You are on page 1of 12

Types of Metamorphism

Somnath Dasgupta, Indian Institute of Science Education Research-Kolkata, Mohanpur, India


Santanu Kumar Bhowmik, Department of Geology & Geophysics, Indian Institute of Technology, Kharagpur, India
© 2020 Elsevier Inc. All rights reserved.

Preamble 1
Early Attempts to Look at the Process of Metamorphism as Basis for Classification 1
Complexities in Classification 2
The Basis of the Adopted Metamorphic Classification 3
Types of Metamorphism 4
Types of Metamorphism in Convergent Plate Margin 4
Regional Thermal Metamorphism 4
Regional Dynamic Metamorphism 5
Burial metamorphism (accretionary) 5
Dislocation/cataclastic metamorphism 5
Regional Dynamo-Thermal Metamorphism 6
Orogenic subduction type 6
Orogenic collision type 7
Metamorphism in Divergent Plate Margins 8
Metamorphism in continental extension zone 8
Ocean floor/ocean ridge metamorphism 8
Metamorphism in Plate Interiors 8
Contact metamorphism 8
Burial metamorphism 9
Other Types of Metamorphism 9
Shock/Impact Metamorphism 9
Lightning Metamorphism 9
Hydrothermal Metamorphism 10
Asteroidal Metamorphism 10
Acknowledgments 10
References 10
Further Reading 12

Preamble

The word Metamorphism is derived from the Greek word “Metamorphosis” (meaning change or alteration) due to changes in
physical and chemical parameters, which are known as agents of metamorphism. Broadly, these are: pressure, temperature and
chemically active fluids. The pressure component consists of (a) lithostatic pressure, which is hydrostatic in nature, and
(b) deviatoric stress or directed pressure, which is directional. In the case of terrestrial metamorphism, the lower limit of
metamorphism is conventionally defined by that of diagenesis, which is essentially gradational and overlapping (termed anchi-
metamorphism or anchizonal metamorphism, discussed in Merriman and Peacor, 1999). The beginning of metamorphism is
customarily set at around 150–200  C and 1 kbar (roughly corresponding to a depth of 3 km). Near surface processes, such as
weathering and diagenesis are, therefore, excluded from the definition of metamorphism. The upper thermal limit of metamor-
phism can be safely set at the condition when rocks become largely molten, which depends on the nature and composition of the
original rock (protolith), the extent of heating and the presence or absence of fluids. Over the last four decades the upper limits of
both lithostatic pressure and temperature have been extended significantly, as discussed under ultrahigh temperature (UHT, TMax
 900  C) and ultrahigh pressure (UHP, P at TMax in excess of quartz-coesite transition curve) metamorphism in a subsequent
section.

Early Attempts to Look at the Process of Metamorphism as Basis for Classification

Eskola (1915, 1920, 1939) looked at the interplay of pressure—temperature with bulk compositional variability of precursors of
metamorphic rocks through the concept of metamorphic facies and mineral facies, which laid a physicochemical basis for
classification of metamorphism. The physicochemical principles of heterogeneous equilibrium accounting for transformation of
rocks were emphasized by Harker (1932). A quantum jump in our understanding happened when Miyashiro (1961) put forward
his idea of varying P/T ratios in different metamorphic terranes identifying different types of metamorphic facies series, and later

Encyclopedia of Geology, 2nd edition https://doi.org/10.1016/B978-0-08-102908-4.00114-4 1


2 Types of Metamorphism

correlated these with global tectonics (Miyashiro, 1972, 1973). Almost simultaneously, Coombs (1961) observed regional
metamorphism in a non-orogenic setting, while Miyashiro (1971) described large scale metamorphism on the ocean floor near
mid-oceanic ridges. All these pioneering studies established a much enlarged domain of terrestrial metamorphism in terms of
variability, classification and causative factors.
Over the last half a century, generation of a significant database on metamorphism across all tectonic domains and geologic age,
startling discoveries (UHP metamorphism—Chopin, 1984; UHT metamorphism—Dallwitz, 1968), and improved experimental
and theoretical analysis contributed towards a comprehensive classification of metamorphic processes and their products. It was
also realized that several methods of classification are required, which results in overlaps and repetition.

