analyses of the flow of ice sheets and glaciers.

      As ice tends to build up in the accumulation area of a glacier, a surface slope toward the ablation zone is developed. This slope and the weight of the ice induce a shear stress throughout the mass. In a case with simple geometry, the shear stress can be given by the following formula:

      where τ is the shear stress, ρ the ice density, h the ice thickness, and α the surface slope. Each element of ice deforms (deformation and flow) according to the magnitude of the shear stress, as determined by (4), at a rate determined by the Glen flow law, stated above. By adding up, or integrating, the shear deformation of each element throughout the glacier thickness, a velocity profile can be produced. It can be given numerical expression as:

      where u1 is the surface velocity caused by internal deformation and k1 a constant involving ice properties and geometry. In this simple case, velocity is approximately proportional to the fourth power of the depth (h4). Therefore, if the thickness of a glacier is only slightly altered by changes in the net mass balance, there will be great changes in the rate of flow.

      Glaciers that are at the melting temperature at the base may also slide on the bed. Two mechanisms operate to permit sliding over a rough bed. First, small protuberances on the bed cause stress concentrations in the ice, an increased amount of plastic flow, and ice streams around the protuberances. Second, ice on the upstream side of protuberances is subjected to higher pressure, which lowers the melting temperature and causes some of the ice to melt; on the downstream side the converse is true, and meltwater freezes. This process, termed regelation, is controlled by the rate at which heat can be conducted through the bumps. The first process is most efficient with large knobs, and the second process is most efficient with small bumps. Together these two processes produce bed slip. Water-filled cavities may form in the lee of bedrock knobs, further complicating the process. In addition, studies have shown that sliding varies as the basal water pressure or amount changes. Although the process of glacier sliding over bedrock is understood in a general way, none of several detailed theories has been confirmed by field observation. This problem is largely unsolved.

      A formula in common use for calculating the sliding speed is:

      where u2 is the sliding speed at the base, pi and pa are the ice pressure and water pressure at the base of the ice, and k2 is another constant involving a measure of the roughness of the bed. The total flow of a glacier can thus be given by the sum of equations (5) and (6), u1 and u2. The total sum would be an approximation, because the formulas ignore longitudinal changes in velocity and thickness and other complicating influences, but it has proved to be useful in analyzing situations ranging from small mountain glaciers to huge ice sheets.

      Other studies have suggested that many glaciers and ice sheets do not slide on a rigid bed but “ride” on a deforming layer of water-charged sediment. This phenomenon is difficult to analyze because the sediment layer may thicken or thin, and thus its properties may change, depending on the history of deformation. In fact, the process may lead to an unsteady, almost chaotic, behaviour over time. Some ice streams in West Antarctica seem to have exhibited such unsteady behaviour.

Response of glaciers to climatic change (climate change)
      The relationship of glaciers and ice sheets to fluctuations in climate is sequential. The general climatic or meteorological environment determines the local mass and heat-exchange processes at the glacier surface, and these in turn determine the net mass balance of the glacier. Changes in the net mass balance produce a dynamic response—that is, changes in the rate of ice flow. The dynamic response causes an advance or retreat of the terminus, which may produce lasting evidence of the change in the glacier margin. If the local climate changes toward increased winter snowfall rates, the net mass balance becomes more positive, which is equivalent to an increase in ice thickness. The rate of glacier flow depends on thickness, so that a slight increase in thickness produces a larger increase in ice flow. This local increase in thickness and flow propagates down-glacier, taking some finite amount of time. When the change arrives at the terminus, it causes the margin of the glacier to extend farther downstream. The result is known as a glacier fluctuation—in this case an advance—and it incorporates the sum of all the changes that have taken place up-glacier during the time it took them to propagate to the terminus.

      The process, however, cannot be traced backward with assurance. A glacier advance can, perhaps, be related to a period of positive mass balances, but to ascertain the meteorological cause is difficult because either increased snowfall or decreased melting can produce a positive mass balance.

      The dynamic response of glaciers to changes in mass balance can be calculated several ways. Although the complete, three-dimensional equations for glacier flow are difficult to solve for changes in time, the effect of a small change or perturbation in climate can be analyzed readily. Such an analysis involves the theory of kinematic waves, which are akin to small pulses in one-dimensional flow systems such as floods in rivers or automobiles on a crowded roadway. The length of time it takes the glacier to respond in its full length to a change in the surface mass balance is approximately given as the ratio of ice thickness to (negative) mass balance at the terminus. The time scale for mountain glaciers is typically on the order of 10 to 100 years—although for thick glaciers or those with low ablation rates it can be much longer. Ice sheets normally have time scales several orders of magnitude longer.

Glaciers and sea level
      Sea level is currently rising at about 1.8 millimetres (0.07 inch) per year. Between 0.3 and 0.7 millimetres (0.01 to 0.03 inch) per year has been attributed to thermal expansion of ocean water, and most of the remainder is thought to be caused by the melting of glaciers and ice sheets on land. There is concern that the rate in sea-level rise may increase markedly in the future owing to global warming. Unfortunately, the state of the mass balance of the ice on the Earth is poorly known, so the exact contributions of the different ice masses to rising sea level is difficult to analyze. The mountain (small) glaciers of the world are thought to be contributing 0.2 to 0.4 millimetres (0.01 to 0.02 inch) per year to the rise. Yet the Greenland Ice Sheet is thought to be close to balance, the status of the Antarctic Ice Sheet is uncertain, and, although the floating ice shelves and glaciers may be in a state of negative balance, the melting of floating ice should not cause sea level to rise, and the grounded portions of the ice sheets seem to be growing. Thus, the cause of sea-level rise is still not well understood.

