6 Metamorphic Rocks

Contributing Author: Dr. Peter Davis, Pacific Lutheran University

Learning Objectives

By the end of this chapter, students will be able to:

The rock cycle (clockwise). Magma turns to Igneous rock through crystallization. Igneous rocks turn to sediment through weathering. Sediment turns to sedimentary rocks through transport and deposition and burial and lithification. Sedimentary rocks turn to metamorphic rocks through textural and or chemical damage. Metamorphic rocks turn to magma through melting. Igneous rocks turn to metamorphic rocks through textural and/or chemical damage. metamorphic and sedimentary rocks endure weathering by exhumation of rock back to Earth's surface.
Figure 6.1: Rock cycle showing the five materials (such as igneous rocks and sediment) and the processes by which one changes into another (such as weathering).

Metamorphic rocks, meta- meaning change and –morphos meaning form, is one of the three rock categories in the rock cycle (see chapter 1). Metamorphic rock material has been changed by temperature, pressure, and/or fluids. The rock cycle shows that both igneous and sedimentary rocks can become metamorphic rocks. And metamorphic rocks themselves can be re-metamorphosed. Because metamorphism is caused by plate tectonic motion, metamorphic rock provides geologists with a history book of how past tectonic processes shaped our planet.

6.1 Metamorphic Processes

Metamorphism occurs when solid rock changes in composition and/or texture without the mineral crystals melting, which is how igneous rock is generated. Metamorphic source rocks, the rocks that experience the metamorphism, are called the parent rock or protolith, from proto– meaning first, and lithos- meaning rock. Most metamorphic processes take place deep underground, inside the earth’s crust. During metamorphism, protolith chemistry is mildly changed by increased temperature (heat), a type of pressure called confining pressure, and/or chemically reactive fluids. Rock texture is changed by heat, confining pressure, and a type of pressure called directed stress.

6.1.1 Temperature (Heat)

Temperature measures a substance’s energy—an increase in temperature represents an increase in energy. Temperature changes affect the chemical equilibrium or cation balance in minerals. At high temperatures atoms may vibrate so vigorously they jump from one position to another within the crystal lattice, which remains intact. In other words, this atom swapping can happen while the rock is still solid.

The temperatures of metamorphic rock lies in between surficial processes (as in sedimentary rock) and magma in the rock cycle. Heat-driven metamorphism begins at temperatures as cold as 200˚C, and can continue to occur at temperatures as high as 700°C-1,100°C. Higher temperatures would create magma, and thus, would no longer be a metamorphic process. Temperature increases with increasing depth in the Earth along a geothermal gradient (see chapter 4) and metamorphic rock records these depth-related temperature changes.

6.1.2 Pressure

Pressure is the force exerted over a unit area on a material. Like heat, pressure can affect the chemical equilibrium of minerals in a rock. The pressure that affects metamorphic rocks can be grouped into confining pressure and directed stress. Stress is a scientific term indicating a force. Strain is the result of this stress, including metamorphic changes within minerals.

Confining Pressure

Three 3-D diagrams: the top diagram is labeled Pressure and shows a cube with a single arrow pointing inward on each face, labeled S1, S2, and S3 which correlate with the Z-axis, Y-axis, and X-axis, respectively. The text next to this diagram says "Pressure is a state where all stresses on a body are equal. The magnitude of these balanced stresses increases with increasing depth within the earth. These stresses can not deform rocks other than to decrease their volume. Pressure is the term used because the concept of pressure is used in chemistry, which it the discipline of science used to understand the mineral reactions that occur within the rock." The lower two diagrams are labeled Directed Stresses with the accompanying text "One or more directions of stress are not equal in magnitude and or not in line with each other (non-coaxial). Unlike balanced stresses, the difference in these stresses can deform rocks within the earth." One of these diagrams is labeled Pure Shear (co-axial) and shows a stretched cuboid with S1 arrows pointing downward toward the top face and upward toward the bottom face; the other diagram is labeled Simple Shear (non-co-axial) and shows a sheared cube with S1 shear force arrows parallel to the top and bottom faces pointed in opposite directions; for both of these diagrams, S1 is greater than S2 and S3, with S2 equaling S3.
Figure 6.2: Difference between pressure and stress and how they deform rocks. Pressure (or confining pressure) has equal stress (forces) in all directions and increases with depth under the Earth’s surface. Under directed stress, some stress directions (forces) are stronger than others, and this can deform rocks.

Pressure exerted on rocks under the surface is due to the simple fact that rocks lie on top of one another. When pressure is exerted from rocks above, it is balanced from below and sides, and is called confining or lithostatic pressure. Confining pressure has equal pressure on all sides (see figure 6.2) and is responsible for causing chemical reactions to occur just like heat. These chemical reactions will cause new minerals to form.