Complexities in Classification

Traditionally, terrestrial metamorphism has been classified in terms of parameters such as (a) agents of metamorphism (T and/or
P), (b) spatial extent of metamorphism (regional vs. local), (c) geological setting of metamorphism (contact, shock, cataclastic or
regional), (d) particular cause of metamorphism (e.g., contact, impact, hydrothermal, combustion, lightning), (e) orogenic setting
of metamorphism (orogenic vs. anorogenic), (f ) single or multiple phases of metamorphism and (g) plate tectonic setting of
metamorphism (plate boundary vs. plate interior) (Spear, 1993; Best, 2002; Bucher and Grapes, 2011; IUGS classification,
Smulikowski et al., 2007; Winter, 2010). Table 1 shows a compilation of the different schemes of classification of major types of
metamorphism primarily adopted from Spear (1993), with additional or alternate names inserted from Smulikowski et al. (2007)
and Bucher and Grapes (2011). Other less pervasive types of metamorphism are mentioned in a later section. Definitions of the
types of metamorphism mentioned in Table 1 are given in Table 2.

Table 1 Summary table showing types of metamorphism.

Principal criterion Types of metamorphism

I. Agents of Thermal (T) Dynamic Dynamo-thermal (T


metamorphism Cataclastic or Burial (PLithostatic) + PLithostatic + PDirected)
(T, P or T + P) dislocation
(PDirected)
II. Geological setting of Local Regional
metamorphism
Contact Shock or High strain (around Burial Ocean-ridge Orogenic or collisional
Impact fault/ductile shear or Ocean- and subduction
zone) floor
III. Orogenic setting of Anorogenic (e.g., Burial) Orogenic
metamorphism
IV. Plate tectonic setting Plate interior e.g., Divergent plate margin (e.g., ocean- Transform plate Convergent plate margin (e.g., orogenic,
of metamorphism contact, burial, pyro- ridge, contact around intrusives in margin (e.g., dynamo-thermal, regional, regional
metamorphism oceanic crust) cataclastic) contact, cataclastic, subduction and
collisional)

After Spear FS (1993) Metamorphic Phase Equilibria and Pressure–Temperature—Time Paths. Monograph Series, vol. 1. Mineralogical Society of America: Washington, DC;
Smulikowski W, Desmons J, Harte B, Sassi FP and Schmid R (2007) A Systematic Nomenclature for Metamorphic Rocks: 2. Types, Grade and Facies of Metamorphism.
Recommendations by the IUGS Subcommission on the Systematics of Metamorphic Rocks. SCMR website: Hwww.bgs.ac.uk/SCMRH; Bucher K and Grapes R (2011) Petrogenesis of
Metamorphic Rocks. Heidelberg: Springer, pp. 426.

Table 2 A summary table showing the key features of different types of metamorphism.

Thermal or contact metamorphism Caused by intrusion of magmatic bodies in any tectonic setting. The affected part of the host rock is called contact
aureole. Fluids extracted from cooling plutons may cause additionally hydrothermal metamorphism
Orogenic, regional or dynamo-thermal Associated with mountain building processes in convergent plate boundaries, and can be broadly subdivided into
metamorphism collisional and subduction types
Ocean floor or Ocean ridge Occurs mostly in the ocean ridges, associated with circulation of hydrothermal fluids (causing hydrothermal
metamorphism metamorphism), and these are carried away by spreading oceanic crust, and thus occupy large areas of the ocean
floor
Burial metamorphism Occurs in the continents, typically caused by burial of sediments in large basins and is not associated with any large
scale deformation
Cataclastic or dislocation Metamorphism in zones of high strain, The rocks are usually fault gouge, but also includes pseudo-tachylytes.
metamorphism
Shock and impact metamorphism Produced by meteorite impact, which raises the temperature and pressure to exceptionally high values, but for very
short duration
Pyrometamorphism A special variety of contact metamorphism that is driven by extreme heat at or near surface conditions
Hydrothermal metamorphism Mineral transformation in rocks caused by passage of hot solutions in a variety of tectonic settings
Types of Metamorphism 3