      With global warming, the melting of mountain glaciers will certainly increase, although this process is limited: the total volume of small glaciers is equivalent to only about 0.6 metre (2 feet) of sea-level rise. Melting of the marginal areas of the Greenland Ice Sheet will likely occur under global warming conditions, and this will be accompanied by the drawing down of the inland ice and increased calving of icebergs; yet these effects may be counterbalanced to some extent by increased snow precipitation on the inland ice. The Antarctic Ice Sheet, on the other hand, may actually serve as a buffer to rising sea level: increased melting of the marginal areas will probably be exceeded by increased snow accumulation due to the warmer air (which holds more moisture) and decreased sea ice (bringing moisture closer to the ice sheet). Modeling studies that predict sea-level rise up to the time of the doubling of greenhouse gas concentrations (i.e., concentrations of atmospheric carbon dioxide, methane, nitrous oxide, and certain other gases) about the year 2050 suggest a modest rise of about 0.3 metre (1 foot).

The great ice sheets
      Two great ice masses, the Antarctic and Greenland ice sheets, stand out in the world today and may be similar in many respects to the large Pleistocene ice sheets. About 99 percent of the world's glacier ice is in these two ice masses, 91 percent in Antarctica alone.

Antarctic Ice Sheet
Dimensions
 The bedrock of the continent of Antarctica is almost completely buried under ice. Mountain ranges and isolated nunataks (a term derived from Greenland's Inuit language, used for individual mountains surrounded by ice (ice shelf)) locally protrude through the ice. Extensive in area are the ice shelves, where the ice sheet extends beyond the land margin and spreads out to sea. The ice sheet, with its associated ice shelves, covers an area of 13,829,000 square kilometres (5,340,000 square miles); exposed rock areas total less than 200,000 square kilometres. The mean thickness of the ice is about 1,829 metres (6,000 feet) and the volume of ice more than 25.4 million cubic kilometres (6 million cubic miles). The land surface beneath the ice is below sea level in many places, but this surface is depressed because of the weight of the ice. If the ice sheet were melted, uplift of the land surface would eventually leave only a few deep troughs and basins below sea level—even though the sea level itself also would rise about 80 metres from the addition of such a large amount of water. Because of the thick ice cover, Antarctica has by far the highest mean altitude of the continents (2 kilometres [1.3 miles]); all other continents have mean altitudes less than 1 kilometre (0.6 mile).

      Antarctica can be divided into three main parts: the smallest and the mildest in climate is the Antarctic Peninsula, extending from latitude 63° S off the tip of South America to a juncture with the main body of West Antarctica at a latitude of about 74° S. The ice cover of the Antarctic Peninsula is a complex of ice caps, piedmont and mountain glaciers, and small ice shelves.

West Antarctica
      The part of the main continent lying south of the Americas, between longitudes 45° W and 165° E, is characterized by irregular bedrock and ice-surface topography and numerous nunataks and deep troughs. Two large ice shelves occur in West Antarctica: the Filchner (Filchner Ice Shelf)-Ronne (Ronne Ice Shelf) Ice Shelf (often considered to be two separate ice shelves), south of the Weddell Sea, and the Ross Ice Shelf, south of the Ross Sea. Each has an area exceeding 500,000 square kilometres.

East Antarctica
      The huge ice mass of East Antarctica, about 10,200,000 square kilometres, is separated from West Antarctica by the Transantarctic Mountains. This major mountain range extends from the eastern margin of the Ross Ice Shelf almost to the Ronne-Filchner Ice Shelf. The bedrock of East Antarctica is approximately at sea level, but the ice surface locally exceeds 4,000 metres above sea level on the highest parts of the polar plateau.

Climatic conditions
      At the South Pole the snow surface is 2,800 metres in altitude, and the mean annual temperature is about -50° C (-58° F), but at the Russian Vostok Station (78°27′ S, 106°52′ E), 3,500 metres above sea level, the mean annual temperature is -58° C (-73° F), and in August 1960 (the winter season) the temperature reached a low of -88.3° C (-127° F). The temperatures on the polar plateau of Εast Αntarctica are by far the coldest on Εarth; the climate of the Arctic is quite mild by comparison. Along the coast of East or West Antarctica, where the climate is milder, mean annual temperatures range from -20° to -9° C (-4° to 16° F), but temperatures exceed the melting point only for brief periods in summer, and then only slightly. Katabatic (katabatic wind) (drainage) winds, however, are very strong along the coast; the mean annual wind speed at Commonwealth Bay is 20 metres per second (45 miles per hour).

      The Greenland Ice Sheet, though subcontinental in size, is huge compared with other glaciers in the world except that of Antarctica. Greenland is mostly covered by this single large ice sheet (1,730,000 square kilometres), while isolated glaciers and small ice caps totaling between 76,000 and 100,000 square kilometres occur around the periphery. The ice sheet is almost 2,400 kilometres long in a north-south direction, and its greatest width is 1,100 kilometres at a latitude of 77° N, near its northern margin. The mean altitude of the ice surface is 2,135 metres. The term Inland Ice, or, in Danish, Indlandsis, is often used for this ice sheet.

      The bedrock surface is near sea level over most of the interior of Greenland, but mountains occur around the periphery. Thus, this ice sheet, in contrast to the Antarctic Ice Sheet, is confined along most of its margin. The ice surface reaches its greatest altitude on two north-south elongated domes, or ridges. The southern dome reaches almost 3,000 metres at latitudes 63°–65° N; the northern dome reaches about 3,290 metres at about latitude 72° N. The crests of both domes are displaced east of the centre line of Greenland.