Confining pressure is measured in bars and ranges from 1 bar at sea level to around 10,000 bars at the base of the crust. For metamorphic rocks, pressures range from a relatively low-pressure of 3,000 bars around 50,000 bars, which occurs around 15-35 kilometers below the surface.

Directed Stress

Thinly layered rock with flattened parallel pebbles embedded throughout.
Figure 6.3: Pebbles (that used to be spherical or close to spherical) in quartzite deformed by directed stress.

Directed stress, also called differential or tectonic stress, is an unequal balance of forces on a rock in one or more directions (see previous figure). Directed stresses are generated by the movement of lithospheric plates. Stress indicates a type of force acting on rock. Strain describes the resultant processes caused by stress and includes metamorphic changes in the minerals. In contrast to confining pressure, directed stress occurs at much lower pressures and does not generate chemical reactions that change mineral composition and atomic structure. Instead, directed stress modifies the parent rock at a mechanical level, changing the arrangement, size, and/or shape of the mineral crystals. These crystalline changes create identifying textures, which is shown in the figure below comparing the phaneritic texture of igneous granite with the foliated texture of metamorphic gneiss.

Two coarse-grained rock samples that both contain pink, white, and black minerals; one rock sample has random alignment of minerals while the other sample has dark minerals forming thin, black parallel lines with pink and white minerals elongated in the same direction.
Figure 6.4: An igneous rock granite (left) and foliated high-temperature and high-pressure metamorphic rock gneiss (right) illustrating a metamorphic texture.

Directed stresses produce rock textures in many ways. Crystals are rotated, changing their orientation in space. Crystals can get fractured, reducing their grain size. Conversely, they may grow larger as atoms migrate. Crystal shapes also become deformed. These mechanical changes occur via recrystallization, which is when minerals dissolve from an area of rock experiencing high stress and precipitate or regrow in a location having lower stress. For example, recrystallization increases grain size much like adjacent soap bubbles coalesce to form larger ones. Recrystallization rearranges mineral crystals without fracturing the rock structure, deforming the rock like silly putty; these changes provide important clues to understanding the creation and movement of deep underground rock faults.

6.1.3 Fluids

A third metamorphic agent is chemically reactive fluids that are expelled by crystallizing magma and created by metamorphic reactions. These reactive fluids are made of mostly water (H2O) and carbon dioxide (CO2), and smaller amounts of potassium (K), sodium (Na), iron (Fe), magnesium (Mg), calcium (Ca), and aluminum (Al). These fluids react with minerals in the protolith, changing its chemical equilibrium and mineral composition, in a process similar to the reactions driven by heat and pressure. In addition to using elements found in the protolith, the chemical reaction may incorporate substances contributed by the fluids to create new minerals. In general, this style of metamorphism, in which fluids play an important role, is called hydrothermal metamorphism or hydrothermal alteration. Water actively participates in chemical reactions and allows extra mobility of the components in hydrothermal alteration.

Fluids-activated metamorphism is frequently involved in creating economically important mineral deposits that are located next to igneous intrusions or magma bodies. For example, the mining districts in the Cottonwood Canyons and Mineral Basin of northern Utah produce valuable ores such as argentite (silver sulfide), galena (lead sulfide), and chalcopyrite (copper iron sulfide), as well as the native element gold. These mineral deposits were created from the interaction between a granitic intrusion called the Little Cottonwood Stock and country rock consisting of mostly limestone and dolostone. Hot, circulating fluids expelled by the crystallizing granite reacted with and dissolved the surrounding limestone and dolostone, precipitating out new minerals created by the chemical reaction. Hydrothermal alternation of mafic mantle rock, such as olivine and basalt, creates the metamorphic rock serpentinite, a member of the serpentine subgroup of minerals. This metamorphic process happens at mid-ocean spreading centers where newly formed oceanic crust interacts with seawater.

A black smoker hydrothermal vent at the bottom of the sea floor. There is a plume of black smoke coming from a cone-shaped extrusion of rock and a colony of tube worms are attached to the cone-shaped rock.
Figure 6.5: Black smoker hydrothermal vent with a colony of giant (6’+) tube worms.

Some hydrothermal alterations remove elements from the parent rock rather than deposit them. This happens when seawater circulates down through fractures in the fresh, still-hot basalt, reacting with and removing mineral ions from it. The dissolved minerals are usually ions that do not fit snugly in the silicate crystal structure, such as copper. The mineral-laden water emerges from the sea floor via hydrothermal vents called black smokers, named after the dark-colored precipitates produced when the hot vent water meets cold seawater (see chapter 4). Ancient black smokers were an important source of copper ore for the inhabitants of Cyprus (Cypriots) as early as 4,000 BCE, and later by the Romans.