Any classification scheme based on multiple parameters will have overlaps. A few points of concern are mentioned here to
highlight the problems. There is a distinct overlap between plate tectonic setting and geological setting and even the nature of
orogeny in one type of plate boundary (say convergent) can be of several types, each with characteristic outputs. Contact
metamorphism, caused by igneous intrusions, can be of local scale (plate interior, small intrusions) or of regional scale (arc
environment with multiple intrusions and at different depth levels). Additionally, several recent studies reveal that the relationship
between the two above is more complicated. Thus, some of the regional metamorphic terranes record more than one cycle of
heating-cooling history and that the last cycle is anorogenic-scale heating (cf. a regional contact metamorphism) superposed on an
earlier regional metamorphic event (e.g., Sizova et al., 2019; Faryad and Cuthbert, 2020). It has also been increasingly realized that
the drivers of metamorphism (e.g., P and T ) progressively change in a particular plate tectonic setting and give rise to metamorphic
rocks that differ in P-T history. Thus, in a convergent plate boundary of oceanic subduction, deep burial of oceanic lithosphere in the
mature stages of subduction produces cold blueschists and eclogites that are traditionally studied as markers of ocean closure in
suture zones (e.g., Ernst, 1988). However, during the infancy of the same subduction cycle, basaltic crust and oceanic sediments are
accreted at the base of the warm mantle wedge, producing significantly warmer rocks (cf. metamorphic sole rocks) such as high-
pressure granulites and amphibolites (e.g., Agard et al., 2016).

The Basis of the Adopted Metamorphic Classification

It becomes clear from the review above that all these schemes of classification are useful to understand the range of metamorphic
processes that operates in the Planet Earth. The review also predicts significant variability of metamorphic processes within and
across plate tectonic settings. This necessitates an integration of the different parameters of classification and considering meta-
morphism in a larger perspective where processes and products are interlinked. In case of so-called regional scale metamorphism,
both pressure and temperature act as agents, and, as first pointed out by Miyashiro (1961), it is the ratio P (lithostatic)/T that
changes in different metamorphic terranes. This gives rise to different metamorphic facies series (or different types of metamor-
phism per se), which can be correlated with different plate tectonic settings (Miyashiro, 1972, 1973). The three types of
metamorphism after Miyashiro (1961) are: low-P/T, medium-P/T and high-P/T. This concept has been extended with incorporation
of a very large dataset that was collected later and with the emergence of new concepts within the broad realm of plate tectonics in
subsequent years (Brown, 2009; Brown and Johnson, 2019a). Like Miyashiro (1961), Brown and his-co-workers also utilized T/P
(note the reverse) ratios at peak metamorphism to identify three types of metamorphism: low-T/P, intermediate-T/P and high-T/P,
which they correlated with discrete tectonic settings. The P-T fields in the two sets of classification (Miyashiro (M) vs. Brown and
co-workers (BJ)) are compared in Fig. 1.

Fig.1 A scaled P-T diagram showing the fields of three thermobaric (after Brown M and Johnson T (2019a) The 51st Hallimond lecture: Time’s arrow, time’s cycle:
Granulite metamorphism and geodynamics. Mineralogical Magazine 83: 323–338, https://doi.org/10.1180/mgm.2019.19) or baric (after Miyashiro A (1961)
Evolution of metamorphic belts. Journal of Petrology 2: 277–311) types of metamorphism is contoured with model thermal gradients (values in GPa/ C). The
alumina-silicate triple point junction is after Pattison DRM (1992) Stability of andalusite and sillimanite and the Al2SiO5 triple point: Constraints from the Ballachulish
aureole, Scotland. The Journal of Geology 100: 423–446, https://doi.org/10.1086/629596. The metamorphic types in the two classifications are referred by BJ and
M respectively. The P-T fields for the Miyashiro classification are taken from Spear FS (1993) Metamorphic Phase Equilibria and Pressure–Temperature—Time
Paths. Monograph Series, vol. 1. Mineralogical Society of America: Washington, DC. See text for further details.
4 Types of Metamorphism

Four important generalizations emerge from Fig. 1:

(1) The Brown dataset shows a bias towards metamorphic rocks along intermediate- and high-T/P ratios with a TMax in excess of
600  C and with its upper limit reaching 1100  C. For rocks along low T/P ratio, P at TMax generally exceeds 1 GPa, reaching
UHP metamorphic conditions as high as P ¼ 7 GPa. This is obviously related to new data accumulated over last 60 years since
Miyashiro presented his compilation.
(2) The Brown classification did not consider the low-P/T type metamorphism of Miyashiro.
(3) The intermediate T/P metamorphism type in the Brown dataset, despite some overlaps with the high- and intermediate-T/P
metamorphic types of the Miyashiro classification, is quite distinct in P-T space.
(4) The high-P/T and intermediate-P/T metamorphic types in the Miyashiro classification are broadly comparable with the low-T/P
and high-T/P metamorphic types in the Brown classification.