      The unconfined ice sheet does not reach the sea along a broad front anywhere in Greenland, and no large ice shelves occur. The ice margin just reaches the sea, however, in a region of irregular topography in the area of Melville Bay southeast of Thule. Large outlet glaciers, which are restricted tongues of the ice sheet, move through bordering valleys around the periphery of Greenland to calve off into the ocean, producing the numerous icebergs that sometimes penetrate North Atlantic shipping lanes. The best known of these is the Jakobshavn Glacier, which, at its terminus, flows at speeds of 20 to 22 metres per day.

      The climate of the Greenland Ice Sheet, though cold, is not as extreme as that of central Antarctica. The lowest mean annual temperatures, about -31° C (-24° F), occur on the north-central part of the north dome, and temperatures at the crest of the south dome are about -20° C (-4° F).

Mass balance of the ice sheets
      The rate of precipitation on the Antarctic Ice Sheet is so low that it may be called a cold desert. Snow accumulation over much of the vast polar plateau is less than five centimetres (two inches) water equivalent per year. Only around the margin of the continent, where cyclonic storms penetrate frequently, does the accumulation rise to values of more than 30 centimetres. The mean for Antarctica is 15 centimetres or less. In Greenland values are higher: less than 15 centimetres in a comparatively small area of north-central Greenland, 30 centimetres along the crests of the domes, and more than 80 centimetres along the southeast and southwest margins; the mean annual snow accumulation is about 37 centimetres of water equivalent.

      Snow accumulation occurs mainly as direct snowfall when cyclonic storms move inland. At high altitudes on the Greenland Ice Sheet and in central Antarctica, ice crystals form in the cold air during clear periods and slowly settle out as fine “diamond dust.” Hoarfrost and rime deposition are generally minor items in the snow-accumulation totals. It is almost impossible to measure the precipitation directly in these climates; precipitation gauges are almost useless for the measurement of blowing snow, and the snow is blown about almost constantly in some areas. The thickness and density of snow deposited on the ground equals precipitation plus hoarfrost and rime deposition, less evaporation, less snow blown away, and plus snow blown in from somewhere else. The last two phenomena are thought to cancel each other approximately—except in the coastal areas, where fierce drainage, or katabatic, winds move appreciable quantities of snow out to sea.

      The snow surface may be smooth where soft powder snow is deposited with little wind, or very hard packed and rough when high winds occur during or after snowfall. Two features are prominent: snow dunes are depositional features resembling sand dunes in their several shapes; sastrugi are jagged erosional features (often cut into snow dunes) caused by strong prevailing winds that occur after snowfall. Sharp, rugged sastrugi, which can be one to two metres high, make travel by vehicle or on foot difficult. The annual snow layers exposed in the side of a snow pit can usually be distinguished by a low density layer (depth hoar) that forms by the burial of surface hoarfrost or by metamorphism of the snow deposited in the fall at a time when the temperature is changing rapidly.

      Almost all of the Antarctic Ice Sheet lies within the dry-snow zone. The percolation, soaked, and superimposed ice zones occur only in a very narrow strip in a small area along the coast. In Greenland only the central part of the northern half of the ice sheet, or about 30 percent of the total area, is within the dry-snow zone. Almost half of the area of the Greenland ice sheet is considered to be in the percolation zone. In flat areas near the equilibrium line, especially in west-central Greenland, there are notorious snow swamps, or slush fields, in summer; some of this water runs off, but much of it refreezes. (For an explanation of a glacier's surface zones, see above Formation and characteristics of glacier ice: Mass balance (glacier).)

      The ice sheets lose material by several processes, including surface melting, evaporation, wind erosion (deflation), iceberg calving, and the melting of the bottom surfaces of floating ice shelves by warmer seawater.

      In Antarctica, calving of ice shelves and outlet glacier tongues clearly predominates among all the processes of ice loss, but calving is very episodic and cannot be measured accurately. The amount of surface melt and evaporation is small, amounting to about 22 centimetres of ice lost from a five-kilometre ring around half the continent. Wind erosion is difficult to evaluate but probably accounts for only a very small loss in the mass balance. The undersides of ice shelves near their outer margins are subject to melting by the ocean water. The rate of melting decreases inland, and at that point some freezing of seawater onto the base of the ice shelves must occur, but farther inland, near the grounding line, the tidal circulation of warm seawater may produce basal melting.

      In Greenland, surface melt is more important, calving is less so, and undershelf melting is important only on floating glacier tongues (seaward projections of a glacier). Most of the calving is from the termini of a relatively few large, fast-moving outlet glaciers. In Greenland, vertical-walled melt pits in the ice are a well-known feature of the ice surface at the ablation zone. Ranging from a few millimetres to a metre in diameter, these pits are floored with a dark, silty material called cryoconite, once thought to be of cosmic origin but now known to be largely terrestrial dust. The vertical melting of the holes is due to the absorption of solar radiation by the dark silt, possibly augmented by biological activity.

Net mass balance
      Because two great ice sheets contain 99 percent of the world's ice, it is important to know whether this ice is growing or shrinking under present climatic conditions. Although just such a determination was a major objective of the International Geophysical Year (1957–58) and more has been learned each year since, even the sign of the net mass balance has not yet been determined conclusively.

