metamorphic rock

metamorphic rock
Any of a class of rocks that result from the alteration of preexisting rocks in response to changing geological conditions, including variations in temperature, pressure, and mechanical stress.

The preexisting rocks may be igneous, sedimentary, or other metamorphic rocks. The structure and mineralogy reflect the particular type of metamorphism that produced the rock and the composition of the parent rock. Metamorphic rocks are commonly classified by type of facies, predictable mineral assemblages associated with certain temperature and pressure conditions (see, e.g., granulite facies).

* * *

Introduction

      any of a class of rocks that result from the alteration of preexisting rocks in response to changing environmental conditions, such as variations in temperature, pressure, and mechanical stress, and the addition or subtraction of chemical components. The preexisting rocks may be igneous (igneous rock), sedimentary (sedimentary rock), or other metamorphic rocks.

      The word metamorphism is taken from the Greek for “change of form”; metamorphic rocks are derived from igneous or sedimentary rocksthat have altered their form (recrystallized) as a result of changes in their physical environment. Metamorphism comprises changes both in mineralogy and in the fabric of the original rock. In general, these alterations are brought about either by the intrusion of hot magma into cooler surrounding rocks (contact metamorphism) or by large-scale tectonic movements of the Earth's lithospheric plates that alter the pressure-temperature conditions of the rocks (regional metamorphism; see also plate tectonics (metamorphic rock)). Minerals within the original rock, or protolith, respond to the changing conditions by reacting with one another to produce a new mineral assemblage that is thermodynamically stable under the new pressure-temperature conditions. These reactions occur in the solid state but may be facilitated by the presence of a fluid phase lining the grain boundaries of the minerals. In contrast to the formation of igneous rocks, metamorphic rocks do not crystallize from a silicate melt, although high-temperature metamorphism can lead to partial melting of the host rock.

      Because metamorphism represents a response to changing physical conditions, those regions of the Earth's surface where dynamic processes are most active will also be regions where metamorphic processes are most intense and easily observed. The vast region of the Pacific margin, for example, with its seismic and volcanic activity, is also an area in which materials are being buried and metamorphosed intensely. In general, the margins of continents and regions of mountain building are the regions where metamorphic processes proceed with intensity. But in relatively quiet places, where sediments accumulate at slow rates, less spectacular changes also occur in response to changes in pressure and temperature conditions. Metamorphic rocks are therefore distributed throughout the geologic column.

      Because most of the Earth's mantleis solid, metamorphic processes may also occur there. Mantle rocks are seldom observed at the surface because they are too dense to rise, but occasionally a glimpse is presented by their inclusion in volcanic materials. Such rocks may represent samples from a depth of a few hundred kilometres, where pressures of about 100 kilobars (3,000,000 inches of mercury) may be operative. Experiments at high pressure have shown that few of the common minerals that occur at the surface will survive at depth within the mantle without changing to new high-density phases in which atoms are packed more closely together. Thus, the common form of SiO2, quartz, with a density of 2.65 grams per cubic centimetre, transforms to a new phase, stishovite, with a density of 4.29 grams per cubic centimetre. Such changes are of critical significance in the geophysical interpretation of the Earth's interior.

      In general, temperatures increase with depth within the Earth along curves referred to as geotherms. The specific shape of the geotherm beneath any location on Earth is a function of its corresponding local tectonic regime. Metamorphism can occur either when a rock moves from one position to another along a single geotherm or when the geotherm itself changes form. The former can take place when a rock is buried or uplifted at a rate that permits it to maintain thermal equilibrium with its surroundings; this type of metamorphism occurs beneath slowly subsiding sedimentary basins and also in the descending oceanic plate in some subduction zones. The latter process occurs either when hot magma intrudes and alters the thermal state of a stationary rock or when the rock is rapidly transported by tectonic processes (e.g.,thrust faulting or large-scale folding) into a new depth-temperature regime in, for example, areas of collision between two continents. Regardless of which process occurs, the result is that a collection of minerals that are thermodynamically stable at the initial conditions are placed under a new set of conditions at which they may or may not be stable. If they are no longer in equilibrium with one another under the new conditions, the minerals will react in such a way as to approach a new equilibrium state. This may involve a complete change in mineral assemblage or simply a shift in the compositions of the preexisting mineral phases. The resultant mineral assemblage will reflect the chemical composition of the original rock and the new pressure-temperature conditions to which the rock was subjected.

      Because protolith compositions and the pressure-temperature conditions under which they may be placed vary widely, the diversity of metamorphic rock types is large. Many of these varieties are repeatedly associated with one another in space and time, however, reflecting a uniformity of geologic processes over hundreds of millions of years. For example, the metamorphic rock associations that developed in the Appalachian Mountains of eastern North America in response to the collision between the North American and African lithospheric plates during the Paleozoic are very similar to those developed in the Alps of south-central Europe during the Mesozoic-Cenozoic collision between the European and African plates. Likewise, the metamorphic rocks exposed in the Alps are grossly similar to metamorphic rocks of the same age in the Himalayas of Asia, which formed during the continental collision between the Indian and Eurasian plates. Metamorphic rocks produced during collisions between oceanic and continental plates from different localities around the world also show striking similarities to each other (see below Regional metamorphism (metamorphic rock)) yet are markedly different from metamorphic rocks produced during continent-continent collisions. Thus, it is often possible to reconstruct tectonic events (tectonics) of the past on the basis of metamorphic rock associations currently exposed at the Earth's surface.

Metamorphic variables
      Metamorphism results from a complex interplay between physical and chemical processes that operate on a scale ranging from micrometres (e.g.,fine mineral grain sizes, thickness of intergranular fluid, diffusion distances for chemical species) to tens or hundreds of kilometres (e.g.,crustal thickness, width of collision zone between lithospheric plates, depth to subducting plate). Despite this wide range and the many processes involved in the recrystallization of sedimentary and igneous protoliths into metamorphic rocks, there are relatively few variables that effect metamorphic changes. Those of greatest importance are temperature, pressure, and the original chemical composition of the protolith; each is briefly discussed below.

      Temperatures at which metamorphism occurs range from the conditions of diagenesis (approximately 150°–200° C) up to the onset of melting. Rocks of different compositions begin to melt at different temperatures, with initial melting occurring at roughly 650°–750° C in rocks of granitic or shaley composition and approximately 900°–1,200° C in rocks of basaltic composition. Above these temperatures, metamorphic processes gradually give way to igneous processes. Hence, the temperature realm of metamorphism spans an interval of about 150°–1,100° C and is strongly dependent on the composition of the protolith.

      The temperature at any point within the Earth's crust is controlled by the local heat-flow regime, which is a composite function of heat flow (heat transfer)upward from the mantle into the crust, heat generated by radioactive decay in nearby regions of the crust, heat transported into the crust by silicate melts, and tectonic transport of hot or cold rocks at rates faster than those needed to maintain thermal equilibrium with the surrounding rocks. The temperature gradient at any location in the Earth, known as the geothermal gradient, is the increase in temperature per unit distance of depth; it is given by the tangent to the local geotherm. The magnitude of the geothermal gradient thus varies with the shape of the geotherm. In regions with high surface heat flow, such as areas of active volcanism or mantle upwelling beneath thinned continental crust, geothermal gradients of 40° to 100° C (104° to 212° F) per kilometre (0.6 mile) prevail, giving rise to high temperatures at relatively shallow levels of the crust. Within the stable interiors of old continents, geothermal gradients of 25° to 35° C per kilometre are more typical, and in zones of active subduction, where the relatively cold oceanic crust is rapidly transported to great depths, geothermal gradients range from 10° to 20° C per kilometre. These large variations in geotherms and geothermal gradients give rise to different metamorphic regimes, or combinations of pressure-temperature conditions, associated with the different tectonic provinces.

      In addition to the variation of geotherms as a function of position in the Earth, individual geotherms at a single location can vary with time. Geotherms are at steady state (i.e.,do not change with time) in tectonically quiescent areas of the Earth, such as the middle regions of large continents, and also in areas where tectonic processes like subduction have operated at similar rates over long periods. Transient geotherms, on the other hand, are generated in tectonically active regions, such as zones of continent-continent collisionor rapid uplift and erosion, in which the tectonic processes are relatively short-lived; in these areas, the temperature at a given depth in the Earth is time-dependent, and individual geotherms can have very complex shapes that with time approach smooth curves. These complex geotherms can produce wide temperature fluctuations at a given depth within the Earth; rocks metamorphosed in response to these variations may show considerable textural and chemical evidence of disequilibrium, reflecting the fact that temperatures changed at rates that were more rapid than reaction rates among the constituent minerals.

      The pressure experienced by a rock during metamorphism is due primarily to the weight of the overlying rocks (i.e.,lithostatic pressure) and is generally reported in units of bars or kilobars. The standard scientific notation for pressure is expressed in pascals or megapascals (1 pascal is equivalent to 10 bars). For typical densities of crustal rocks of two to three grams per cubic centimetre, one kilobar of lithostatic pressure is generated by a column of overlying rocks approximately 3.5 kilometres high. Typical continentalcrustal thicknesses are on the order of 30–40 kilometres but can be as great as 60–80 kilometres in mountain belts such as the Alps and Himalayas. Hence, metamorphism of continental crust occurs at pressures from a few hundred bars (adjacent to shallow-level intrusions) to 10–20 kilobars at the base of the crust. Oceanic crustis generally 6–10 kilometres in thickness, and metamorphic pressures within the oceanic crust are therefore considerably less than in continental regions. In subduction zones, however, oceanic and, more rarely, continental crust may be carried down to depths exceeding 100 kilometres, and metamorphism at very high pressures may occur. Metamorphic recrystallization also occurs in the mantle at pressures up to hundreds of kilobars.

      Changes in lithostatic pressure experienced by a rock during metamorphism are brought about by burialor upliftof the sample. Burial can occur in response either to ongoing deposition of sediments above the sample or tectonic loading brought about, for example, by thrust-faulting or large-scale folding of the region. Uplift, or more properly unroofing, takes place when overlying rocks are stripped off by erosional processes or when the overburden is tectonically thinned.