This variability of thermobaric ratios in the two datasets that also cover the extreme limits of metamorphism (in Fig. 1) allows
recognitions of four distinct metamorphic trends in the global metamorphic belts. These are: (a) low-T/P type (ratio between 150
and 375  C/GPa), (b) intermediate-T/P type (between 375 and 775  C/GPa), (c) high-T/P ratio type (between 775 and 1500  C/
GPa) and (d) very high-T/P ratio type (>1500  C/GPa).
These four types of thermo-baric ratios in metamorphic rocks, when linked with the various plate tectonic settings, provide a
fundamental basis to classify large-scale terrestrial metamorphism. Before the classification is described in detail, the following
cautionary notes may be exercised: (a) There may be overlaps of thermobaric ratios of the four metamorphic groupings, indicating a
commonality of metamorphic conditions and tectonic settings of their origin. (b) Secondly, the thermal gradient for each grouping,
inherently implied from the thermo-baric ratios does not correspond to any specific steady-state or transient and evolving thermal
condition. This suggests that the higher T sub-groups within each broad grouping do not necessarily pass through the P-T conditions
recorded in their lower-T counterparts. This has the implication that the metamorphic rocks in the same thermal grouping can be
produced in a number of tectonic settings. (c) Thirdly, the recovery of peak P-T estimates in HT/UHT metamorphic rocks is often
complicated due to the endothermic nature of melting reactions, which consume a substantial portion of the heat budget during
metamorphism, and act as effective thermal buffers (Schorn et al., 2018). One implication of this is that metapelite migmatites in
UHT metamorphic terranes may underestimate TMax in comparison to more refractory lithologies (Schorn et al., 2018). With these
caveats, the different types of terrestrial metamorphism can be classified on the basis of plate tectonic settings (four types, namely
convergent, divergent, transform plate margin and plate interior types) and agents of metamorphism (T, P and both). Being aware of
the fact that the agents are mostly not mutually exclusive, dominance of particular agent(s) for different types of metamorphism is
the criterion used.
Additionally, identification of new natural phenomena extended the scope of metamorphism beyond the classical domain of
terrestrial, particularly crustal domain. Some of these may be of local extent. These types do not fit exactly in the scheme of
classification and are discussed separately in “Other Types of Metamorphism” section.
Table 3 shows the classification scheme for regional terrestrial metamorphism adopted largely from literature incorporating later
accumulated data. Fig. 2 is a cartoon showing the positioning of the types of metamorphism in different tectonic settings and roles
of the respective variables of metamorphism.

Types of Metamorphism
Types of Metamorphism in Convergent Plate Margin
Under this category, three broad types of metamorphism, referred to as sections “Regional Thermal Metamorphism,” “Regional
Dynamic Metamorphism,” “Regional Dynamo-Thermal Metamorphism”, are recognized.