QR code
Take this quiz to check your comprehension of this section.

If you are using an offline version of this text, access the quiz for section 6.1 via the QR code.

 

6.2 Metamorphic Textures

Metamorphic texture is the description of the shape and orientation of mineral grains in a metamorphic rock. Metamorphic rock textures are foliated, non-foliated, or lineated are described below.

Identify rock's foliation Textural features Mineral composition Rock name Parent rock
Foliated (layered texture) Fine-grained or no visible grains Flat, slaty cleavage is well developed. Dense, microscopic grains may exhibit slight sheen (or dull luster). Clanky sound when struck. Breaks into hard, flat sheets. Fine, microscopic clay or mica Slate Shale
Finely crystalline; micas hardly discernible, but impart a sheen or luster. Breaks along wavy surfaces. Dark silicates and micas Phyllite Siltstone or shale
Medium- to coarse-grained Schistose texture. Foliation formed by alignment of visible crystals. Rock breaks along scaly foliation surfaces. Medium to fine-grained. Sparkling appearance. Common minerals include chlorite, biotite, muscovite, garnet, and hornblende. Recognizable minerals used as part of rock name. Porphyroblasts common. Mica schist Siltstone or shale
Garnet schist
Gneissic banding. Coarse-grained. Foliation present as minerals arranged into alternating light and dark layers giving the rock a banded texture in side view. Crystalline texture. No cleavage. Light-colored quartz and feldspar; dark ferromagnesian minerals Gneiss Shale or granitic rocks
Non-foliated (no layered texture) Fine-grained or no visible grains Medium- to coarse-grained crystalline structure. Crystals of amphibole (hornblende) in blade-like crystals Amphibolite Basalt, gabbro, or ultramafic igneous rocks
Microcrystalline texture. Glassy black sheen. Conchoidal fracture. Low density. Fine, tar-like, organic makeup Anthracite coal Coal
Dense and dark-colored. Fine or microcrystalline texture. Very hard. Color can range from gray, gray-green to black. Microscopic dark silicates Hornfels Many rock types
Microcrystalline or no visible grains with smooth, wavy surfaces. May be dull or glossy. Usually shades of green. Serpentine. May have fibrous asbestos visible. Serpentinite Ultramafic igneous rocks or peridotite
Fine- to coarse-grained Microcrystalline or no visible grains. Can be scratched with a fingernail. Shades of green, gray, brown, or white. Soapy feel. Talc Soapstone or talc schist Ultramafic igneous rocks
Crystalline. Hard (scratches glass). Breaks across grains. Sandy or sugary texture. Color variable; can be white, pink, buff, brown, red, purple. Quartz grains fused together. Grains will not rub off like sandstone. Quartzite Quartz sandstone
Finely crystalline (resembling a sugar cube) to medium or coarse texture. Color variable; white, pink, gray, among others. Fossils in some varieties. Calcite or dolomite crystals tightly fused together. Calcite effervesces with HCl; dolomite effervesces only when powdered. Marble Limestone or dolostone
Texture of conglomerate, but breaks across clasts as easily as around them. Pebbles may be stretched (lineated) or cut by rock cleavage. Granules or pebbles are commonly granitic or jasper, chert, quartz, or quartzite. Meta-conglomerate Conglomerate

Table 6.1: Metamorphic rock identification table.

6.2.1 Foliation and Lineation

Foliation is a term used that describes minerals lined up in planes. Certain minerals, most notably the mica group, are mostly thin and planar by default. Foliated rocks typically appear as if the minerals are stacked like pages of a book, thus the use of the term ‘folia’, like a leaf. Other minerals, with hornblende being a good example, are longer in one direction, linear like a pencil or a needle, rather than a planar-shaped book. These linear objects can also be aligned within a rock. This is referred to as a lineation. Linear crystals, such as hornblende, tourmaline, or stretched quartz grains, can be arranged as part of a foliation, a lineation, or foliation/lineation together. If they lie on a plane with mica, but with no common or preferred direction, this is foliation. If the minerals line up and point in a common direction, but with no planar fabric, this is lineation. When minerals lie on a plane AND point in a common direction; this is both foliation and lineation.