      Fluidstrapped in the pores of rocks during metamorphism exert pressure on the surrounding grains. At depths greater than a few kilometres within the Earth, the magnitude of the fluid pressure is equal to the lithostatic pressure, reflecting the fact that mineral grain boundaries recrystallize in such a way as to minimize pore space and to seal off the fluid channelways by which solutions rise from depth. At shallow depths, however, interconnected pore spaces can exist, and hence the pressure within a pore is related to the weight of an overlying column of fluid rather than rock. Because metamorphic fluids (dominantly composed of water and carbon dioxide) are less dense than rocks, the fluid pressure at these conditions is lower than the lithostatic pressure.

      Deformation (deformation and flow)of rocks during metamorphism occurs when the rock experiences an anisotropic (anisotropy) stress—i.e.,unequal pressures operating in different directions. Anisotropic stresses rarely exceed more than a few tens or hundreds of bars but have a profound influence on the textural development of metamorphic rocks (see below Textural features (metamorphic rock)and Structural features (metamorphic rock)).

Rock composition
Classification into four chemical systems
      Common metamorphic rock types have essentially the same chemical composition as what must be their equally common igneous or sedimentary precursors. Common greenschists have essentially the same compositions as basalts; marbles are like limestones; slates are similar to mudstones or shales; and many gneisses are like granodiorites. In general, then, the chemical composition of a metamorphic rock will closely reflect the primary nature of the material that has been metamorphosed. If there are significant differences, they tend to affect only the most mobile (soluble) or volatile elements; water and carbon dioxide contents can change significantly, for example.

      Despite the wide variety of igneous and sedimentary rock types that can recrystallize into metamorphic rocks, most metamorphic rocks can be described with reference to only four chemical systems: pelitic, calcareous, felsic, and mafic. Pelitic rocksare derived from mudstone (shale) protoliths and are rich in potassium (K), aluminum (Al), silicon (Si), iron (Fe), magnesium (Mg), and water (H2O), with lesser amounts of manganese (Mn), titanium (Ti), calcium (Ca), and other constituents. Calcareous rocksare formed from a variety of chemical and detrital sediments such as limestone, dolostone, or marl and are largely composed of calcium oxide (CaO), magnesium oxide (MgO), and carbon dioxide (CO2), with varying amounts of aluminum, silicon, iron, and water. Felsic rocks (felsic and mafic rocks)can be produced by metamorphism of both igneous and sedimentary protoliths (e.g.,granite and arkose, respectively) and are rich in silicon, sodium (Na), potassium, calcium, aluminum, and lesser amounts of iron and magnesium. Mafic rocks (mafic rock)derive from basalt protoliths and some volcanogenic sediments and contain an abundance of iron, magnesium, calcium, silicon, and aluminum. Ultramaficmetamorphic rocks result from the metamorphism of mantle rocks and some oceanic crust and contain dominantly magnesium, silicon, and carbon dioxide, with smaller amounts of iron, calcium, and aluminum. For the purposes of this discussion, ultramafic rocks are considered to be a subset of the mafic category.

       Common minerals of metamorphic rocks Common metamorphic minerals as a function of pressure, temperature, and protolith compositionThe particular metamorphic mineralsthat develop in each of these four rock categories are controlled above all by the protolith chemistry. The mineral calcite (CaCO3), for example, can occur only in rocks that contain sufficient quantities of calcium. The specific pressure-temperature conditions to which the rock is subjected will further influence the minerals that are produced during recrystallization; for example, at high pressures calcite will be replaced by a denser polymorph of CaCO3called aragonite. In general, increasing pressure favours denser mineral structures, whereas increasing temperature favours anhydrous and less dense mineral phases. Many of the minerals developed during metamorphism, along with their chemical compositions, are given in alphabetical order in the Table (Common minerals of metamorphic rocks). The most common metamorphic minerals that form in rocks of the four chemical categories described above are listed in the Table (Common metamorphic minerals as a function of pressure, temperature, and protolith composition)as a function of pressure and temperature. Although some minerals, such as quartz, calcite, plagioclase, and biotite, develop under a variety of conditions, other minerals are more restricted in occurrence; examples are lawsonite, which is produced primarily during high-pressure, low-temperature metamorphism of basaltic protoliths, and sillimanite, which develops during relatively high-temperature metamorphism of pelitic rocks.

Thermodynamics of metamorphic assemblages
       Common metamorphic minerals as a function of pressure, temperature, and protolith compositionDespite the large number of minerals listed in the Table (Common metamorphic minerals as a function of pressure, temperature, and protolith composition)for each of the four bulk compositions, the actual number of minerals present in an individual metamorphic rock is limited by the laws of thermodynamics. The number of mineral phases that can coexist stably in a metamorphic rock at a particular set of pressure-temperature conditions is given by the Gibbs phase rule:number of mineral phases = number of chemical components − number of degrees of freedom+ 2, where the 2 stands for the two variables of pressure and temperature. The degrees of freedom of the system are the parameters that can be independently varied without changing the mineral assemblage of the rock. For example, a rock with no degrees of freedom can only exist at a single set of pressure-temperature conditions; if either the pressure or the temperature is varied, the minerals will react with one another to change the assemblage. A rock with two degrees of freedom can undergo small changes in pressure or temperature or both without altering the assemblage. Most metamorphic rocks have mineral assemblages that reflect two or more degrees of freedom at the time the rock recrystallized. Thus, a typical pelitic rock made up of the six chemical components silica (SiO2), aluminum oxide (Al2O3), ferrous oxide (FeO), magnesium oxide (MgO), potash (K2O), and water would contain no more than six minerals; the identity of those minerals would be controlled by the pressure and temperature at which recrystallization occurred. In such a rock taken from the Earth's surface, the identity of the six minerals could be used to infer the approximate depth and temperature conditions that prevailed at the time of its recrystallization. Rocks that contain more mineral phases than would be predicted by the phase rule often preserve evidence of chemical disequilibrium in the form of reactions that did not go to completion. Careful examination of such samples under the microscope can often reveal the nature of these reactions and provide useful information on how pressure and temperature conditions changed during the burial and uplift history of the rock.

      Metamorphic rocks only rarely exhibit a chemical composition that is characteristically “metamorphic.” This statement is equivalent to saying that diffusion of materials in metamorphism is a slow process, and various chemical units do not mix on any large scale. But occasionally, particularly during contact metamorphism, diffusion may occur across a boundary of chemical dissimilarity, leading to rocks of unique composition. This process is referred to as metasomatism. If a granite is emplaced into a limestone, the contact region may be flooded with silica and other components, leading to the formation of a metasomatic rock. Often such contacts are chemically zoned. A simple example is provided by the metamorphism of magnesium-rich igneous rocks in contact with quartz-rich sediments. A zonation of the type serpentine-talc-quartz may be found such as:

      In this case the talc zone has grown by silica diffusion into the more silica-poor environment of the serpentine. Economic deposits are not uncommon in such situations—e.g.,the formation of the CaWO4(calcium tungstate) scheelitewhen tungstate in the form of WO3moves from a granite into a limestone contact. The reaction can be expressed as:

Metamorphic reactions

Reactions in a kaolinite- quartz system
      A very simple mineralogical system and its response to changing pressure and temperature provide a good illustration of what occurs in metamorphism. An uncomplicated sediment at the Earth's surface, a mixture of the clay mineral kaolinite [Al4Si4O10(OH)8] and the mineral quartz (SiO2), provides a good example. Most sediments have small crystals or grain sizes but great porosity and permeability, and the pores are filled with water. As time passes, more sediments are piled on top of the surface layer, and it becomes slowly buried. Accordingly, the pressure to which the layer is subjected increases because of the load on top, known as overburden. At the same time, the temperature will increase because of radioactive heating within the sediment and heat flow from deeper levels within the Earth.

      In the first stages of incremental burial and heating, few chemical reactions will occur in the sediment layer, but the porosity decreases, and the low-density pore water is squeezed out. This process will be nearly complete by the time the layer is buried by five kilometres of overburden. There will be some increase in the size of crystals; small crystals with a large surface area are more soluble and less stable than large crystals, and throughout metamorphic processes there is a tendency for crystals to grow in size with time, particularly if the temperature is rising.

      Eventually, when the rock is buried to a depth at which temperatures of about 300° C obtain, a chemical reaction sets in, and the kaolinite and quartz are transformed to pyrophyllite and water:

      The exact temperature at which this occurs depends on the fluid pressure in the system, but in general the fluid and rock-load pressures tend to be rather similar during such reactions. The water virtually fights its way out by lifting the rocks. Thus, the first chemical reaction is a dehydration reaction leading to the formation of a new hydrate. The water released is itself a solvent for silicates and promotes the crystallization of the product phases.

      If heating and burial continue, another dehydration sets in at about 400° C, in which the pyrophyllite is transformed to andalusite, quartz, and water:

      After the water has escaped, the rock becomes virtually anhydrous, containing only traces of fluid in small inclusions in the product crystals and along grain boundaries. Both of these dehydration reactions tend to be fast, because water, a good silicate solvent, is present.

      Although the mineral andalusite is indicated as the first product of dehydration of pyrophyllite, there are three minerals with the chemical composition Al2SiO5. Each has unique crystal structures, and each is stable under definite pressure-temperature conditions. Such differing forms with identical composition are called polymorphs (polymorphism). If pyrophyllite is dehydrated under high-pressure conditions, the polymorph of Al2SiO5formed would be the mineral kyanite (the most dense polymorph). With continued heating, the original andalusite or kyanite will invert to sillimanite, the highest-temperature Al2SiO5polymorph:

      The kinetics of these polymorphic transformations are sufficiently sluggish, however, that kyanite or andalusite may persist well into the stability field of sillimanite.