Regional Thermal Metamorphism


In this metamorphic type, T is the dominating agent with or without influence of PLithostatic (usually 4–5 kbar) and PDirected,
resulting in a very high-T/P ratio (generally >1500  C/GPa). This is distinguished and separated from contact metamorphism at the
plate interior by the regional-scale distribution of the metamorphic rocks, metamorphic isograds and zones and their common
association with convergent plate boundary tectonic settings. The type regions of this metamorphism are: Ryoke metamorphic belt
in Japan, Buchan region, the Scottish Highlands, the Acadian metamorphic region in Maine, United States, the Cooma metamor-
phic complex, Lachlan Fold Belt, SE Australia and others. The metamorphism corresponds to the andalusite-sillimanite facies series
(Buchan-type of Miyashiro, 1961) and produces schists, granofels, hornfels and migmatites, depending on the extent of TMax and
the presence or absence of deformation. Rocks in this type of metamorphism record near isobaric heating-cooling P-T paths along
hairpin clockwise or counter-clockwise geometry.
The regional thermal metamorphism, however, shows a diversity of heat sources from magmatic intrusions and associated fluid
flows in arcs and continental collision zones to mantle-scale processes due to lithospheric extension and oceanic-ridge subduction
(DeLong et al., 1979). Miyashiro (1961, 1973) attributed the very high T/P ratio at shallow crustal levels in the type Ryoke
metamorphic belt to heat advection by multiple syn-metamorphic intrusions of granitoids and broadly coeval felsic volcanism in an
arc setting. Recent studies, however, suggest a more complicated relationship between magmatic intrusion and regional thermal
Types of Metamorphism 5

Table 3 Major types of terrestrial metamorphism.

Agents T PLithostatic PDirected T and PLithostatic/Directed

Plate tectonic framework, spatial extent and geological setting


A. Convergent A.1. Regional thermal A.2. Regional dynamic A.3. Regional dynamo-thermal
plate margin A.2.1. Burial A.2.2.
(Accretionary) Dislocation
Very high T/P ratio (>1500  C/GPa) A.3.1. Orogenic A.3.2. Orogenic collision
Subduction
A.3.1.1. Low T/P A.3.2.1. Intermediate A.3.2.2. High T/P
ratio T/P ratio (between ratio (between
(<375  C/GPa) 775  C/GPa and 1500  C/GPa
A.3.1.2. 375  C/GPa) and 775  C/GPa)
Intermediate to
high T/P ratio
(700–1190  C/
GPa)
B. Divergent B.1. Regional thermal
plate margin B.1.1. B.1.2. Ocean-
Continental ridge/ocean
extension floor
Diastathather-
mal
High T/P ratio High T/P ratio
C. Transform C.1. Regional
plate margin dynamic
Dislocation
D. Plate D.1. Local thermal D.2. Regional
interior dynamic
(regional/ D.1.1. Contact D.1.2. Pyro- Burial
local) metamorphism
High T/P ratio Extremely high High to medium
(>1500  C/ T/P (! infinity) T/P ratios
GPa)

metamorphism, including recognition of temporally separate emplacement history of granitoids and possible diachronous
character of metamorphism in the Ryoke belt (Skrzypek et al., 2016).
A kyanite-sillimanite type regional thermal metamorphism has been reported from the Tananao Complex, Taiwan, where
intrusion of a granodiorite suite in greenstones in a fore-arc tectonic setting led to the development of a regional thermal aureole,
characterized by short-lived heating and rapid cooling (Wintsch et al., 2011).

Regional Dynamic Metamorphism


Under this heading, two types of metamorphism are considered-a part of burial metamorphism (discussed in section “Burial
metamorphism”) that occurs in accretionary complexes and dislocation or cataclastic metamorphism.

Burial metamorphism (accretionary)


This is discussed under section “Burial metamorphism”.

Dislocation/cataclastic metamorphism
Basically related to transformations along faults or fault zones, this is also referred to as high strain metamorphism. High strain rates
associated with deviatoric stress are responsible for producing cataclastites. Although more common in plate boundaries, it is by no
means restricted to any specific tectonic environment. At low temperatures near the surface and high strain rate, brittle deformation
produces fault zone rocks or fault gouge, which do not show recrystallization. However, the nature of deformation is dependent on
the mineral properties. At shallow depth quartz and calcite can show brittle and ductile deformation respectively. At greater depth
and with increasing temperature, rocks are affected by ductile deformation, and the characteristic product is a mylonite that shows
varying degrees of recrystallization (Wise et al., 1984). There is a complete array of transitions from purely brittle (cataclastic) to
purely ductile (mylonite) deformation in natural rocks. Accumulation of high strain rates along dislocation surfaces often lead to
frictional melting and the formation of pseudotachylite. Examples of continental-scale dislocation metamorphism are the San
Andreas fault in California, the Brevard in the Appalachian Mountains, the Moine Thrust in Scotland and the Alpine fault in New
Zealand. The above examples show that dislocation metamorphism is not confined to convergent plate margins, but also in
transform plate margins. Therefore, it is additionally listed as C.2 in Table 3.
6 Types of Metamorphism