      It appears that accumulation on the surface of the Antarctic Ice Sheet is approximately balanced by iceberg calving and basal melting from the ice shelves. Compilations from many authors and the Intergovernmental Panel on Climate Change (IPCC), Third Scientific Assessment (2001), suggest the following average values, given in gigatons (billions of tons) per year (1 gigaton is equivalent to 1.1 cubic kilometres of water):

      Accumulation

      Accumulation on grounded ice+ 1,829±87

      Accumulation on grounded ice and ice shelves+ 2,233±86

      Ablation

      Calving of ice shelves and glaciers− 2,072±304

      Bottom melting, ice shelves− 540±218

      Melting and runoff− 10±10

      Net mass balance− 389±384

      The net difference, however, is on the same order as the margin of error in estimating the various quantities. Furthermore, some authors have suggested that the values stated above for calving and ice-shelf melting are too high and that the discharge of ice to the sea, as measured by ice-flow studies, is clearly less than the accumulation. Thus, even the sign of the net balance is not well defined. It appears that the net balance of the grounded portion of the Antarctic Ice Sheet is positive, while that of the floating ice shelves is negative. Studies of fluctuations in the extent of floating ice have been inconclusive.

      The net mass balance of the Greenland Ice Sheet also appears to be close to zero, but here, too, the margin of error is too large for definite conclusions. The estimated balance is as follows, again from the IPCC and in gigatons per year.

      Accumulation

      Snow accumulation± 21

      Ablation

      Iceberg calving− 235±33

      Melting and runoff− 297±32

      Bottom melting− 32± 3

      Net mass balance− 42±51

      Uncertainties in the quantities given above are due to the difficulty of analyzing the spatial and temporal distributions of accumulation, the relatively few annual measurements of iceberg calving, and a lack of knowledge of the amount of surface meltwater that refreezes in the cold snow and ice at depth. Many of the outlet glaciers and portions of the ice-sheet margin in the southwestern part of Greenland, where many observations have been made, have stopped the retreats that were observed from the 1950s through the 1970s. After a period of relative stability and advance during the 1980s, glacier retreats have both resumed and accelerated in Greenland since the mid-1990s.

Flow of the ice sheets
      In general, the flow of the Antarctic and Greenland ice sheets is not directed radially outward to the sea. Instead, ice from central high points tends to converge into discrete drainage basins and then concentrate into rapidly flowing ice streams. (Such so-called streams are currents of ice that move several times faster than the ice on either side of them.) The ice of much of East Antarctica has a rather simple shape with several subtle high points or domes. Greenland resembles an elongated dome, or ridge, with two summits. West Antarctica is a complex of converging and diverging flow because of the jumble of ridges and troughs in the subglacial bedrock and the convergence of ice streams.

      Flow rates in the interior of an ice sheet are very low, being measured in centimetres or metres per year, because the surface slope is minuscule and the ice is very cold. As the ice moves outward, the rate of flow increases to a few tens of metres per year, and this rate of flow increases still further, up to one kilometre per year, as the flow is channeled into outlet glaciers or ice streams. Ice shelves continue the flow and even cause it to increase, because ice spreads out in ever thinner layers. At the edge of the Ross Ice Shelf, ice is moving out about 900 metres per year toward the ocean.

      This simple picture of ice flow is made more complicated by the dependence of the flow law of ice on temperature. Because a temperature increase of about 15° C (27° F) causes a 10-fold increase in the deformation rate of ice, the temperature distribution of an ice sheet partly determines its flow structure. The cold ice of the central part of an ice sheet is carried down into warmer zones. This shift modifies the static temperature distribution, and the shear deformation is concentrated in a thin zone of warmer ice at the base. The forward velocity may be almost uniform throughout the depth to within a few tens or hundreds of metres from the bedrock.

      Another important effect on ice flow is the heat produced by friction, caused by the sliding of the ice on bedrock or by internal shearing within the basal ice. If a portion of the ice sheet deforms more rapidly than its surroundings, the slight amount of extra heat production raises the temperature of this portion, causing it to deform even more readily. This increased deformation may explain the phenomena of ice streams. Ice streams are very effective in moving ice from large drainage areas of Antarctica and Greenland out to ice shelves or to the sea. It is known that at least one Antarctic ice stream moves rapidly on a layer of water-charged deforming sediment; a nearby ice stream appears to have ceased rapid movement in the past several hundred years, perhaps owing to loss of its sediment layer.

Information from deep cores (core sampling)
      Most of the Antarctic and Greenland ice sheets are below freezing throughout. Continuous cores, taken in some cases to the bedrock below, allow the sampling of an ice sheet through its entire history of accumulation. Records obtained from these cores represent exciting new developments in paleoclimatology and paleoenvironmental studies. Because there is no melting, the layered structure of the ice preserves a continuous record of snow accumulation and chemistry, air temperature and chemistry, and fallout from volcanic, terrestrial, marine, cosmic, and man-made sources. Actual samples of ancient atmospheres are trapped in air bubbles within the ice. This record extends back more than 400,000 years.

      Near the surface it is possible to pick out annual layers by visual inspection. In some locations, such as the Greenland Ice core Project/Greenland Ice Sheet Project 2 (GRIP/GISP2) sites at the summit of Greenland, these annual layers can be traced back more than 40,000 years, much like counting tree rings. The result is a remarkably high-resolution record of climatic change. When individual layers are not readily visible, seasonal changes in dust, marine salts, and isotopes can be used to infer annual chronologies. Precise dating of recent layers can be accomplished by locating radioactive fallout from known nuclear detonations or traces of volcanic eruptions of known date. Other techniques must be used to reconstruct a chronology from some very deep cores. One method involves a theoretical analysis of the flow. If the vertical profile of ice flow is known, and if it can be assumed that the rate of accumulation has been approximately constant through time, then an expression for the age of the ice as a function of depth can be developed.