Reactions of other mineral systems
      Owing to the very simple bulk composition of the protolith in this example (a subset of the pelitic system containing only SiO2-Al2O3-H2O), no other mineralogical changes will occur with continued heating or burial. The original sediment composed of kaolinite, quartz, and water will thus have been metamorphosed into a rock composed of sillimanite and quartz, and perhaps some metastable andalusite or kyanite, depending on the details of the burial and heating history. In the case of a more typical pelite containing the additional chemical components potash, ferrous oxide, and magnesium oxide, the reaction history would be correspondingly more complex. A typical shalethat undergoes burial and heating in response to continent-continent collision would develop the minerals muscovite, chlorite, biotite, garnet, staurolite, kyanite, sillimanite, and alkali feldspar, in approximately that order, before beginning to melt at about 700° C. Each of these minerals appears in response to a chemical reaction similar to those presented above. Most of these reactions are dehydration reactions, and the shale thus loses water progressively throughout the entire metamorphic event. As discussed above, the total number of minerals present in the rock is controlled by the Gibbs phase rule, and the addition of new minerals generally results from the breakdown of old minerals. For example, the following reaction,

      occurs at temperatures of about 500°–550° C and typically consumes all the preexisting chlorite in the rock, introduces the new mineral staurolite, and adds more biotite and quartz to the biotite and quartz generated by earlier reactions. Some garnet and muscovite usually remain after the reaction, although examination of the sample under the microscope would probably reveal partial corrosion (wearing away due to chemical reactions) of the garnets resulting from their consumption.

Isograds
      Reactions that introduce new minerals in rocks of a specific bulk composition are referred to as mineral appearance isograds. Isograds can be mapped in the field as lines across which the metamorphic mineral assemblage changes. Caution must be exercised to note the approximate bulk composition of the rocks throughout the map area, however, because the same mineral can develop at quite different sets of pressure-temperature conditions in rocks of dissimilar chemical composition. For example, garnet generally appears at a lower temperature in pelitic rocks than it does in mafic rocks; hence, the garnet isograd in pelitic rocks will not be the same map line as that in metabasaltic (i.e.,metamorphosed basalt) compositions. Isograd patterns are discussed further in Structural features (metamorphic rock)below.

Principal types
      Metamorphic reactions can be classified into two types that show different degrees of sensitivity to temperature and pressure changes: net-transfer reactionsand exchange reactions. Net-transfer reactions involve the breakdown of preexisting mineral phases and corresponding nucleationand growth of new phases. (Nucleation is the process in which a crystal begins to grow from one or more points, or nuclei.) They can be either solid-solid reactions(mineral A + mineral B = mineral C + mineral D) or devolatilization reactions(hydrous mineral A = anhydrous mineral B + water), but in either case they require significant breaking of bonds and reorganization of material in the rock. They may depend most strongly on either temperature or pressure changes. In general, devolatilization reactions are temperature-sensitive, reflecting the large increase in entropy (disorder) that accompanies release of structurally bound hydroxyl groups (OH) from minerals to produce molecular water. Net-transfer reactions that involve a significant change in density of the participating mineral phases are typically more sensitive to changes in pressure than in temperature. An example is the transformation of albite (NaAlSi3O8) to the sodic pyroxene jadeite (NaAlSi2O6) plus quartz (SiO2); albite and quartz have similar densities of about 2.6 grams per cubic centimetre, whereas jadeite has a density of 3.3 grams per cubic centimetre. The increased density reflects the closer packing of atoms in the jadeite structure. Not surprisingly, the denser phase jadeite is produced during subduction zone (high-pressure) metamorphism. Net-transfer reactions always involve a change in mineral assemblage, and textural evidence of the reaction often remains in the sample (see below Textural features (metamorphic rock)); isograd reactions are invariably net-transfer reactions.

      In contrast to net-transfer reactions, exchange reactions involve redistribution of atoms between preexisting phases rather than nucleation and growth of new phases. The reactions result simply in compositional changes of minerals already present in the rock and do not modify the mineral assemblage. For example, the reaction

      results in redistribution of iron and magnesium between garnet and biotite but creates no new phases. This reaction is limited by the rates at which iron and magnesium can diffuse through the garnet and biotite structures. Because diffusion processes are strongly controlled by temperature but are nearly unaffected by pressure, exchange reactions are typically sensitive to changes only in metamorphic temperature. Exchange reactions leave no textural record in the sample and can be determined only by detailed microanalysis of the constituent mineral phases. The compositions of minerals as controlled by exchange reactions can provide a useful record of the temperature history of a metamorphic sample.

      The types of reactions cited here are typical of all metamorphic changes. Gases are lost (hydrous minerals lose water, carbonates lose carbon dioxide), and mineral phases undergo polymorphic or other structural changes; low-volume, dense mineral species are formed by high pressures, and less dense phases are favoured by high temperatures. Considering the immense chemical and mineralogical complexity of the Earth's crust, it is clear that the number of possible reactions is vast. In any given complex column of crustal materials some chemical reaction is likely for almost any incremental change in pressure and temperature. This is a fact of immense importance in unraveling the history and mechanics of the Earth, for such changes constitute a vital record and are the primary reason for the study of metamorphic rocks.

Retrograde metamorphism
      In general, the changes in mineral assemblage and mineral composition that occur during burial and heating are referred to as prograde metamorphism, whereas those that occur during uplift and cooling of a rock represent retrograde metamorphism. If thermodynamic equilibrium were always maintained, one might expect all the reactions that occur during prograde metamorphism to be reversed during subsequent uplift of the rocks and reexposure at the Earth's surface; in this case, metamorphic rocks would never be seen in outcrop. Two factors mitigate against complete retrogression of metamorphic rocks, however, during their return to the Earth's surface. First is the efficient removal of the water and carbon dioxide released during prograde devolatilization reactions by upward migration of the fluid along grain boundaries and through fractures. Because almost all the water released during heating by reactions such as

      is removed from the site of reaction, the reaction cannot be reversed during cooling unless water is subsequently added to the rock. Thus, garnet can be preserved at the Earth's surface even though it is thermodynamically unstable at such low temperatures and pressures. The second reason that metamorphic reactions do not typically operate in reverse during cooling is that reaction rates are increased by rising temperatures. During cooling, reaction kinetics become sluggish, and metastable mineral assemblages and compositions can be preserved well outside their normal stability fields. Thus, prograde reactions are generally more efficient than retrograde reactions, and metamorphic assemblages indicative of even extremely high temperatures or pressures or both are found exposed throughout the world. It is common, however, to find at least some signs of retrogression in most metamorphic rocks. For example, garnets are often rimmed by small amounts of chlorite and quartz, indicating that limited quantities of water were available for the reverse of the reaction given above to proceed during cooling. Retrograde features such as these reaction rims can be mapped to yield information on pathways of fluid migration through the rocks during uplift and cooling. In other rocks, such as high-temperature gneisses, mineral compositions often reflect temperatures too low to be in equilibrium with the preserved mineral assemblage; in these samples, it is clear that certain exchange reactions operated in a retrograde sense even when the net-transfer reactions were frozen in during prograde metamorphism.

Textural features
      The fabricof a metamorphic rock results from the combined effects of mineral reactions and deformation throughout the metamorphic event and the subsequent return of the rock to the terrestrial surface. The study of metamorphic fabrics in outcrop and under the microscope has become a highly specialized subject aimed at revealing the nature and direction of the forces acting during dynamic processes within the Earth. Much of this work is an outgrowth of a classic investigation conducted in 1930 by the Austrian geologist Bruno Sander, coupled with more recent experimental work on the pressure-temperature stabilities of metamorphic minerals and their responses to deformation.

      Observations show that pressure is only rarely hydrostatic (equal in all directions) at any point within the Earth's crust. In real cases, consequently, anisotropic stresses operate that may lead to flow or fracture of materials. Such occurrences produce certain characteristic fabrics or structures in metamorphic rocks that may be observed at the scale of the orientation of small crystals in a rock or as a pattern of folds in a mountain range. One of the principal characteristics of most metamorphic rocks is that the arrangement of crystals is not isotropic, or random, but that there is a strong preferred orientation related to the direction of stress components of pressure. Such preferred orientation of crystals and mineral grains is perhaps the most striking difference between metamorphic rocks and other rock types.

Major features
      The most obvious features of metamorphic rocks are certain planar features that are often termed s-surfaces. The simplest planar features may be primary bedding (stratification)(akin to the layering in sedimentary rocks). As the rock crystallizes or recrystallizes under directed pressure, new crystals may grow in some preferred direction, sometimes subparallel to the primary bedding but often at new angles defining new planar structures. At the same time, folding (fold)of layers may occur, leading to folds with amplitudes on scales of kilometres or millimetres. Fabric symmetry may be represented by the nature of deformed fossils, pebbles in a conglomerate, or any objects with a known shape prior to deformation.

      A few terms that commonly are used to describe several types of preferred orientation in metamorphic rocks include foliation, a general term describing any type of s-surface, bedding, or crystal orientation; slaty cleavage, a planar structure leading to facile cleavage that is normally caused by the preferred orientation of mica crystals; schistosity, a term used to describe repetitive and pronounced foliation of the type that is present in schists; and lineation, which is any linear structure, such as the axis of the fold, grooves on a fault plane, or the direction of stretching of pebbles.

      The various mineral phases of a metamorphic rock have different physical properties and symmetries. When a rock is subjected to recrystallization in a stress field, different substances will behave differently according to such physical properties and symmetries. Some minerals always tend to grow in better-formed crystals than others; rates of nucleation may differ, and this can lead to different patterns of growth of crystals—there may be a few large crystals or a mass of small crystals. Minerals can be arranged in order of their tendency to form crystals showing planar surfaces—namely, magnetite, garnet, epidote, mica, calcite, quartz, and feldspar. Minerals that have a tendency to form large single crystals (e.g.,garnet) are termed porphyroblasts.