Fig. 2 The interplay of metamorphic agents shown in the corners of a tetrahedron (adapted after Best MG (2002) Igneous and Metamorphic Petrology. 2nd edn,
Wiley-Blackwell, pp. 752) leads to classification of metamorphism into different types (numbered as 1–8) (a). These are linked with four varieties of plate tectonic
settings, namely (A) metamorphism in convergent plate margin, (B) metamorphism in divergent plate margin, (C) metamorphism in transform plate margin, and
(D) metamorphism in plate interior. The schematic sketches of some of these plate tectonic settings (adapted after Curray JR (1991) Possible greenschist
metamorphism at the base of a 22-km sedimentary section, Bay of Bengal. Geology 19(11): 1097–1100; Spear FS (1993) Metamorphic Phase Equilibria and
Pressure–Temperature—Time Paths. Monograph Series, vol. 1. Mineralogical Society of America: Washington, DC; Bosch D, Jamais M, Boudier F, Nicolas A,
Dautria JM and Agrinier P (2004) Deep and high-temperature hydrothermal circulation in the Oman ophiolite—Petrological and isotopic evidence. Journal of
Petrology 45: 1181–1208; Agard P, Yamato P, Soret M, Prigent C, Guillot S, Plunder A, Dubacq B, Chauvet A and Monié P (2016) Plate interface rheological switches
during subduction infancy: Control on slab penetration and metamorphic sole formation. Earth and Planetary Science Letters 451: 208–220, https://doi.org/10.1016/
j.epsl.2016.06.054) are presented in Figs. 2B–K. Abbreviations used: LT/HT/VHT, Low/high/very high temperature.

Regional Dynamo-Thermal Metamorphism


In this metamorphic type, both T and P are the agents of metamorphism as during metamorphism in orogenic belts. Depending on
the tectonic setting, this type of metamorphism has been further classified into orogenic subduction (Section “Orogenic subduction
type”) and orogenic collision (Section “Orogenic collision type”) types.

Orogenic subduction type


This is further subdivided into two types on the basis of T/P ratios [Low-T/P (Section “Low-T/P orogenic Subduction subtype”) and
intermediate-to high-T/P (Section “Intermediate to high T/P orogenic subduction subtype”) subtypes]

Low-T/P orogenic subduction subtype


The metamorphism of this type is characterized by transformations along a low apparent thermal gradient at metamorphic peak
(between 150 and 375  C/GPa) and corresponds to glaucophane + jadeite type facies series of Miyashiro (1961) during the cold,
mature stage of oceanic subduction or during a transition from oceanic to continental subduction (Chopin, 1984; Sobolev and
Shatsky, 1990; Liou et al., 1994; Carswell et al., 2003; de Sigyoer et al., 2004; Agard et al., 2009; Davis and Whitney, 2006; Hacker,
2006; Angiboust et al., 2009; Zhang et al., 2009; Ao and Bhowmik, 2014; O’Brien, 2018; Rajkakati et al., 2019). The metamorphism
Types of Metamorphism 7