      A very useful technique (isotopic fractionation) for tracing past temperatures involves the measurement of oxygen isotopes—namely, the ratio of oxygen-18 to oxygen-16. Oxygen-16 is the dominant isotope, making up more than 99 percent of all natural oxygen; oxygen-18 makes up 0.2 percent. However, the exact concentration of oxygen-18 in precipitation, particularly at high latitudes, depends on the temperature. Winter snow has a smaller oxygen-18–oxygen-16 ratio than does summer snow. A similar isotopic method for inferring precipitation temperature is based on measuring the ratio of deuterium (hydrogen-2 (hydrogen)) to normal hydrogen (hydrogen-1). The relation between these oxygen and hydrogen isotopic ratios, termed the deuterium excess, is useful for inferring conditions at the time of evaporation and precipitation. The temperature scale derived from isotopic measurements can be calibrated by the observable temperature-depth record near the surface of ice sheets.

      Results of ice core measurements are greatly extending the knowledge of past climates. For instance, air samples taken from ice cores show an increase in methane, carbon dioxide, and other “greenhouse gas” concentrations with the rise of industrialization and human population. On a longer time scale, the concentration of carbon dioxide in the atmosphere can be shown to be related to atmospheric temperature (as indicated by oxygen and hydrogen isotopes)—thus confirming the global-warming greenhouse effect, by which heat in the form of long-wave infrared radiation is trapped by atmospheric carbon dioxide and reflected back to the Earth's surface.

      Perhaps most exciting are recent ice core results that show surprisingly rapid fluctuations in climate (climate change), especially during the last glacial period (160,000 to 10,000 years ago) and probably in the interglacial period that preceded it. Detectable variations in the dustiness of the atmosphere (a function of wind and atmospheric circulation), temperature, precipitation amounts, and other variables show that, during this time period, the climate frequently alternated between full-glacial and nonglacial conditions in less than a decade. Some of these changes seem to have occurred as sudden climate fluctuations, called Dansgaard-Oeschger events, in which the temperature jumped 5° to 7° C (9° to 13° F), remained in that state for a few years to centuries, jumped back, and repeated the process several times before settling into the new state for a long time—perhaps 1,000 years. These findings have profound and unsettling implications for the understanding of the coupled ocean-atmosphere climate system.

Mountain glaciers
 In this discussion the term mountain glaciers includes all perennial ice masses other than the Antarctic and Greenland ice sheets. These ice masses are not necessarily associated with mountains. Sometimes the term small glaciers is used, but only in a relative sense: a glacier 10,000 square kilometres (4,000 square miles) in surface area would not be called “small” in many parts of the world.

Classification of mountain glaciers
      Mountain glaciers are generally confined to a more or less marked path directing their movement. The shape of the channel and the degree to which the glacier fills it determine the type of glacier. Valley glaciers are a classic type; they flow at least in part down a valley and are longer than they are wide. Cirque glaciers, short and wide, are confined to cirques, or amphitheatres, cut in the mountain landscape. Other types include transection glaciers or ice fields, which fill systems of valleys, and glaciers in special situations, such as summit glaciers, hanging glaciers, ice aprons, crater glaciers, and regenerated or reconstituted glaciers. Glaciers that spread out at the foot of mountain ranges are called piedmont glaciers. Outlet glaciers are valley glaciers that originate in ice sheets, ice caps, and ice fields. Because of the complex shapes of mountain landscapes and the resulting variety of situations in which glaciers can develop, it is difficult to draw clear distinctions among the various types of glaciers.

      Mountain glaciers also are classified as polar, subpolar, or temperate and their surfaces by the occurrence of dry-snow, percolation, saturation, and superimposed-ice zones, as for ice sheets.

Surface features
      The snow surface of the accumulation area of a mountain glacier displays the same snow dune and sastrugi features found on ice sheets, especially in winter, but normally these features are neither as large nor as well developed. Where appreciable melting of the snow occurs, several additional features may be produced. During periods of clear, sunny weather, sun cups (cup-shaped hollows usually between 5 and 50 centimetres [2 and 20 inches] in depth) may develop. On very high-altitude, low-latitude snow and firn fields these may grow into spectacular narrow blades of ice, up to several metres high, called nieves penitentes. Rain falling on the snow surface (or very high rates of melt) may cause a network of meltwater runnels (shallow grooves trending downslope) to develop.

      Other features are characteristic of the ablation zone. Below icefalls (steep reaches of a valley glacier), several types of curved bands can be seen. The surface of the glacier may rise and fall in a periodic manner, with the spacing between wave crests approximately equal to the amount of ice flow in a year. Called wave ogives (pointed arches), these arcs result from the great stretching of the ice in the rapidly flowing icefall. The ice that moves through the icefall in summer has more of its surface exposed to melting and is greatly reduced in volume compared with the ice moving through in winter. Dirtband ogives also may occur below icefalls; these are caused by seasonal differences in the amount of dust or by snow trapped in the icefall. In plan view, the ogives are invariably distorted into arcs or curves convex downglacier; hence the name ogive.

      The ice of the ablation zone normally shows a distinctive layered structure. This can be relict stratification developed by the alternation of dense and light or of clean and dirty snow accumulations from higher on the glacier. This stratification is later subdued by recrystallization accompanying plastic flow. A new layering called foliation is developed by the flow. Foliation is expressed by alternating layers of clear and bubbly or coarse-grained and fine-grained ice. Although the origin of this structure is not fully understood, it is analogous to the process that produces foliated structures in metamorphic rocks.