      Porphyroblastic crystals may grow before, during, or after an episode of deformation (pre-, syn-, and postkinematic growth, respectively); the relative timing of mineral growth and deformation can often be determined by examining the sample under a microscope. Prekinematic porphyroblasts may be fractured by subsequent deformation; the orientation of the fractures and any offset of the grains along them provide information on the directed stresses responsible for the deformation. Prekinematic grains may also be surrounded by pressure shadows produced by minerals such as quartz that dissolve in zones of maximum compressive stress and reprecipitate in zones of lesser compressive stress adjacent to the rigid porphyroblasts. The texture of the shadows is different from that of the host rock. Samples exhibiting asymmetric pressure shadows around porphyroblasts can yield information on the orientation of shear stresses during deformation. A spectacular example of synkinematic prophyroblast growth is provided by the so-called snowball garnets, which have spiral trails of inclusions that indicate rotation of either the garnet or the matrix foliation during garnet growth. Postkinematic porphyroblasts typically overgrow all previous fabrics in the rock and may contain trails of mineral inclusions that define microfolds or an earlier schistosity.

      In some samples, it is possible to use the compositions of the porphyroblasts to calculate the depth and temperature conditions at which they grew and thereby constrain the conditions at which deformation occurred. Studies of this sort add immeasurably to the understanding of crustal rheologies and the response of rocks to large-scale orogenic events. Because a particular metamorphic event may be accompanied by either several isolated episodes of deformation or a single continuum of deformation, there may be many fabric generations recorded in one sample; individual minerals may be postkinematic with respect to the earliest deformation but prekinematic relative to younger deformation in the same rock. Thus, the study of porphyroblast fabrics in metamorphic rocks can be complex but has the potential to yield important information on the structural history of metamorphic regions.

 Because changes in pressure and temperature often occur at faster rates than those of mineral reaction and recrystallization, metamorphic rocks may display fabrics that result from incomplete reactions. Such disequilibrium features provide a wealth of information on the reaction history of the sample and, by comparison with experimental studies of mineral stabilities, can also constrain the quantitative pressure-temperature history of the rock during metamorphism. An example of a reaction texture is shown in Figure 1—>, in which a corroded garnet is surrounded by a corona(reaction rim) of the mineral cordierite; other minerals present in the matrix include sillimanite, quartz, biotite, and alkali feldspar. The sample does not contain garnet in contact with sillimanite or quartz. These textural features suggest the following reaction relationship between garnet, sillimanite, quartz, and cordierite:

      This reaction has been shown experimentally to occur at temperatures of approximately 725° ± 50° C and to be very sensitive to pressure, with cordierite occurring under low-pressure conditions. The textural evidence that preexisting garnet was partially replaced by cordierite thus implies that the rock underwent decompression while still at high temperatures and that the decompression occurred too rapidly for the rock to recrystallize completely (i.e.,for garnet to be totally replaced by cordierite).

Lamination
      There is also a tendency for many types of metamorphic rocks to become laminated, and the separate laminae may have distinct chemical compositions. A macroscopically rather homogeneous sediment may prove to be inhomogeneous on a minute scale. When graywackes are metamorphosed within the greenschist facies, for example, laminae rich in quartz and feldspar alternate with others rich in epidote, chlorite, and muscovite. The precise causes of this process are not well known, but it may result from a combination of extensive deformation accompanied by recrystallization. In a sense, it is a type of flow unmixing. It is important to recognize that this type of structure need have no relation to original bedding in the unmetamorphosed sediments.

Structural features
      Metamorphic rocks are often intimately related to large-scale (kilometres of tens of kilometres) structural features of the Earth. Such features include folds, nappes, and faults with a wide variety of geometries. In many cases, the correlation of metamorphic isograds and their position in the structure implies a genetic relationship between the two. For example, one of the major structural features in the Himalayan mountain belt is the Main Central Thrust, a thrust faultthat runs for hundreds of kilometres from east to west and was responsible for the transportation of rocks belonging to the Eurasian Plate southward over those of the Indian Plate. Along much of the length of this fault, the metamorphic rocks in the hanging wall (located above the fault) display a pattern of inverted isograds; i.e.,the rocks that reached the highest temperatures of metamorphism overlie rocks that record lower temperatures, implying that metamorphic temperatures decreased with depth to the fault. Several explanations have been proposed to account for this anomalous distribution of temperature with depth. One model suggests the fault transported hot Asian rocks over cooler Indian rocks, which caused cooling of the Asian rocks in the vicinity of the fault. Another model proposes that fluids circulating along the fault zone caused retrograde metamorphism and thus reset the rocks located nearest to the fault to lower temperatures. Although neither of these models provides an adequate explanation for the entire length of the Main Central Thrust, they both emphasize the significant control that structural features can exert on the development of metamorphic rocks.

      Metamorphism associated with nappes (nappe)(large recumbent folds) in the Alps and the Appalachians provides strong evidence that the tectonic transport of rocks typically occurs at rates faster than those of thermal equilibration—in other words, that the nappes can transport hot rocks for large distances without significant cooling. Nappe formation is a major process of crustal thickening during continent-continent collision; emplacement of the nappes results in burial and heating of the underlying rocks. Isograddistributions associated with nappe structures can be either normal or inverted, depending on the relative rates of nappe emplacement and heat transfer.

      Isograd maps can provide information on the relative timing of structural and metamorphic events in much the same way that fabric studies constrain the relative timing of deformation episodes and prophyroblast growth. For example, isograd patterns that are cut by faults clearly indicate that metamorphism predated fault displacement, whereas isograd sequences that overprint structural discontinuities imply the reverse. Isograds that parallel major structures suggest some cause-and-effect relationship between the structural and metamorphic development of the region. Since the 1980s, metamorphic petrologists and structural geologists have increasingly worked together to correlate metamorphic and tectonic events and thereby increase understanding of crustal dynamics in tectonically active regions of the Earth.

Metamorphic facies
      Metamorphic petrologists studying contact metamorphism early in the 20th century introduced the idea of metamorphic facies to correlate metamorphic events. The concept was first defined in 1914 by a Finnish petrologist, Pentti Eelis Eskola, as any rock of a metamorphic formation that has attained chemical equilibrium through metamorphism at constant temperature and pressure conditions, with its mineral composition controlled only by the chemical composition. In current usage, a metamorphic facies is a set of metamorphic mineral assemblages, repeatedly associated in space and time, such that there is a constant and therefore predictable relation between mineral composition and chemical composition.

      The facies concept is more or less observation-based. In a single outcrop, for instance, layers of different chemical composition will display different mineral assemblages despite having all experienced the same pressure and temperature history. A pelitic layer might contain the assemblage garnet + chlorite + biotite + muscovite + quartz, whereas a basaltic horizon a few centimetres away would contain the assemblage chlorite + actinolite + albite. Both of these rocks belong to the same facies, meaning that, in another region, a geologist who observed the assemblage chlorite + actinolite + albite in a metabasalt could predict that associated pelitic rocks would contain the garnet + chlorite + biotite + muscovite + quartz assemblage.

      Experimental work on the relative stabilities of metamorphic minerals and assemblages has permitted correlation of the empirically derived facies with quantitative pressure and temperature conditions. The names of metamorphic facies in common usage are derived from the behaviour of a rock of basaltic (basalt)bulk composition during metamorphism at various sets of pressure-temperature conditions. For example, a basalt metamorphosed during subduction to high pressures at low temperatures recrystallizes into a rock containing glaucophane, lawsonite, and albite; glaucophane is a sodic amphibole that is blue to black in hand sample and lavender to blue under the microscope. Because of their distinctive bluish coloration, such samples are called blueschists. The same rock type metamorphosed at more moderate pressures and temperatures in the range of 400°–500° C would contain abundant chlorite and actinolite, minerals that are green both in hand sample and under the microscope, and would be referred to as a greenschist (greenschist facies). At somewhat higher temperatures, the rock would become an amphibolite, reflecting a mineralogy composed predominantly of the amphibole hornblende along with plagioclase and perhaps some garnet. At still higher temperatures, a metabasalt recrystallizes into a rock containing hypersthene, diopside, and plagioclase; in general, these minerals form relatively equant crystals and hence do not develop a preferred orientation. The granular texture of these rocks has resulted in the name granulitefor a high-temperature metabasalt. A pelitic or calcareous rock will develop very different mineral assemblages from a metabasalt, yet the same facies names apply. Thus, one can refer to a greenschist facies pelitic schist, an amphibolite facies calcsilicate rock, or a granulite facies garnet gneiss.

      The boundaries between the different facies are regions of pressure and temperature in which chemical reactions occur that would significantly alter the mineralogy of a rock of basaltic bulk composition. For example, the boundary between the greenschist and amphibolite facies marks a transition from amphibole of actinolitic composition to hornblende and of a sodic plagioclase into a more calcic plagioclase. The reactions that bring about these transformations depend on the specific composition of the rock.

Facies series
      Different types of tectonic processes produce different associations of metamorphic facies in the field. For example, regions associated with subductionof oceanic material beneath either oceanic or continental crust are characterized by blueschist, greenschist, and eclogite facies rocks, whereas areas thought to reflect continent-continent collisionare more typically distinguished by greenschist and amphibolite facies rocks. Still other regions, usually containing an abundance of intrusive igneous material, show associations of low-pressure greenschist, amphibolite, and granulite facies rocks. These observations led a Japanese petrologist, Akiho Miyashiro, working in the 1960s and '70s, to develop the concept of baric types, or metamorphic facies series. Miyashiro described the three facies associations given above as high-pressure, medium-pressure, and low-pressure facies series, respectively, and correlated the development of these characteristic series with the shape of the geotherm in different tectonic settings. Subsequent thermal modeling studies have shown that metamorphism generally occurs in response to tectonically induced perturbation of geotherms rather than along steady-state geotherms and, hence, that the facies series do not record metamorphic geotherms. Nonetheless, the concept of metamorphic facies series is a useful one in that it emphasizes the strong genetic relationship between metamorphic style and tectonic setting.