produces high-pressure [HP, blueschists and low-temperature (LT) eclogites] and UHP metamorphic rocks within a P-T window of
1–7 GPa and 300–1000  C (Fig. 1). While oceanic and continental subduction shows a similar range of T/P ratios in the HP-UHP
metamorphic rocks (Brown and Johnson, 2019b), oceanic subduction characteristically produces lawsonite eclogite and blueschists
(Wei and Clarke, 2011). Additionally, the preserved rock record reveals statistically lower PMax in oceanic subduction-related rocks
(<2.3 GPa, Agard et al., 2009; <2.7 GPa, Erdman and Lee 2014) than during continental subduction (>2.7 GPa).
Blueschists and eclogites occur in a variety of modes in the fossil accretionary prisms: as structurally coherent slices in the Alps
(e.g., Zermatt-Saas, Sesia Lanzo zone), New Caledonia, Cyclades (e.g., Syros), as boudins or relicts in amphibolite facies rocks (e.g.,
Norway) or as “knockers” in low-T blueschists, as in the Franciscan Complex. Eclogites in these settings record TMax (generally in the
range of 520–580  C), broadly coincident in time with PMax (in the range of 2.0–3.0 Ga), hairpin type clockwise P-T paths and the
rare presence of coesite as the UHP indicator mineral. In the higher T (TMax between 750 and as high as 950  C) and P end (between
3.5 and 5.5 GPa), and still following the same low-T/P thermo-baric ratio, the eclogites lack lawsonite and contain epidote as the
hydrous phase, as in the Western Gneiss coesite province, Dabie-Sulu metamorphic belt and Kokchetav Massif of Kazakhstan
(Carswell et al., 2003; Zhang et al., 2009; Sobolev and Shatsky, 1990) or as smaller tectonic slivers such as Dora Moira (Chopin,
1984) and Tso Morari in the NW Himalayas (de Sigyoer et al., 2004). Diamond has also been reported from a few of them such as
the Dabie-Sulu belt and the Western Gneiss UHP domain. The UHP eclogites show clockwise P-T paths and show two-stage
exhumation: an initial one along a steep dP/dT gradient to lower crustal levels and a later one along a shallower dP/dT gradient to
upper crustal levels.

Intermediate to high T/P orogenic subduction subtype


This metamorphic category deals with metamorphic transformations during subduction infancy as slabs initiate descent into the
mantle. During this descent, upper crustal materials are scraped off from the downgoing slab, became accreted and welded beneath
high-temperature (HT) over plate mantle (Agard et al., 2016; Bhowmik and Ao, 2016). The welded rocks at mantle depths become
heated due to heat transfer from the hanging wall mantle and/or shear heating (e.g., Hacker, 1990), producing characteristically
warmer “metamorphic sole” rocks. The sole rocks are classified into two varieties (after Agard et al., 2016): HT sole rocks of HP
granulites and amphibolites (TMax >700–850  C at P  1.0 GPa and T/P ratio between 700  C/GPa and 1190  C/GPa) and LT sole
rocks of amphibolite to greenschist facies metasediments (TMax 550–650  C and at lower pressures).

Orogenic collision type


On the basis of T/P ratio and the P-T limits of metamorphism, the orogenic collision-type is further subdivided into two types:
Intermediate T/P type orogenic collision metamorphism (Section “Intermediate T/P type orogenic collision metamorphism”) and
high T/P type orogenic collision metamorphism (Section “High-T/P type orogenic collision metamorphism”).

Intermediate T/P type orogenic collision metamorphism


The intermediate-T/P metamorphic type is transitional with eclogites of low-T/P orogenic subduction type at the higher P end and
common granulites and UHT rocks of high-T/P type orogenic collision metamorphism at the lower P end. The metamorphism,
restricted in the kyanite stability field (within the metamorphic conditions of 750 to >1000  C and from 1.5 to >2.5 GPa),
produces medium-T (MT ) eclogites and HP granulite facies rocks (Brown, 2009) depending on the depths of burial and bulk rock
compositions (Johnson and Harley, 2012). MT eclogite and HP granulites are reported from Proterozoic to Phanerozoic collisional
orogenic belts such as the Orosirian Usagaran Orogen of Tanzania (Collins et al., 2004), Grenvillian belts of the Proto-Atlantic
region (Rivers, 2009) and the Caledonian and Variscan belts (O’Brien, 2008). In these tectonic settings, higher pressures of
metamorphism along a colder thermal gradient when compared with HT granulite and UHT metamorphism (see section “High-
T/P type orogenic collision metamorphism” below) are related with subduction burial of crustal rocks to mantle depths, in the core
regions of large orogens (e.g., Beaumont et al., 2010). P-T paths are commonly of clockwise geometry with some of these rocks
recording TMax after PMax, implying the onset of thermal relaxation due to radioactive self-heating at the end of collision. At the other
end of the tectonic spectrum, MT eclogite to HP granulite facies rocks can also form as part of continuous magmatic activity at arc
roots (e.g., in Fiordland granulites, De Paoli et al., 2009).