Three images: on the left is a photo of a rock sample with elongated tan and black linear minerals visible along on the sides which are seen as dots where they terminate at the rock face perpendicular to the lines; in the middle is a schematic black and white drawing of the rock sample; and on the right is a photo of a bundle of plastic drinking straws that are held together with a rubber band wrapped around them.
Figure 6.6: Example of lineation where minerals are aligned like a stack of straws or pencils.
Two images: on the left is a photo of a grayish tan rock sample with visible layering and elongated linear mineral alignment along the sides; on the face perpendicular to the lineations, layering is still visible but the lineations appear as dots; the right image shows a grayscale schematic drawing of the rock with arrows that label foliation, aligned tourmaline crystals on foliation, and alignment direction.
Figure 6.7: An example of foliation WITH lineation.
Three images: on the left are two photos of a rock sample, one viewed from the top and one viewed from the side; along the top, elongate black crystals have random orientation embedded within tan matrix but on the side, the black crystals have visible layering with the tan matrix; on the right is a schematic grayscale drawing of the rock sample with two arrows: one pointing to the top labeled "Foliated surface displays non-lineated hornblende grains" and another pointing to the side labeled "This surface displays a cross section of foliated plagioclase and hornblende."
Figure 6.8: An example of foliation WITHOUT lineation.

Foliated metamorphic rocks are named based on the style of their foliations. Each rock name has a specific texture that defines and distinguishes it, with their descriptions listed below.

Slate is a fine-grained metamorphic rock that exhibits a foliation called slaty cleavage that is the flat orientation of the small platy crystals of mica and chlorite forming perpendicular to the direction of stressThe minerals in slate are too small to see with the unaided eye. The thin layers in slate may resemble sedimentary bedding, but they are a result of directed stress and may lie at angles to the original strata. In fact, original sedimentary layering may be partially or completely obscured by the foliation. Thin slabs of slate are often used as a building material for roofs and tiles.

Outcrop of tan to brown thinly layered rock.
Figure 6.9: Rock breaking along flat even planes.
Two images: on the left is a photo of a gray rock sample with visible thin layering that runs vertically through the sample and also a bedding plane that runs roughly left to right; the right image shows a grayscale schematic drawing of the rock with arrows that label the foliation direction and bedding direction.
Figure 6.10: Foliation versus bedding. Foliation is caused by metamorphism. Bedding is a result of sedimentary processes. They do not have to align.
Sample of tan rock with a slight sheen that has a small fold and thin layering; a US quarter rests on top of it for scale.
Figure 6.11: Phyllite with a small fold.

Phyllite is a foliated metamorphic rock in which platy minerals have grown larger and the surface of the foliation shows a sheen from light reflecting from the grains, perhaps even a wavy appearance, called crenulations. Similar to phyllite but with even larger grains is the foliated metamorphic rock schist, which has large platy grains visible as individual crystals. Common minerals are muscovite, biotite, and porphyroblasts of garnets. A porphyroblast is a large crystal of a particular mineral surrounded by small grains. Schistosity is a textural description of foliation created by the parallel alignment of platy visible grains. Some schists are named for their minerals such as mica schist (mostly micas), garnet schist (mica schist with garnets), and staurolite schist (mica schists with staurolite).

Sample of grayish-tan rock with a silky sheen that has thin layering; a scale bar rests on the sample that says 3 cm.
Figure 6.12: Schist.
Two photos of similar rocks that have visible foliation: the rocks have a silvery sheen and sparse small glassy brown crystals throughout. Arrows labeled "Garnet staurolite muscovite schist" point to the small brown crystals.
Figure 6.13: Garnet staurolite muscovite schist.
Sample of rock that has distinct black and white parallel bands throughout.
Figure 6.14: Gneiss.

Gneissic banding is a metamorphic foliation in which visible silicate minerals separate into dark and light bands or lineations. These grains tend to be coarse and often folded. A rock with this texture is called gneiss. Since gneisses form at the highest temperatures and pressures, some partial melting may occur. This partially melted rock is a transition between metamorphic and igneous rocks called a migmatite.

Rock that has distinct black and white swirling bands; a person's finger points to the rock for scale.
Figure 6.15: Migmatite, a rock which was partially molten.

Migmatites appear as dark and light banded gneiss that may be swirled or twisted some since some minerals started to melt. Thin accumulations of light colored rock layers can occur in a darker rock that are parallel to each other, or even cut across the gneissic foliation. The lighter colored layers are interpreted to be the result of the separation of a felsic igneous melt from the adjacent highly metamorphosed darker layers, or injection of a felsic melt from some distance away.

6.2.2 Non-foliated

Two chunks of rock, one pink in color and the other white in color; both have interlocking crystals; a US quarter rests on the pink sample for scale.
Figure 6.16: Marble.
Chunk of pinkish tan rock with interlocking crystals; a US quarter rests on the sample for scale.
Figure 6.17: Baraboo quartzite.