Pressure-temperature-time paths
 Interaction between metamorphic petrologists and geophysicists in the 1980s led to the realization that each metamorphic rock follows its own unique path through pressure- (depth-) temperature space during metamorphism and that these paths bear little or no resemblance to steady-state geotherms. The specific shape of a pressure-temperature-time (P-T-t) path depends on the tectonic setting in which the rock was metamorphosed, which in turn controls the relative rates at which burial or uplift and heating or cooling occur. For example, a rock that is uplifted rapidly from depth, perhaps in response to extensional faulting (that caused by the stretching of the Earth's crust), may transport heat with it to near-surface depths. Its P-T-t path would show a phase of nearly isothermal decompression (uplift at approximately constant temperature), reflecting the fact that uplift rates were more rapid than those of heat transfer. In contrast, if the same rock remained at depth for a long period of time and then experienced very slow uplift, its P-T-t path would show cooling during uplift or even a phase of isobaric (constant-pressure) cooling. Rocks belonging to medium-pressure facies series generally follow P-T-t paths that are clockwise loops on a pressure-temperature diagram, reflecting rapid burial during a collisional event followed by heating and relatively rapid uplift. In contrast, low-pressure facies series rocks may follow counterclockwise P-T-t paths in response to rapid heating of the crust due to magma intrusion prior to uplift (Figure 2—>). P-T-t paths followed by rocks of a high-pressure facies series are less predictable and depend strongly on the mechanism by which the rocks are transferred from the subducting slab to the overlying continental crust. In general, the mineral assemblage preserved in a metamorphic rock is frozen at the highest temperature experienced during metamorphism (see above Retrograde metamorphism (metamorphic rock)), and thus the facies and facies series to which the rock would be assigned reflect only a single point on its P-T-t path.

      One of the principal goals of much of the work that is done on metamorphic rocks is the reconstruction of the P-T-t paths followed by rocks presently exposed at the Earth's surface. Because these paths are so strongly linked to dynamic processes, their reconstruction provides a means by which tectonic processes operative in the geologic past may be understood. Owing to the continuous recrystallization of rocks that occurs during progressive metamorphism, much of the early record of metamorphic changes within a sample is eradicated by later events. It is, therefore, not possible to determine the entire P-T-t path followed by an individual sample, but often enough disequilibrium features are preserved to permit reconstruction of a few thousand bars and a couple of hundred degrees of the path; such a portion may represent anywhere from a few million to a hundred million years of Earth history, as revealed by geochronologic determinations involving different minerals or fabric generations in the sample. Techniques for determining the pressure-temperature history of a metamorphic rock include using compositions of coexisting minerals to calculate pressures and temperatures of equilibration (geobarometry and geothermometry, respectively), comparing the mineral assemblage to experimentally determined stability fields for the phases, utilizing mineral inclusions enclosed within porphyroblasts to constrain assemblages present in the early history of the sample, and making use of the densities of small inclusions of fluids trapped within the minerals to determine possible pressures and temperatures experienced at different stages in the burial and uplift history.

Origin of metamorphic rocks: types of metamorphism
      It is convenient to distinguish several general types of metamorphism in order to simplify the description of the various metamorphic phenomena. Recognized here are hydrothermal, dynamic, contact, and regional metamorphism, each of which will be described in turn.

Hydrothermal metamorphism
      Changes that occur in rocks near the surface, where there is intense activity of hot water, are categorized as hydrothermal metamorphism. Such areas include Yellowstone National Park in the northwestern United States, the Salton Sea in California, and Wairakei in New Zealand. It is now generally recognized that the circulating groundwaters that often become heated by their proximity to igneous materials produce the metamorphism. Migration of chemical elements, vein formation, and other kinds of mineral concentration may be extreme on account of the large volumes of water circulated.

Dynamic metamorphism
      When directed pressure or stress is the dominant agent of metamorphism, it is termed dynamic; other terms are dislocation, kinematic, and mechanical metamorphism. Mineralogical changes occurring on a fault plane provide an obvious example. In some such cases, the action may simply be a grinding up of existing grains or realignment of minerals that have non-equant crystals. If the action is intense, friction may even lead to melting.

Contact metamorphism
      Whenever silicate melts (magmas (magma), from which igneous rocks crystallize within the Earth) invade the crust at any level, they perturb the normal thermal regime and cause a heat increase in the vicinity. If a mass of basaltic liquid ascending from the upper mantle is trapped in the crust and crystallizes there, it will heat the surrounding area; the amount of heating and its duration will be a direct function of the mass and shape of the igneous material. Contact-metamorphic phenomena thus occur in the vicinity of hot igneous materials and at any depth. Under such circumstances, pressure and temperature are not simply correlated. Thermal gradients are often steep unless the igneous mass is extremely large. Contact aureoles—the surrounding zones of rock that become altered or metamorphosed—vary in thickness from several centimetres (around tabular bodies such as dikes and thin sills) to several kilometres (around large granitic intrusions). The contact metamorphic rocks of the aureole zone often lack any obvious schistosity or foliation.

      The facies associated with contact metamorphism include the sanidinite, pyroxenite-hornfels, hornblende-hornfels, and albite-epidote-hornfels facies.

      Rocks of the sanidinite facies are represented by small fragments of aureole materials that have often been totally immersed in silicate liquids or by the aureole rocks surrounding volcanic pipes. Very high temperatures are attained, often at very low pressures. The dominant feature of the mineralogy of this facies is an almost complete lack of minerals containing water or carbon dioxide. Many of the minerals show similarity to those of igneous rocks themselves. If the duration of heating is short, adjustment to the imposed temperature is often imperfect.

      Pelitic rocks (high in aluminum oxide) contain minerals such as mullite, sillimanite, sanidine, cordierite, spinel, hypersthene, anorthite, tridymite, and even glass. One of the classic localities of such rocks is the island of Mull, off the west coast of Scotland, but these rocks can be found in most regions of volcanism.

      Calcareous rocks (originally impure limestones or dolomites) tend to lose nearly all their carbon dioxide, but pure calcite may survive. Typical metamorphic minerals are quartz, wollastonite, anorthite, diopside, periclase, and in some places (the classic is Scawt Hill in Northern Ireland) an array of complex calcium silicates such as spurrite, larnite, rankinite, melilite, merwinite, and monticellite. These minerals result from the addition of varying amounts of silica to impure mixtures of calcite and dolomite. In a general way the minerals of this facies are reminiscent of those of industrial slags.

Pyroxene-hornfels facies
      Rocks of the pyroxene-hornfels facies are characteristically formed near larger granitic or gabbroic bodies at depths of a few kilometres or at pressures of a few hundred bars. The mineral assemblages are again largely anhydrous, but, unlike the sanidinite facies, the minerals reflect distinctly lower temperatures. One of the classic descriptions of such rocks is from the Oslo district of Norway.

      In pelitic rocks, minerals such as quartz, orthoclase, andalusite, sillimanite, cordierite, orthopyroxene, and plagioclase occur. Sometimes the hydrate biotite is developed. In calcareous rocks the minerals found include plagioclase, diopside, grossularite, vesuvianite, wollastonite, and sometimes the more complex calcium silicates monticellite, melilite, spurrite, tilleyite, and clinohumite.

Hornblende-hornfels facies
      A generally deeper level of contact metamorphism at pressures of a few kilobars is represented by the hornblende-hornfels facies. Hydrated phases become stable, and the transition to regional metamorphism becomes apparent. Because of the generally greater depth, this type of aureole is often superposed on a metamorphism at more normal pressure-temperature conditions, and the rocks may appear schistose and exhibit new thermally generated minerals on a preexisting assemblage. This type of metamorphism develops the classic “spotted” texture in which new porphyroblasts grow in slates and phyllites of a previous episode of metamorphism. Typically, such rocks are developed near most of the world's large granite batholiths.

      Typical minerals of pelitic assemblages include quartz, muscovite, biotite, andalusite, sillimanite, cordierite, plagioclase, microcline, and staurolite. Calcareous assemblages include calcite, quartz, diopside, grossular, plagioclase, wollastonite, brucite, talc, forsterite, tremolite, and clinozoisite. Basaltic compositions include plagioclase, hornblende, diopside, quartz, biotite, and almandine garnet.

      When rather pure limestoneand dolomitecome into direct contact with granitic rocks, elements such as silicon, iron, magnesium, and aluminum diffuse into the limestone, forming spectacular rocks termed skarns (skarn). These rocks often consist of large garnet crystals (grossular) with green diopside and vesuvianite or epidote.

Albite-epidote-hornfels facies
      Rocks of the albite-epidote-hornfels facies are characteristically found as the outer zones of contact aureoles where the thermal episode fades out and the rocks pass into their regional grade of metamorphism. The mineral assemblages are quite similar to those found in regional greenschist-facies metamorphism, except for the presence of low-pressure phases such as andalusite. Characteristic minerals include quartz, muscovite, biotite, chlorite, andalusite, actinolite, calcite, dolomite, albite, and epidote.

Regional metamorphism
      Regional metamorphism is associated with the major events of Earth dynamics, and the vast majority of metamorphic rocks are so produced. They are the rocks involved in the cyclic processes of erosion, sedimentation, burial, metamorphism, and mountain building, events that are all related to major convective processes in the Earth's mantle.

      Most regionally metamorphosed rocks develop primarily in response to continent-continent collision and to collision between oceanic and continental plates. As a result, young metamorphic belts aligned roughly parallel to the present-day continental margins (e.g.,the Pacific margin) as well as older metamorphic belts are used to infer the geometries of the continental margins at earlier periods in Earth history. Most of the world's mountain beltsare at least partially composed of regionally metamorphosed rocks, with spectacular examples provided by the Alps, the Himalayas, the northern Appalachians, and the Highlands of Scotland. Although the processes that formed each of these mountain belts are broadly similar, in almost all such crustal events at different times and places, there is uniqueness as well as conformity to a general pattern. Metamorphic events in the Alps, the Urals, and the Himalayas all show specific differences: to unravel such differences and their significance is one of the major tasks of metamorphic petrology.