      The ice crystals in strongly deformed (deformation and flow), foliated ice invariably have a preferred orientation, relative to the stress directions. In some situations, more often in polar than in temperate ice, the hexagonal axes are aligned perpendicularly to the plane of foliation. This alignment places the crystal glide planes parallel to the planes of (presumed) greatest shearing. In many other locations the hexagonal crystal axes are preferentially aligned in four different directions, none perpendicular to the foliation. This enigmatic pattern has resisted explanation so far.

      Crevasses (crevasse) are common to both the accumulation and ablation zones of mountain glaciers, as well as of ice sheets. Transverse crevasses, perpendicular to the flow direction along the centre line of valley glaciers, are caused by extending flow. Splaying crevasses, parallel to the flow in midchannel, are caused by a transverse expansion of the flow. The drag of the valley walls produces marginal crevasses, which intersect the margin at 45°. Transverse and splaying crevasses curve around to become marginal crevasses near the edge of a valley glacier. Splaying and transverse crevasses may occur together, chopping the glacier surface into discrete blocks or towers, called seracs.

      Crevasses deepen until the rate of surface stretching is counterbalanced by the rate of plastic flow tending to close the crevasses at depth. Thus, crevasse depths are a function of the rate of stretching and the temperature of the ice. Crevasses deeper than 50 metres (160 feet) are rare in temperate mountains, but crevasses to 100 metres or more in depth may occur in polar regions. Often the crevasses are concealed by a snow bridge, built by accumulations of windblown snow.

Mass balance of mountain glaciers
      The rate of accumulation and ablation on mountain glaciers depends on latitude, altitude, and distance downwind from sources of abundant moisture, such as the oceans. The glaciers along the coasts of Washington, British Columbia, southeastern Alaska, South Island of New Zealand, Iceland, and southwestern Norway receive prodigious snowfall. Snow accumulation of three to five metres of water equivalent in a single season is not uncommon. With this large income, glaciers can exist at low altitudes in spite of very high melt rates. The rate of snowfall increases with increasing altitude; thus, the gradient of net mass balance with altitude is steep. This gradient also expresses the rate of transfer of mass by glacier flow from high to low altitudes and is called the activity index.

      Typical of the temperate, maritime glaciers is South Cascade Glacier, in western Washington. Its activity index is high, normally about 17 millimetres per metre (0.2 inch per foot); the yearly snow accumulation averages about 3.1 metres of water-equivalent; and the equilibrium line is at the relatively low altitude of 1,900 metres. This glacier contains only ablation and saturation zones; the winter chill is so slight that no superimposed ice is formed.

      In the maritime environment of southeastern Alaska are many very large glaciers; Bering and Seward-Malaspina glaciers (piedmont glaciers) cover about 5,800 and 5,200 square kilometres (2,200 and 2,000 square miles) in area, respectively. Equilibrium lines are lower than those in Washington state, but the rates of accumulation and ablation and the activity indices are about the same. Because these mountains are high, and some glaciers extend over a great range of altitude, all surface zones except the dry-snow zone are represented.

      In more continental (inland) environments, the rate of snowfall is much less, and the summer climate is generally warmer. Thus, glaciers can exist only at high altitudes. High winds may concentrate the meagre snowfall in deep, protected basins, however, allowing glaciers to form even in areas of low precipitation and high melt rates. Glaciers formed almost entirely of drift snow occur at high altitudes in Colorado and in the polar Ural Mountains and are often referred to as Ural-type glaciers. Superimposed ice and soaked zones are found in the accumulation area; in higher areas the percolation zone is found, and in some local extreme areas the dry-snow zone occurs. Because of the decrease in melt rates, continental glaciers in high latitudes occur at lower altitudes and have lower accumulation totals and activity indices. McCall Glacier, in the northwestern part of the Brooks Range in Alaska, has the lowest activity index (two millimetres per metre) measured in western North America. Glaciers in intermediate climates have intermediate equilibrium-line altitudes, accumulation or ablation totals, and activity indices.

Flow of mountain glaciers
      Ice flow in valley glaciers has been studied extensively. The first measurements date from the mid-18th century, and the first theoretical analyses date from the middle of the 19th century. These glaciers generally flow at rates of 0.1 to 2 metres per day, faster at the surface than at depth, faster in midchannel than along the margins, and usually fastest at or just below the equilibrium line. Cold, polar glaciers flow relatively slowly, because the constitutive law of ice is sensitive to temperature and because they generally are frozen to their beds. In some high-latitude areas, such as the Svalbard archipelago north of Norway, polythermal glaciers are common; these consist of subfreezing ice overlying temperate ice, and, because they are warm-based, they actively slide on their beds.

      The fastest glaciers (other than those in the act of surging) are thick, temperate glaciers in which high subglacial water pressures produce high rates of sliding. Normal temperate glaciers ending on land generally have subglacial water pressures in the range of 50 to 80 percent of the ice pressure, but glaciers that end in the sea may have subglacial water pressures almost equal to the ice pressure—that is, they almost float. The lower reach of Columbia Glacier in southern Alaska, for instance, flows between 20 and 30 metres (66 and 100 feet) per day, almost entirely by sliding. Such a high sliding rate occurs because the glacier, by terminating in the ocean, must have a subglacial water pressure high enough to drive water out of the glacier against the pressure of the ocean water.

Glacier hydrology (hydrologic cycle)
      A temperate glacier is essentially a reservoir that gains precipitation in both liquid and solid form, stores a large share of this precipitation, and then releases it with little loss at a later date. The hydrologic characteristics of this reservoir, however, are complex, because its physical attributes change during a year.