Non-foliated textures do not have lineations, foliations, or other alignments of mineral grains. Non-foliated metamorphic rocks are typically composed of just one mineral, and therefore, usually show the effects of metamorphism with recrystallization in which crystals grow together, but with no preferred direction. The two most common examples of non-foliated rocks are quartzite and marble. Quartzite is a metamorphic rock from the protolith sandstone. In quartzites, the quartz grains from the original sandstone are enlarged and interlocked by recrystallization. A defining characteristic for distinguishing quartzite from sandstone is that when broken with a rock hammer, the quartz crystals break across the grains. In a sandstone, only a thin mineral cement holds the grains together, meaning that a broken piece of sandstone will leave the grains intact. Because most sandstones are rich in quartz, and quartz is a mechanically and chemically durable substance, quartzite is very hard and resistant to weathering.

Marble is metamorphosed limestone (or dolostone) composed of calcite (or dolomite). Recrystallization typically generates larger interlocking crystals of calcite or dolomite. Marble and quartzite often look similar, but these minerals are considerably softer than quartz. Another way to distinguish marble from quartzite is with a drop of dilute hydrochloric acid. Marble will effervesce (fizz) if it is made of calcite.

A third non-foliated rock is hornfels identified by its dense, fine grained, hard, blocky or splintery texture composed of several silicate minerals. Crystals in hornfels grow smaller with metamorphism, and become so small that specialized study is required to identify them. These are common around intrusive igneous bodies and are hard to identify. The protolith of hornfels can be even harder to distinguish, which can be anything from mudstone to basalt.

Zoomed-in view of a chunk of pinkish tan rock with interlocking glassy crystals.
Figure 6.18: Macro view of quartzite. Note the interconnectedness of the grains.
Close-up photo of unconsolidated amber-colored glassy grains with rounded edges; a scale bar at the lower right says 1.0 mm.
Figure 6.19: Unmetamorphosed, unconsolidated sand grains have space between the grains.

QR code
Take this quiz to check your comprehension of this section.

If you are using an offline version of this text, access the quiz for section 6.2 via the QR code.

 

6.3 Metamorphic Grade

Rock with a silvery sheen and large polygonal brown crystals scattered throughout; a Euro coin rests on the outcrop for scale.
Figure 6.20: Garnet schist.

Metamorphic grade refers to the range of metamorphic change a rock undergoes, progressing from low (little metamorphic change) grade to high (significant metamorphic change) grade. Low-grade metamorphism begins at temperatures and pressures just above sedimentary rock conditions. The sequence slatephylliteschistgneiss illustrates an increasing metamorphic grade.

Geologists use index minerals that form at certain temperatures and pressures to identify metamorphic grade. These index minerals also provide important clues to a rock’s sedimentary protolith and the metamorphic conditions that created it. Chlorite, muscovite, biotite, garnet, and staurolite are index minerals representing a respective sequence of low-to-high grade rock. The figure shows a phase diagram of three index minerals—sillimanite, kyanite, and andalusite—with the same chemical formula (Al2SiO5) but having different crystal structures (polymorphism) created by different pressure and temperature conditions.

QR code
Complete this interactive activity to check your understanding.

If you are using an offline version of this text, access this interactive activity via the QR code.

 

Some metamorphic rocks are named based on the highest grade of index mineral present. Chlorite schist includes the low-grade index mineral chlorite. Muscovite schist contains the slightly higher grade muscovite, indicating a greater degree of metamorphism. Garnet schist includes the high grade index mineral garnet, and indicating it has experienced much higher pressures and temperatures than chlorite.

QR code
Take this quiz to check your comprehension of this section.

If you are using an offline version of this text, access the quiz for section 6.3 via the QR code.

 

6.4 Metamorphic Environments

As with igneous processes, metamorphic rocks form at different zones of pressure (depth) and temperature as shown on the pressure-temperature (P-T) diagram. The term facies is an objective description of a rock. In metamorphic rocks facies are groups of minerals called mineral assemblages. The names of metamorphic facies on the pressure-temperature diagram reflect minerals and mineral assemblages that are stable at these pressures and temperatures and provide information about the metamorphic processes that have affected the rocks. This is useful when interpreting the history of a metamorphic rock.

Pressure increases upward on the vertical axis and temperature increases toward the right on the horizontal axis. High pressure with high temperature is associated with subduction metamorphism while low pressure and high temperature is associated with volcanic arc and continental metamorphism.
Figure 6.21: Pressure–temperature graphs of various metamorphic facies.

In the late 1800s, British geologist George Barrow mapped zones of index minerals in different metamorphic zones of an area that underwent regional metamorphism. Barrow outlined a progression of index minerals, named the Barrovian Sequence, that represents increasing metamorphic grade: chlorite (slates and phyllites) -> biotite (phyllites and schists) -> garnet (schists) -> staurolite (schists) -> kyanite (schists) -> sillimanite (schists and gneisses).

Map of Scotland with metamorphic zones highlighted; metamorphic grade increases toward the northwest, away from the Highland Boundary Fault in the southeast.
Figure 6.22: Barrovian sequence in Scotland.