      In areas of collision between oceanicand continentallithospheric plates such as the circum-Pacific region, the denser oceanic plate is subducted (carried into the Earth's mantle) beneath the more buoyant continental lithosphere. Rapid subductionof the cool oceanic lithosphere perturbs the thermal regime in such a way that high pressures can be obtained at relatively low temperatures, thereby generating blueschists and eclogites (high-pressure facies series) from ocean-floor basalts transported down the subduction zone. Continued subduction of these rocks to great depth may eventually result in either (1) rising temperatures and partial melting of subducted rocks or (2) the melting of hydrated peridotite created by fluids released from metamorphic reactions in the subduction zone that rise into the overlying mantle wedge. These melts contribute to the formation of the volcanoes that overlie subduction zones in areas such as the Andes of South America, Japan, and the Aleutian Islands. Upward migration of subduction-related magmas also contributes to the development of paired metamorphic belts, in which high-pressure, low-temperature metamorphic rocks are flanked on the continental side by a parallel belt of low-pressure, high-temperature rocks. The latter rocks are thought to reflect perturbation of the crustal thermal regime by the passage of silicate melts generated above the subducting slab. Continued intrusion of magma over a period of time would cause an increase in crustal temperatures at relatively shallow depths and produce the high-temperature rocks adjacent to the high-pressure rocks generated in the subduction zone. Well-developed paired metamorphic belts are exposed in Japan, California, the Alps, and New Zealand.

      Data obtained from deep earthquakes in subduction zones indicate that a descending slab of oceanic lithosphere can remain intact to depths of several hundred kilometres before undergoing complete melting or fragmentation or both and being incorporated into the surrounding mantle. Clearly, the blueschists and eclogites exposed in orogenic belts around the world did not undergo such a process and were instead returned to the Earth's surface. Most of the high-pressure rocks that have been studied from Japan, California, New Caledonia, the Alps, and Scandinavia record maximum pressures of 10–20 kilobars, corresponding to subduction to depths of approximately 35–70 kilometres. A few samples have been discovered in Norway, the Alps, and China that contain the mineral coesite, a high-pressure polymorph of quartz. Experimental studies on the stability of coesite imply minimum pressures of 30 kilobars for these rocks, indicating burial or subduction to depths of approximately 100 kilometres. This is termed ultra-high pressure metamorphism (UHPM). These pressures are particularly noteworthy in that they are recorded in rocks derived from sedimentary rather than basaltic protoliths. Because of the low density, and hence greater buoyancy, of sediments relative to basalts, many geologists have argued that sediment subduction must be a rather limited process; the coesite-bearing metapelites (metamorphosed pelites) provide important evidence that sediment subduction can and does occur under certain circumstances.

      The processes by which rocks that have been partially subducted are returned to the surface are not well understood. Models have been proposed to account for uplift and exposure of these high-pressure, high-density rocks; they include scraping material from the subducting plate against the overlying crustal lithosphere, upward flow of material in response to forced convection above the subducted slab, and removal of overlying thickened crust by low-angle extensional faulting. Testing these models requires considerable petrologic and structural work in areas where high-pressure rocks are exposed.

      Most of the high-pressure rocks that are currently displayed in metamorphic belts around the world were metamorphosed in Mesozoic or Cenozoic time (e.g.,the circum-Pacific belt, the Alps, the Greek Cyclades, and the Cordillera Betica in Spain). Older high-pressure rocks are known from only a few isolated occurrences in, for example, Wales, Bavaria, the ële de Groix off the coast of Brittany, and the Norwegian Caledonides (on the west coast of Norway). The general absence of high-pressure samples in the early rock record raises a number of interesting questions concerning Earth history. Some geologists have argued that the lack of well-developed Precambrian and Paleozoic high-pressure belts indicates that plate tectonic processes have changed significantly throughout geologic time. Specifically, they claim that greater heat production in Archean time (about 3.8 to 2.5 billion years ago) would have produced hotter crustal geotherms, resulting in thin, hot lithospheric plates whose mechanical behaviour may have been quite different from that of the present-day plates and hence may not have permitted formation of subduction zones. The increasing abundance of subduction-related metamorphic rocks with decreasing age in the rock record would thus reflect the gradual onset of plate tectonics as operative today. Others argue that the rock record is biased owing to preferential erosion or thermal overprinting (development of a new mineralogy that may obliterate the original one) of old blueschists and eclogites. Thermal modeling studies suggest that blueschists will generally undergo heating and be converted to greenschist assemblages if exposure at the Earth's surface does not occur within 100 million to 200 million years after high-pressure metamorphism. Early exposure at the surface also increases the chances for removal by erosion, however, resulting in a low probability for preserving blueschists greater than 100 million to 200 million years old. Geologists favouring generation of blueschists throughout Earth history but only selective preservation of these rocks also point to crustal rocks more than 2.5 billion years old that record metamorphism at depths of 25–40 kilometres; these medium-pressure facies series rocks imply that crustal thicknesses in the early Earth were similar to those of the present day and thus that modern plate tectonic processes may have operated from the early Precambrian to the present. This debate, though unresolved, emphasizes the substantial knowledge of the thermal structure of the Earth and plate tectonic processes that can be obtained from the study of metamorphic rocks.

      Depending on the original geometry of the Earth's lithospheric plates, subduction of oceanic crust beneath continental lithosphere may result in complete consumption of an ocean basin and subsequent collision between two continents. Collisions of this type have a long and complex history that may include initial formation of a paired metamorphic belt followed by extreme crustal thickening in response to the actual collision of the continents. The overthickened crust produced by the collision event will be gravitationally unstable and will undergo subsequent rapid erosion and possibly extensional faulting in order to return to a normal crustal thickness. Rocks metamorphosed in the early stages of collision may belong to a high-pressure facies series, reflecting the final stages of subduction of oceanic lithosphere, whereas the younger facies more typically belong to medium-pressure facies series. Metamorphic rocks exposed in former collision zones may thus have followed a variety of pressure-temperature-time paths, but paths showing rapid burial followed by heating and subsequent unroofing at moderate to high temperatures have been reported from many mountain belts around the world. Owing to the strong directed forces operative during collision, deformation typically accompanies metamorphism; rocks metamorphosed in response to continent-continent collision generally have fabrics showing a strong preferred orientation of mineral grains, folds on a variety of scales, and pre-, syn-, and postkinematic porphyroblasts. Examples of metamorphic belts produced in response to this type of collision include the Paleozoic Appalachian and Caledonides belts and the Mesozoic-Cenozoic Alpine and Himalayan belts.

      Regionally metamorphosed rocks are also exposed in areas where the crust has been thinned by extensional faulting, such as the Basin and Range province of the western United States. In this type of occurrence, areas of medium- and low-pressure facies series rocks that measure a few tens of kilometres in diameter are juxtaposed against unmetamorphosed sediments or very low-grade metamorphic rocks along low-angle extensional faults. (Metamorphic gradesrefer to the degree and intensity of the metamorphism: they are determined by the pressure and temperatures to which the rock has been subjected.) Such areas are generally referred to as metamorphic core complexes. Metamorphism in these complexes may or may not be related to the extensional event. In some instances, metamorphic rocks produced during much earlier events are simply unroofed and exposed by the faulting but show little or no recrystallization related to extension. In other cases, prolonged extension has resulted in an increased crustal geotherm, and relatively high-temperature metamorphism and magmatism is thus directly related to the extensional event. Immediately adjacent to the faults, the rocks may also be affected by dynamic metamorphism.

      The facies associated with regional metamorphism include, at low grade, the zeolite and prehnite-pumpellyite facies. In areas belonging to high-pressure facies series, the rocks are predominantly in the blueschist and eclogite facies. Medium- and low-pressure facies series are typified by rocks belonging to the greenschist, amphibolite, and granulite facies.

      In the zeolite facies, sediments and volcanic debris show the first major response to burial. Reactions are often not complete, and typical metamorphic fabrics may be poorly developed or not developed at all. This is the facies of burial metamorphism.

      The zeolite facies was first described from southern New Zealand, but similar rocks have now been described from many younger mountain regions of the Earth, particularly around the Pacific margin and the European Alps. Typically, the rocks are best developed where reactive volcanic materials (often partly glassy) are common and the characteristic minerals include zeolites, which are low-density, hydrated silicates, stable at temperatures rarely exceeding 300° C. Typical mineral assemblages include heulandite, analcite, quartz with complex clay minerals (montmorillonite), micaceous phases such as chlorite and celadonite, and the potassium feldspar, adularia. At higher grades of metamorphism, the zeolite laumonite and the feldspar albite dominate the mineral assemblage. In New Zealand these are developed in a rock column that is about 15 kilometres thick. Calcareous rocks (impure limestones) show very little response to this grade of metamorphism.

Prehnite-pumpellyite facies
      Along with the zeolite facies, the prehnite-pumpellyite facies received little attention until about 1950. The first rocks of the facies were described in New Zealand and Celebes. The facies is transitional, bridging the path to the blueschist facies or the greenschist facies. It is particularly well developed in graywacke-type sediments. The two minerals prehnite and pumpellyite replace the zeolite minerals of the zeolite facies and are themselves replaced by epidote minerals in the greenschist facies and by lawsonite and pyroxenes in the blueschist facies. Typical minerals in this facies are quartz, albite, prehnite, pumpellyite, chlorite, stilpnomelane, muscovite, and actinolite. Almost all the minerals are hydrated, and, except for chlorite, they bear little resemblance to the minerals of sediments. This facies has been most described from younger mountain ranges of the Pacific margin.