      In late spring the glacier is covered by a thick snowpack at the melting temperature. Meltwater and liquid precipitation must travel through the snowpack by slow percolation until reaching well-defined meltwater channels in the solid ice below. In summer the snowpack becomes thinner, and drainage paths within the snow are more defined, so that meltwater and liquid precipitation are transmitted through the glacier rapidly. In winter, snow accumulates, and the surface layer freezes, stopping the movement of meltwater and precipitation at the surface. The rest of the ice reservoir may continue to drain, but in the process the conduits within and under the ice tend to close.

      The runoff from a typical Northern Hemisphere temperate glacier reaches a peak in late July or early August. Solar radiation, the chief source of heat to promote melt, reaches a peak in June. The delay in the peak melt rates is primarily because of the changing albedo (surface reflectivity) during the summer; initially the snow is very reflective and covers the whole glacier, but as the summer wears on the snow becomes wet (less reflective), and in addition more and more ice of much lower albedo is exposed. Thus, even though the incoming radiation decreases during midsummer, the proportion of it that is absorbed to cause melt is greatly increased. Other heat-exchange processes, such as turbulent transfer from warm air, also become more important during midsummer and late summer.

      This albedo variation produces a runoff “buffering effect” against unusually wet or dry years. An unusually heavy winter snowpack (snow) causes high-albedo snow to persist longer over the glacier in summer; thus, less meltwater is produced. Conversely, an unusually light winter snowfall causes older firn and ice of lower albedo to be exposed earlier in the summer, producing increased melt and runoff. Thus, glaciers naturally regulate the runoff, seasonally and from year to year. When glacier runoff is combined with nonglacier runoff in roughly equal amounts, the result is very stable and even streamflow. This condition is part of the basis for the extensive hydrologic development that is found in regions such as the Alps, Norway, and western Washington.

      Glacier streams (river) are characterized by high sediment concentrations. The sediment ranges from boulders to a distinctive fine-grained material called rock flour, or glacier flour, which is colloidal in size (often less than one micrometre in diameter). The suspended sediment concentration decreases with distance from the glacier, but the rock-flour component may persist for great distances and remain suspended in lakes for many years; it is responsible for the green colour of Alpine lakes. Glacier streams vary in discharge with the time of day, and this variation causes a continual readjustment of the stream channel and the transportation of reworked debris, adding to the sediment load. Rates of glacier erosion (that is, sediment production) are typically on the order of one millimetre per year, averaged over the glacier area, but they are higher in particularly steep terrain or where the bedrock is especially soft.

Glacier floods
      Glacier outburst floods, or jökulhlaups, can be spectacular or even catastrophic. These happen when drainage within a glacier is blocked by internal plastic flow and water is stored in or behind the glacier. The water eventually finds a narrow path to trickle out. This movement will cause the path to be enlarged by melting, causing faster flow, more melting, a larger conduit, and so on until all the water is released quite suddenly. The word jökulhlaup is Icelandic (Iceland) in origin, and Iceland has experienced some of the world's most spectacular outburst floods. The 1922 Grimsvötn outburst released about 7.1 cubic kilometres (1.7 cubic miles) of water in a flood that was estimated to have reached almost 57,000 cubic metres (2,000,000 cubic feet) per second. Outburst floods occur in many glacier-covered mountain ranges; some break out regularly each year, some at intervals of two or more years, and some are completely irregular and impossible to predict.

Glacier surges
      Most glaciers follow a regular and nonspectacular pattern of advance and retreat in response to a varying climate. A very different behaviour pattern has been reported for glaciers in certain, but not all, areas. Such glaciers may, after a period of normal flow, or quiescence, lasting 10 to 100 or more years, suddenly begin to flow very rapidly, to up to five metres per hour. This rapid flow, lasting only a year or two, causes a sudden depletion of the upper part of the glacier, accompanied by a swelling and advance of the lower part, although these usually do not reach positions beyond the limits of previous surges. Advances of several kilometres in as many months have been recorded. Even more interesting is the fact that these glaciers periodically repeat cycles of quiescence and activity, irrespective of climate. These unusual glaciers are called surging glaciers.

      Although surging glaciers are not rare in some areas (e.g., Alaska Range and St. Elias Mountains), they are totally absent in other areas of similar topography, bedrock, climate, and so forth (e.g., western Chugach Mountains and Coast Mountains). Furthermore, glaciers of all shapes and sizes, from tiny cirque glaciers to major portions of a large ice cap, have been known to surge. The flow instability that results in glacier surges is generally caused by an abrupt decoupling of the glacier from its bed. This decoupling is the result of a breakdown in the normal subglacier water flow system, but the exact mechanisms that cause some glaciers to surge are not fully understood.

Tidewater glaciers
      Many glaciers terminate in the ocean with the calving of icebergs. Known as tidewater glaciers, these glaciers are the seaward extensions of ice streams originating in ice fields, ice caps, or ice sheets. Some tidewater glaciers are similar to surging glaciers in that they flow at high speeds—as much as 35 metres (115 feet) per day—but they do so continuously. Tidewater glaciers share another characteristic with surging glaciers in that they may advance and retreat periodically, independent of climatic variation.