The first of the Barrovian sequence has a mineral group that is commonly found in the metamorphic greenschist facies. Greenschist rocks form under relatively low pressure and temperatures and represent the fringes of regional metamorphism. The “green” part of the name is derived from green minerals like chlorite, serpentine, and epidote, and the “schist” part is applied due to the presence of platy minerals such as muscovite.

Many different styles of metamorphic facies are recognized, tied to different geologic and tectonic processes. Recognizing these facies is the most direct way to interpret the metamorphic history of a rock. A simplified list of major metamorphic facies is given below.

6.4.1 Burial Metamorphism

Burial metamorphism occurs when rocks are deeply buried, at depths of more than 2000 meters (1.24 miles). Burial metamorphism commonly occurs in sedimentary basins, where rocks are buried deeply by overlying sediments. As an extension of diagenesis, a process that occurs during lithification (chapter 5), burial metamorphism can cause clay minerals, such as smectite, in shales to change to another clay mineral illite. Or it can cause quartz sandstone to metamorphose into the quartzite such the Big Cottonwood Formation in the Wasatch Range of Utah. This formation was deposited as ancient near-shore sands in the late Proterozoic (see chapter 7), deeply buried and metamorphosed to quartzite, folded, and later exposed at the surface in the Wasatch Range today. Increase of temperature with depth in combination with an increase of confining pressure produces low-grade metamorphic rocks with a mineral assemblages indicative of a zeolite facies.

6.4.2 Contact Metamorphism

Contact metamorphism occurs in rock exposed to high temperature and low pressure, as might happen when hot magma intrudes into or lava flows over pre-existing protolith. This combination of high temperature and low pressure produces numerous metamorphic facies. The lowest pressure conditions produce hornfels facies, while higher pressure creates greenschist, amphibolite, or granulite facies.

As with all metamorphic rock, the parent rock texture and chemistry are major factors in determining the final outcome of the metamorphic process, including what index minerals are present. Fine-grained shale and basalt, which happen to be chemically similar, characteristically recrystallize to produce hornfels. Sandstone (silica) surrounding an igneous intrusion becomes quartzite via contact metamorphism, and limestone (carbonate) becomes marble.

Outcrop of thick beds of flat-lying dull tan rock on top of pink rock with a chalky appearance.
Figure 6.23: Contact metamorphism in outcrop.

When contact metamorphism occurs deeper in the Earth, metamorphism can be seen as rings of facies around the intrusion, resulting in aureoles. These differences in metamorphism appear as distinct bands surrounding the intrusion, as can be seen around the Alta Stock in Little Cottonwood Canyon, Utah. The Alta Stock is a granite intrusion surrounded first by rings of the index minerals amphibole (tremolite) and olivine (forsterite), with a ring of talc (dolostone) located further away.

6.4.3 Regional Metamorphism

Regional metamorphism occurs when parent rock is subjected to increased temperature and pressure over a large area, and is often located in mountain ranges created by converging continental crustal plates. This is the setting for the Barrovian sequence of rock facies, with the lowest grade of metamorphism occurring on the flanks of the mountains and highest grade near the core of the mountain range, closest to the convergent boundary.

An example of an old regional metamorphic environment is visible in the northern Appalachian Mountains while driving east from New York state through Vermont and into New Hampshire. Along this route the degree of metamorphism gradually increases from sedimentary parent rock, to low-grade metamorphic rock, then higher-grade metamorphic rock, and eventually the igneous core. The rock sequence is sedimentary rock, slate, phyllite, schist, gneiss, migmatite, and granite. In fact, New Hampshire is nicknamed the Granite State. The reverse sequence can be seen heading east, from eastern New Hampshire to the coast.

6.4.4 Subduction Zone Metamorphism

Sample of dark blue rock with dark green bands throughout; silvery mica grains can also be seen sparsely throughout the sample; a US quarter rests on the sample for scale.
Figure 6.24: Blueschist.

Subduction zone metamorphism is a type of regional metamorphism that occurs when a slab of oceanic crust is subducted under continental crust (see chapter 2). Because rock is a good insulator, the temperature of the descending oceanic slab increases slowly relative to the more rapidly increasing pressure, creating a metamorphic environment of high pressure and low temperature. Glaucophane, which has a distinctive blue color, is an index mineral found in blueschist facies (see metamorphic facies diagram). The California Coast Range near San Francisco has blueschistfacies rocks created by subduction-zone metamorphism, which include rocks made of blueschist, greenstone, and red chert. Greenstone, which is metamorphized basalt, gets its color from the index mineral chlorite.