Blueschist facies (glaucophane facies)
      Rocks of the blueschist facies represent deep metamorphism under conditions of a low thermal gradient. The characteristic locale for this type of metamorphism appears to be along a continental margin being underthrust by an oceanic plate. Regions in which blueschists are found are also regions of great seismic and volcanic activity, such as the Pacific margin. The best described examples of this class of metamorphism come from California, Japan, New Caledonia, Celebes, the Alps, and the Mediterranean region. At present there are no confirmed examples of glaucophane schists predating the Paleozoic Era. Because of the presence of the blue amphibole glaucophane and minerals such as garnet and jadeite, these schists are among the most attractive of metamorphic rocks.

      Characteristic minerals of the facies include quartz, glaucophane, lawsonite, jadeite, omphacite, garnet, albite, chlorite, muscovite, paragonite, epidote, and kyanite. In calcareous rocks, calcite may be replaced by the high-pressure polymorph aragonite. In general, the facies is characterized by many high-density minerals reflecting a high pressure of formation.

Eclogite facies
      The eclogite facies was initially recognized in rocks only of basaltic (basalt)composition, which are transformed at the pressure-temperature conditions of the eclogite facies into spectacular red and green rocks composed of the anhydrous mineral assemblage garnet plus omphacite. The garnet is rich in the high-pressure species pyrope, and the omphacite is rich in the high-pressure pyroxene jadeite. Small amounts of minerals such as kyanite, zoisite, and hornblende may be present. The rocks are of high density and frequently show little or no schistosity. It is now known that protoliths other than basalt also can be metamorphosed to pressures and temperatures characteristic of the eclogite facies, and a wide variety of mineral assemblages can be stable at these conditions, including several hydrous mineral phases. Minerals that have been observed in metapelites include magnesium-rich chloritoid and staurolite, kyanite, garnet, phengite (a muscovite mica with high magnesium and silicon and low aluminum content), chlorite, and talc. Experimental work shows that pelitic rocks composed primarily of talc and kyanite, which are referred to as whiteschists, can be stable from pressures of approximately 6 kilobars up to greater than 30 kilobars. Minerals observed in eclogite-facies calcareous rocks include magnesite, dolomite, zoisite or epidote, and omphacite.

      Because of the high density and composition, it was proposed long ago that part of the upper mantle might be made of eclogite. Such a view is supported by eclogitic intrusions in volcanic rocks and by eclogitic inclusions in diamond-bearing kimberlite, which must come from the upper mantle. Some workers also think that eclogites found in metamorphic terrains in Norway, California, U.S., and the European Alps could also come from the mantle by tectonic processes.

      Early experimental work on eclogites of basaltic bulk composition suggested that eclogites could generally only be stable if water pressure was much lower than the lithostatic pressure, and the facies was thus thought to represent dry, high-pressure metamorphism of basaltic protoliths. Subsequent work on the more diverse protolith compositions reveals, however, that a wide range of water pressures are possible in the eclogite facies and that fluid compositions in equilibrium with the eclogite minerals also probably vary greatly. Indeed, fluid inclusions (tiny bubbles of fluid trapped within mineral grains) in eclogite samples provide evidence of fluids containing nitrogen, salts, and carbon dioxide in addition to water. Eclogite metamorphism is therefore not confined to dry environments but results instead from metamorphism of a variety of rock types at pressures above about 10 kilobars, corresponding to burial to approximately 35 kilometres, and at temperatures ranging from about 400ẓ to 1,000° C. The temperatures of the eclogite facies overlap those of the greenschist, amphibolite, and granulite facies, but the higher pressures result in distinctly different mineral assemblages characterized by high-density mineral phases.

      The greenschist facies was once considered the first major facies of metamorphism proper. The name comes from the abundance of the green mineral chloritein such rocks. Because chlorite and muscovite are ubiquitous and because both exhibit a platy crystal habit, these rocks normally show a highly developed foliation and often exhibit strong metamorphic differentiation. They have been described from practically every metamorphic terrain on Earth, from earliest Precambrian to the young mountain regions. In fact, many of the Earth's oldest rocks (about three billion years old) of the continental-shield areas are in this facies, classic examples of which are in the Appalachians, the Highlands of Scotland, New Zealand, the European Alps, Japan, and Norway.

      The dominant minerals of greenschists formed from silicate-rich sediments include quartz, albite, muscovite, chlorite, epidote, calcite, actinolite, magnetite, biotite, and paragonite. Minerals less common include the manganese-rich garnet spessartine, stilpnomelane, kyanite, rutile, sphene, pyrophyllite, and chloritoid. Calcareous rocks are dominated by calcite, dolomite, and quartz; the major carbonate minerals are thermally stable. It is only when large quantities of water flush away carbon dioxide or keep its partial pressure low that carbonate-silicate reactions take place and liberate carbon dioxide. The typical minerals of this facies have low water contents as compared to the zeolite facies minerals.

      The amphibolite facies is the common high-grade facies of regional metamorphism, and, like the greenschist facies, such rocks are present in all ages from all over the world. Their characteristic feature is the development of the most common amphibole, hornblende, in the presence of a plagioclase feldspar and garnet. The rocks are normally highly foliated or schistose. Many zones or isograds subdividing the facies have been recognized, and classic studies have been made in the Highlands of Scotland, New Hampshire and Vermont in the United States, Switzerland, and the Himalayas.

      Characteristic minerals derived from pelitic rocks are quartz, muscovite, biotite, garnet, plagioclase, kyanite, sillimanite, staurolite, and orthoclase. Minerals derived from basaltic rocks include hornblende, plagioclase, garnet, epidote, and biotite. Those derived from calcareous rocks are calcite, diopside, grossular (garnet), zoisite, actinolite (hornblende), scapolite, and phlogopite. Minerals from magnesium-rich ultramafic rocks are chlorite, anthophyllite, and talc. In most common types, water is present in minerals only of the mica and amphibole families, and, with their water contents of only about 1 to 3 percent, dehydration is nearing its metamorphic climax.

      In rocks of basaltic (basalt)composition, the granulite facies is an anhydrous facies that results from progressive dehydration of amphibolites at high temperature. Rocks of other bulk compositions may retain some hydrous minerals, such as biotite and hornblende, but it is likely that water pressure is lower than lithostatic pressure during most granulite facies metamorphism. Evidence for relatively low water pressures comes from fluid inclusion data indicating carbon dioxide-rich fluid compositions and from preservation of some bulk compositions that should have undergone nearly total melting at granulite temperatures if water pressure had been equal to lithostatic pressure.

      Rocks of this facies frequently have a granular texture quite similar to plutonic igneous rocks. Schistosity is only weakly developed. Typical minerals of the facies are quartz, alkali feldspar, garnet, plagioclase, cordierite, sillimanite, and orthopyroxene. In calcareous members, dolomite, calcite, diopside, and forsterite occur; and it is in this facies that minerals of the scapolite family are best developed. Small amounts of hornblende are often present. A rare mineral occurring in this facies is sapphirine. The rock type charnockite (from Tamil Nādu, India), essentially a orthopyroxene granite, is normally included in this facies.

      It appears from experimental studies that during ultrametamorphism, when melting starts, the basic reactions which take place are of the type

      biotite + other minerals→ melt + residue

      hornblende + other minerals→ melt + residue.

      The first melts to form are partly wet granitic or granodioritic melts, and phases such as biotite and hornblende break down by producing a partly wet melt from the least refractory phases in the rocks. They would persist to much higher temperatures in other systems of their own composition. The residue in the above equations is a granulite-facies metamorphic rock containing phases such as pyroxene and sillimanite. Thus it is probable, but certainly not universally accepted, that many granulites are formed only in the presence of a silicate liquid. This liquid may, of course, move to higher crustal levels.

      Large areas of granulite facies rocks are confined almost entirely to Precambrian areas of the continents (those areas that were formed more than 570 million years ago), with well-developed areas exposed in Canada, India, Africa, Antarctica, Greenland, and the Adirondack Mountains of New York in the northeastern United States. Smaller areas of granulite facies rocks occur in younger mountain belts, with Paleozoic examples in New England (U.S.) and Brittany and Tertiary examples (those formed between 66.4 and 1.6 million years ago) in British Columbia (Can.) and Timor. The apparent decrease in the volume of granulite facies rocks with decreasing age of metamorphism has led some geologists to postulate, as mentioned above, that plate tectonic processes might have changed significantly with time—specifically that steady-state continental geotherms were hotter in the Precambrian than at the present time. Some work on pressure-temperature-time paths in granulites also suggests that Precambrian granulites were metamorphosed along distinctly different paths than younger granulites, lending credence to models invoking changes in tectonic processes. An alternative hypothesis is that large volumes of granulites have been formed throughout Earth history but that they have not yet been exposed by erosion. Pressures calculated from fragments of granulite-facies metamorphic rocks carried to the surface in young volcanic eruptions suggest that the fragments were derived from the lower crust. It is likely that the lower crust is currently composed largely of granulite-facies rocks that may be exposed by future episodes of mountain building, but it is also possible that these granulites will prove to be different from their Precambrian counterparts. In order to resolve some of the controversies surrounding the origin and composition of granulites, it is necessary that considerable studies of these rocks be conducted in the future.

Distribution of metamorphic rocks
      A high-grade metamorphic rock is one that formed at a depth of tens of kilometres and later returned to the surface. Hence, metamorphic regions are also regions of former or recent intense orogeny. More stable regions of the Earth's crusttend to be covered with sediments, and only deep drilling will reveal the metamorphic rocks below.

      The Earth's crust is made up of two basic units, the continents and ocean basins. Exploration of ocean floors has revealed that old, thick sedimentary piles are missing. Doubtless this is related to the processes of continental drift or seafloor spreading; sediments are continuously swept up by continental motion and are added to the continents or returned to the upper mantle (see also plate tectonics (metamorphic rock)). Nearly all studies of metamorphic rocks have concentrated on the continents for this reason.

      There are few large areas of the Earth's crust that are not affected by some type of igneous event from time to time. Although the intensity of volcanism may be focused in certain geographic regions (e.g.,the Pacific margin), volcanism appears to be a rather random phenomenon, at times even occurring in the stable shield areas of the continents. In this sense, contact-metamorphic events may be found almost everywhere at almost any time on Earth. But these metamorphic events are of trivial volumetric significance compared to those of regional metamorphism.