      The physical mechanisms that control the rate of iceberg calving are not yet well understood. Empirical studies of grounded (not floating) tidewater glaciers in Alaska, Svalbard, and elsewhere suggest that the speed of iceberg calving is roughly proportional to water depth at the terminus. This relation can produce an instability and periodic advance-retreat cycles. For example, a glacier terminating in shallow water at the head of a fjord will have a low calving speed that may be exceeded by the ice flow speed, causing advance of the terminus. At the same time, glacial erosion will cause the deposition of sediment as a moraine shoal at the terminus. With time, the glacier will advance, eroding the shoal on the upstream face and depositing sediment on the downstream face. The shoal, by reducing the depth of the water at the glacier's terminus and thereby inhibiting iceberg calving, will allow the glacier to advance into deep water farther down the fjord. This advance phase is slow—typically 9 to 40 metres (30 to 130 feet) per year—and in an Alaskan fjord it may take a period of 1,000 years or more to cover a typical fjord length of 30 to 130 kilometres (20 to 80 miles).

      Such a glacier, in an extended position and terminating in shallow water on a moraine shoal, is in an unstable situation. If, for some reason, the terminus retreats slightly, the deeper water upstream of the shoal will cause an increase in iceberg calving; this will result in further retreat into deeper water, which will further increase the calving until the calving speed becomes so high that the normal processes of glacier flow cannot compensate. A rapid, irreversible retreat will result until the glacier reaches shallow water back at the head of the fjord. In contrast to the slow advance phase, the retreat phase may take only a few decades. The fastest glacier retreats observed during historical time (for instance, the opening of Glacier Bay, Alaska), as well as those inferred during the demise of the great Quaternary ice sheets, were caused by this mechanism. Information on the advance and retreat of tidewater glaciers should not be used to infer climatic change, however.

Mark F. Meier

Additional Reading
A beautifully illustrated introduction to glaciers is contained in Austin Post and E.R. LaChapelle, Glacier Ice, rev. ed. (2000). W.S.B. Paterson, The Physics of Glaciers, 3rd ed. (1994), is the standard text on glaciers and ice sheets, emphasizing process rather than description. J.T. Andrews, Glacial Systems (1975), is a compact, clear introduction to glaciers and their environment. Michael Hambrey and Jürg Alean, Glaciers (1992); and Robert P. Sharp, Living Ice (1988), are well-illustrated introductions to glaciers and their effect on the landscape. National Research Council (U.S.), Ad Hoc Committee on the Relationship between Land Ice and Sea Level, Glaciers, Ice Sheets, and Sea Level: Effect of a CO2-Induced Climatic Change (1985); and D.J. Drewry (ed.), Antarctica: Glaciological and Geophysical Folio (1983), is a large-format compendium on the Earth's largest ice mass. “Fast Glacier Flow: Ice Streams, Surging, and Tidewater Glaciers,” Journal of Geophysical Research, part B, Solid Earth and Planets, 92(9):8835–8841 (1987), is a collection of review papers and scientific contributions from the Chapman Conference on Fast Glacier Flow.Mark F. Meier

* * *


Universalium. 2010.

Игры ⚽ Поможем решить контрольную работу
Synonyms:

Look at other dictionaries:

  • Glacier Re — Glacier Reinsurance AG[1] Unternehmensform Aktiengesellschaft Gründung …   Deutsch Wikipedia

  • Glacier — Gla cier, n. [F. glacier, fr. glace ice, L. glacies.] An immense field or stream of ice, formed in the region of perpetual snow, and moving slowly down a mountain slope or valley, as in the Alps, or over an extended area, as in Greenland. [1913… …   The Collaborative International Dictionary of English

  • Glacier — Glacier, WA U.S. Census Designated Place in Washington Population (2000): 90 Housing Units (2000): 228 Land area (2000): 3.007073 sq. miles (7.788284 sq. km) Water area (2000): 0.000000 sq. miles (0.000000 sq. km) Total area (2000): 3.007073 sq.… …   StarDict's U.S. Gazetteer Places

  • Glacier, WA — U.S. Census Designated Place in Washington Population (2000): 90 Housing Units (2000): 228 Land area (2000): 3.007073 sq. miles (7.788284 sq. km) Water area (2000): 0.000000 sq. miles (0.000000 sq. km) Total area (2000): 3.007073 sq. miles… …   StarDict's U.S. Gazetteer Places

  • Glacier — Allgemeine Informationen Genre(s) Shoegazing, Independent Psychedelic Rock Gründung 2005 Website …   Deutsch Wikipedia

  • glacier — 1. (gla sié ; l r ne se lie jamais ; au pluriel, l s se lie : des gla sié z immenses) s. m. Amas considérable de glace qu on ne rencontre que dans les hautes vallées des montagnes. Les glaciers de la Savoie. Les observateurs ont reconnu que les… …   Dictionnaire de la Langue Française d'Émile Littré

  • glacier — 1744, from Fr. glacier, from Savoy dialect glacière moving mass of ice, from O.Fr. glace ice, from V.L. glacia (Cf. O.Prov. glassa, It. ghiaccia), from L. glacies (see GLACIAL (Cf. glacial)) …   Etymology dictionary

  • Glacier — (fr., spr. Glasiëh), Eisberg, Gletscher …   Pierer's Universal-Lexikon

  • Glacier — (franz., spr. glaßjĕ), Eishändler; Gletscher, Eisberg; Glacière, Eiskeller, Eisgrube …   Meyers Großes Konversations-Lexikon

  • Glacier — (frz., spr. ßĭeh), Eisberg, Gletscher; Glacière (spr. ßĭähr), Eiskeller; glacieren (spr. glaß ), mit einer glatten, glänzenden Fläche überziehen …   Kleines Konversations-Lexikon

  • glacier — [n] mountain of ice, snow berg, floe, glacial mass, iceberg, icecap, ice field, ice floe, snow slide; concept 509 …   New thesaurus

Share the article and excerpts

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