6.4.5 Fault Metamorphism

Two images: on the left is a side view of a rock sample with thin parallel layers and rounded grains embedded in some places, around which the layers are slightly offset; a US quarter rests against the sample for scale; on the right is a grayscale drawing of the sample with arrows pointing in the direction of stress: the arrow above the sample points toward the left and the arrow below the sample points toward the right.
Figure 6.25: Mylonite.

There are a range of metamorphic rocks made along faults. Near the surface, rocks are involved in repeated brittle faulting produce a material called rock flour, which is rock ground up to the particle size of flour used for food. At lower depths, faulting create cataclastites, chaotically-crushed mixes of rock material with little internal texture. At depths below cataclasites, where strain becomes ductile, mylonites are formed. Mylonites are metamorphic rocks created by dynamic recrystallization through directed shear forces, generally resulting in a reduction of grain size. When larger, stronger crystals (like feldspar, quartz, garnet) embedded in a metamorphic matrix are sheared into an asymmetrical eye-shaped crystal, an augen is formed.

Cross sectional view of a rock sample that shows distinct thin layering of dark gray, brown, and black minerals; rounded white crystals that do not deform as easily form lens-shapes among the layers; an arrow above the sample points toward the left and an arrow below the sample points toward the right; a scale bar near the bottom of the sample says 1 cm.
Figure 6.26: Examples of augens.

6.4.6 Shock Metamorphism

A microscopic view of a grain that has visible thin lines forming a V-shape through the grain; it has a prismatic inside with a blue center, purple ring around the center, and yellow outer ring; outside of the ring the mineral is white.
Figure 6.27: Shock lamellae in a quartz grain.

Shock (also known as impact) metamorphism is metamorphism resulting from meteor or other bolide impacts, or from a similar high-pressure shock event. Shock metamorphism is the result of very high pressures (and higher, but less extreme temperatures) delivered relatively rapidly. Shock metamorphism produces planar deformation features, tektites, shatter cones, and quartz polymorphs. Shock metamorphism produces planar deformation features (shock laminae), which are narrow planes of glassy material with distinct orientations found in silicate mineral grains. Shocked quartz has planar deformation features

Person holding a cone-shaped rock with lines running along the exterior that converge toward the point of the cone.
Figure 6.28: Shatter cone.

Shatter cones are cone-shaped pieces of rock created by dynamic branching fractures caused by impacts. While not strictly a metamorphic structure, they are common around shock metamorphism. Their diameter can range from microscopic to several meters. Fine-grained rocks with shatter cones show a distinctive horsetail pattern.

Shock metamorphism can also produce index minerals, though they are typically only found via microscopic analysis. The quartz polymorphs coesite and stishovite are indicative of impact metamorphism. As discussed in chapter 3, polymorphs are minerals with the same composition but different crystal structures. Intense pressure (> 10 GPa) and moderate to high temperatures (700-1200 °C) are required to form these minerals.

Two elongated dark brownish black shiny rocks, one resembling the shape of a dumbbell and the other resembling the shape of a teardrop.
Figure 6.29: Tektites.

Shock metamorphism can also produce glass. Tektites are gravel-size glass grains ejected during an impact event. They resemble volcanic glass but, unlike volcanic glass, tektites contain no water or phenocrysts, and have a different bulk and isotopic chemistry. Tektites contain partially melted inclusions of shocked mineral grains. Although all are melt glasses, tektites are also chemically distinct from trinitite, which is produced from thermonuclear detonations, and fulgurites, which are produced by lightning strikes. All geologic glasses not derived from volcanoes can be called with the general term pseudotachylytes, a name which can also be applied to glasses created by faulting. The term pseudo in this context means ‘false’ or ‘in the appearance of’, a volcanic rock called tachylite because the material observed looks like a volcanic rock, but is produced by significant shear heating.

QR code
Take this quiz to check your comprehension of this section.

If you are using an offline version of this text, access the quiz for section 6.4 via the QR code.

 

QR code
Video 6.1: Identifying metamorphic rocks.

If you are using an offline version of this text, access this YouTube video via the QR code.

 

QR code
Video 6.2: Identifying metamorphic rocks.

If you are using an offline version of this text, access this YouTube video via the QR code.

Summary

Metamorphism is the process that changes existing rocks (called protoliths) into new rocks with new minerals and new textures. Increases in temperature and pressure are the main causes of metamorphism, with fluids adding important mobilization of materials. The primary way metamorphic rocks are identified is with texture. Foliated textures come from platy minerals forming planes in a rock, while non-foliated metamorphic rocks have no internal fabric. Grade describes the amount of metamorphism in a rock, and facies are a set of minerals that can help guide an observer to an interpretation of the metamorphic history of a rock. Different tectonic or geologic environments cause metamorphism, including collisions, subduction, faulting, and even impacts from space.