 During the past 500 million years or so of Earth history, major tectonic, seismic, igneous, and metamorphic events have been concentrated on continental margins(Figure 3—>). This has been a period of depression and uplift of the Earth's crust associated with the formation of the present continental distribution. The processes are still going on at dramatic rates in ocean trench environments. These modern regions of activity form immense linear belts. One such belt runs around virtually the entire Pacific margin and another through the Mediterranean and southern Asia to fuse with the circum-Pacific belt. It is in these belts that the spectacular development of zeolite facies, prehnite-pumpellyite facies, blueschist facies, and, occasionally, eclogite facies, as well as the more universal facies of regional metamorphism, have occurred. The granulite facies is almost missing.

      The central and often dominant feature of most continents is their vast Precambrian-shield (continental shield) area; examples include the Canadian Shield, Brazilian Shield, African Shield, and Australian Shield. In these rocks, dating reveals ages of 1 to 3.5 billion years, and they have been little affected by tectonic events postdating the Cambrian. But these shield areas are themselves complex. They consist of vast areas of granitic or granodioritic gneisses. Inside them, between them, and overlapping onto them are belts of sedimentary rocks quite like those in modern sedimentary belts of the Pacific margin or European Alps. These rocks are frequently metamorphosed in the greenschist, amphibolite, and granulite facies. Low-temperature facies and, in particular, low-temperature–high-pressure facies are missing—or have not yet been found. From marginal areas of these stable shield areas, a complex array of processes has been documented covering the past few hundred million years. The Caledonian orogeny(at the close of the Silurian Period) produced tectonic-metamorphic events along the east coast of North America, Greenland, the British Isles, Fennoscandia, Central Asia, and Australia. The Hercynian, or Variscan, orogeny followed about 300 million years ago, affecting subparallel regions and the Urals and European Alps. In fact, the shield margins appear to have been subjected to a more or less constant battering by forces both destroying and rebuilding the margins of these protocontinents. As geologists study Precambrian areas in greater detail, the number of metamorphic and orogenic events recognized on a global scale increases.

      It is the great task and problem of those who study metamorphic rocks to deduce the record of Earth dynamics and thermal history from metamorphic rocks. Among the questions to be answered are (1) whether the pattern of facies development through time—e.g.,the granulite facies in the Archean to blueschist facies in the Tertiary—is a reflection of a cooling Earth and the decline of radioactivity in the crust, and (2) whether the increase in size of global tectonic-metamorphic belts through time reflects changes in convective patterns in the mantle.

      As understanding of the pressure-temperature regimes of metamorphism increases, and as knowledge of rock mechanics and fluid motion during metamorphism also increases through field and laboratory studies, it may become possible to understand the details of the motion of the chemical elements during such processes and hence much of the subject of economic geology, or the search for man's essential raw materials.

Classification of metamorphic rocks
      Because of the diverse chemistry, mineralogy, and primary origin of metamorphic rocks and because of the diverse fabrics or textures that may develop depending on the stresses that may operate during their formation, there is no simple, universally used classification of these rocks. Any classification of metamorphic rocks tends to stress either their fabric, mineralogy, or primary origin. Some common metamorphic rock types are described here.

      Rocks in which metamorphic minerals are easily seen by eye or hand lens and in which the mineral grains have a highly orientated fabric are called schists. Grains of acicular (needlelike) or platy minerals (e.g.,amphiboles and micas) tend to lie with their long directions parallel or their planar directions parallel. Often the rocks show a pronounced mineralogical layering; quartz layers a few millimetres or centimetres in thickness may lie between mica layers, for example. Other words often qualify schist: as described above, greenschist is a schist rich in the green mineral chlorite; blueschist is rich in the blue amphibole, glaucophane; mica-schist is rich in mica; and a graphite-schist is rich in graphite. Schists that are rich in the amphibole hornblende and are often derived by metamorphism of common igneous rocks of the basalt-gabbro type are called amphibolites.

 A very fine-grained metamorphic rock (usually developed from clay-rich sediments) exhibiting perfect planar layering and perfection of splitting into layers (slaty cleavage) is slate (see photograph—>). Such rocks are normally rich in micas and chlorites. As the intensity of metamorphism increases, porphyroblasts may grow; such slates are sometimes called spotted slates. As metamorphism proceeds, the average crystal size increases, and mineral segregation develops; the rock then may be termed a phyllite.

      A gneiss is produced by intense metamorphism, at high temperature and pressure. The grain size is coarser than that in schists, and layering is often well developed; mineral orientation is less perfect than in schists, however. Very common granitic gneisses of Precambrian areas have been derived from metamorphism of granitic igneous rocks.

Hornfels
      The hornfels are formed by contact metamorphism and typically show little sign of the action of directed pressure. They are fine-grained rocks in which crystals display little orientation.

      Rocks derived from the metamorphism of carbonate sediments containing calcite or dolomite are marbles. The main result of metamorphism is an increase in grain size. Because of the rather equidimensional habit of calcite and dolomite crystals, they rarely appear schistose unless they contain other minerals such as mica.

Mylonites and cataclastites (cataclastite)
      These are rocks in which the texture is the result of ductile shearing or mechanical shattering of grains. They often show only slight, if any, development of new minerals. They form on fault planes or in zones of intense shearing. If the crustal rocks have an appropriate composition, phyllonites may develop where new mica crystals grow parallel to the shearing direction. If shearing is extreme, melting may occur, locally producing a pseudo-tachylite (tachylyte). Tachylite is a term applied to certain types of glass formed by rapid cooling of molten rocks.

Other classes
      Most of the above terms indicate structural or fabric classification of metamorphic rocks. Sometimes terms are used to indicate chemical features. Several types of schists, for example, include the following: pelitic schists contain much aluminum oxide and often are derivatives of clay-rich sediments; quartzofeldspathic schists are high in quartz and feldspars and often are derivatives of sandstones or quartz-rich igneous rocks; calcareous schists have a high content of lime (CaO) and often are derivatives of impure limestones, dolomites, or calcareous muds; and mafic schists contain the elements of mafic igneous rocks—namely, calcium, magnesium, and iron.

William S. Fyfe Jane Selverstone     

Additional Reading
Standard references and basic textbooks include, in order of increasing complexity, Con Gillen, Metamorphic Geology: An Introduction to Tectonic and Metamorphic Processes (1982); Roger Mason, Petrology of the Metamorphic Rocks, 2nd ed. (1990); Bruce W.D. Yardley, An Introduction to Metamorphic Petrology (1989); Akiho Miyashiro, Metamorphism and Metamorphic Belts (1973); K. Bucher and M. Frey, Petrogenesis of Metamorphic Rocks, 6th ed. (1994); Martin Frey and Doug Robinson (eds.) Low-grade Metamorphism: An Overview (1999); and Frank S. Spear, Metamorphic Phase Equilibria and Pressure-Temperature-Time Paths (1995). Metamorphic fabric development is discussed in R.H. Vernon, Metamorphic Processes: Reactions and Microstructure Development (1983); Ralph Kretz, Metamorphic Crystallization (1994); and C.W. Passchier and R.A.J. Truow, Micro-tectonics (1996). Major techniques for studying metamorphic rocks and the metamorphic histories of many of the world's mountain belts are reviewed in J.S. Daly, R.A. Cliff, and Bruce W.D. Yardley (eds.), Evolution of Metamorphic Belts (1988); and Peter J. Treloar and Patrick J. O'Brien (eds.), What Drives Metamorphism and Metamorphic Reactions (1998).Jane Selverstone

* * *


Universalium. 2010.

Игры ⚽ Поможем решить контрольную работу

Look at other dictionaries:

  • Metamorphic rock — Metamorphic redirects here. For other uses, see Metamorphic (disambiguation). Quartzite, a form of metamorphic rock, from the Museum of Geology at University of Tartu collection. Metamorphic rock is the transformation of an existing rock type,… …   Wikipedia

  • metamorphic rock — metamorfinė uoliena statusas T sritis chemija apibrėžtis Metamorfizmo pakeista nuosėdinė arba magminė uoliena. atitikmenys: angl. metamorphic rock rus. метаморфическая порода …   Chemijos terminų aiškinamasis žodynas

  • metamorphic rock —    Rock formed from igneous or sedimentary rock which has been geologically changed in its crystalline structure by great heat and pressure. Examples include marble, schist, and slate. Also see stone …   Glossary of Art Terms

  • metamorphic rock —    Rock of any origin altered in mineralogical composition, chemical composition, or structure by heat, pressure, and movement at depth in the earth s crust. Nearly all such rocks are crystalline.    Examples: schist, gneiss, quartzite, slate,… …   Glossary of landform and geologic terms

  • metamorphic rock — GLOSSARY OF VOLCANIC TERMS In the strictest sense, rocks that have formed in the solid state in response to pronounced changes in temperature and/or pressure without any change in the bulk chemical composition of the rock. Metamorphic processes… …   Glossary of volcanic terms

  • metamorphic rock — noun one of the major groups of rock that makes up the crust of the Earth; consists of pre existing rock mass in which new minerals or textures are formed at higher temperatures and greater pressures than those present on the Earths surface …   Wiktionary

  • metamorphic rock — noun rock altered by pressure and heat • Hypernyms: ↑rock, ↑stone • Hyponyms: ↑amphibolite, ↑marlite, ↑marlstone, ↑hornfels, ↑hornstone, ↑gneiss …   Useful english dictionary

  • metamorphic rock —   rocks created by the chemical alteration of existing rocks under the influence of temperature, pressure, or both …   Geography glossary

  • metamorphic rock — /mɛtəˌmɔfɪk ˈrɒk/ (say metuh.mawfik rok) noun See rock1 (def. 1a) …  

  • Metamorphic facies — Eclogite Blueschist Greenschist Prehnite Pumpellyite …   Wikipedia

Share the article and excerpts

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