QR code
Take this quiz to check your comprehension of this chapter.

If you are using an offline version of this text, access the quiz for chapter 6 via the QR code.

 

Text References

  1. Bucher, K., and Grapes, R., 2011, Petrogenesis of metamorphic rocks: Springer, 341 p.
  2. Jeong, I.-K., Heffner, R.H., Graf, M.J., and Billinge, S.J.L., 2003, Lattice dynamics and correlated atomic motion from the atomic pair distribution function: Phys. Rev. B Condens. Matter, v. 67, no. 10, p. 104301.
  3. Marshak, S., 2009, Essentials of Geology, 3rd or 4th Edition.
  4. Proctor, B.P., McAleer, R., Kunk, M.J., and Wintsch, R.P., 2013, Post-Taconic tilting and Acadian structural overprint of the classic Barrovian metamorphic gradient in Dutchess County, New York: Am. J. Sci., v. 313, no. 7, p. 649–682.
  5. Timeline of Art History, 2007, Reference Reviews, v. 21, no. 8, p. 45–45.

Figure References

Figure 6.1: Rock cycle showing the five materials (such as igneous rocks and sediment) and the processes by which one changes into another (such as weathering). Kindred Grey. 2022. CC BY 4.0.

Figure 6.2: Difference between pressure and stress and how they deform rocks. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.3: Pebbles (that used to be spherical or close to spherical) in quartzite deformed by directed stress. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.4: An igneous rock granite (left) and foliated high-temperature and high-pressure metamorphic rock gneiss (right) illustrating a metamorphic texture. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.5: Black smoker hydrothermal vent with a colony of giant (6’+) tube worms. NOAA. 2006. Public domain. https://commons.wikimedia.org/wiki/File:Main_Endeavour_black_smoker.jpg

Table 6.1: Metamorphic rock identification table. Kindred Grey. 2022. Table data from Belinda Madsen’s graphic in CH 6 of An Introduction to Geology. OpenStax. Salt Lake Community College. CC BY-NC-SA 4.0.

Figure 6.6: Example of lineation where minerals are aligned like a stack of straws or pencils. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.7: An example of foliation WITH lineation. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.8: An example of foliation WITHOUT lineation. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.9: Rock breaking along flat even planes. Uta Baumfelder. 2010. Public domain. https://commons.wikimedia.org/wiki/File:Ehemaliger_Schiefertagebau_am_Brand.JPG

Figure 6.10: Foliation versus bedding. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.11: Phyllite with a small fold. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.12: Schist. Michael C. Rygel. 2012. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Schist_detail.jpg

Figure 6.13: Garnet staurolite muscovite schist. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.14: Gneiss. Siim Sepp. 2005. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Gneiss.jpg

Figure 6.15: Migmatite, a rock which was partially molten. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.16: Marble. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.17: Baraboo quartzite. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.18: Macro view of quartzite. Manishwiki15. 2012. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Sample_of_Quartzite.JPG

Figure 6.19: Unmetamorphosed, unconsolidated sand grains have space between the grains. Wilson44691. 2008. Public domain. https://commons.wikimedia.org/wiki/File:CoralPinkSandDunesSand.JPG

Figure 6.20: Garnet schist. Graeme Churchard (GOC53). 2005. CC BY 2.0. https://commons.wikimedia.org/wiki/File:Garnet_Mica_Schist_Syros_Greece.jpg

Figure 6.21: Pressure–temperature graphs of various metamorphic facies. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.22: Barrovian sequence in Scotland. Woudloper. 2009. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Scotland_metamorphic_zones_EN.svg

Figure 6.23: Contact metamorphism in outcrop. Random Tree. 2012. Public domain. https://commons.wikimedia.org/wiki/File:Metamorphic_Aureole_in_the_Henry_Mountains.JPG

Figure 6.24: Blueschist. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.25: Mylonite. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.26: Examples of augens. Peter Davis. 2017. CC BY-NC-SA 4.0. https://slcc.pressbooks.pub/introgeology/chapter/6-metamorphic-rocks/

Figure 6.27: Shock lamellae in a quartz grain. Glen A. Izett. 2000. Public domain. https://commons.wikimedia.org/wiki/File:820qtz.jpg

Figure 6.28: Shatter cone. JMGastonguay. 2014. CC BY-SA 4.0. https://commons.wikimedia.org/wiki/File:ShatterConeCharlevoix1.jpg

Figure 6.29: Tektites. Brocken Inaglory. 2007. CC BY-SA 3.0. https://en.wikipedia.org/wiki/File:Two_tektites.JPG

 

definition

License

Icon for the Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License

Introduction to Earth Science Copyright © 2023 by Laura Neser is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License, except where otherwise noted.

Share This Book