High-T/P type orogenic collision metamorphism


The metamorphism under this category is delimited by the fields of intermediate T/P type orogenic collision metamorphism at the
higher-P end and regional thermal metamorphism at the lower P end and between two apparent thermal gradients, 775  C/GPa and
1500  C/GPa. Broadly synonymous with the kyanite-sillimanite-type metamorphic facies series of Miyashiro (1961), the high T/P
type orogenic collision metamorphism can be classified into two subgroups: (a) the classical Barrovian type of metamorphism at
the lower T end (between 450  C in the biotite zone to 750  C in the K-feldspar-sillimanite zone) that produces schists,
amphibolites and migmatites and (b) the common granulites and UHT metamorphism at the higher T end (T ¼ 800–1100  C at
P ¼ 0.5–1.5 GPa). The type localities of the Barrovian metamorphism are the Scottish Highlands (Barrow, 1912), the Acadian of
eastern United States (Whitney et al., 1996) and the Sikkim Himalayas (Dasgupta et al., 2009). A special variety of high-T/P type
metamorphism is the UHT metamorphism, which is recognized by diagnostic mineral assemblages in high Mg-Al protoliths, such
as orthopyroxene + sillimanite + quartz, sapphirine + quartz, spinel + quartz or osumilite (Sengupta et al., 1990; Harley, 1998;
Kelsey, 2008).
Although the metamorphic type is attributed to continental collision, the extreme heat required to raise temperatures of
800–1100  C at 20–45 km depths, however, requires high heat production in the deeper levels of large and long-lived collisional
8 Types of Metamorphism

orogens with low erosion rates (Harley, 1998, 2008; Kelsey, 2008; Clark et al., 2011). While Barrovian metamorphism and common
granulites usually show clockwise P-T paths, the P-T paths in UHT granulites are variable [e.g., clockwise paths in Napier Complex
(Mitchell and Harley, 2017)] and counter-clockwise paths in the Eastern Ghats Belt (Sengupta et al., 1990). The counter-clockwise
paths in high-T/P type metamorphism are often explained by post-extension thickening and thermal decay in arc-back-arc systems
(Thompson et al., 2001; Sizova et al., 2014; Bhowmik and Chakraborty, 2017). At another extreme end of the tectonic scenario, the
high-T/P metamorphism results as a lower crustal metamorphic overprint on partially exhumed and cooled HP-UHP metamorphic
rocks (Bohemian massif, Faryad and Cuthbert, 2020). This high-T/P type of metamorphism in the Bohemian massif is attributed to
mantle-scale thermal perturbation of asthenospheric upwelling due to either slab break off or mantle delamination (Faryad and
Cuthbert, 2020).

Metamorphism in Divergent Plate Margins


Metamorphism in continental extension zone
Thinning of continental lithosphere during extensional tectonics is an important mechanism to produce low-P/high-T metamor-
phism, including granulite facies metamorphic rocks at mid-crustal levels. However, crustal extension may also occur in convergent
plate boundary processes synchronous with or succeeding a crustal thickening event as part of an orogenic collapse. Therefore, it is
often difficult to decouple metamorphism during rift-tectonics alone from an overall convergent plate boundary process. In general,
extensional tectonics lead to very heat flow at the base of the attenuated continental crust and subjacent lithospheric mantle (e.g.,
the Betic Cordillera). Platt et al. (2003) established syn-exhumational heating of orogenic lower crust (conductive heating from the
asthenosphere) and mid-crustal granulite facies metamorphism at a very high-T/P ratio (1800  C/GPa) of the crustal carapace
above the Ronda lithospheric mantle peridotite, Spain.

Ocean floor/ocean ridge metamorphism


Introduced by Miyashiro (1972) as ocean floor metamorphism (ocean ridge metamorphism of Spear, 1993), this type of regional
metamorphism occurs in mid-oceanic ridges, but found over large tracts of the ocean floor because of sea floor spreading.
As seafloor spreading is associated with very high thermal gradients (50–500  C/km), the metamorphic transformations of oceanic
crust are controlled by a combination of factors such as: (a) advective heat transfer by mafic intrusions in the oceanic crust and
asthenospheric upwelling due to extensions of the oceanic lithosphere and (b) convective circulation of large amounts of warm sea
water along fracture systems, penetrating the Moho and even the axial magma chamber. Thus, ocean floor metamorphism partially
overlaps hydrothermal metamorphism (discussed under section “Other Types of Metamorphism”). Investigations of hydrous
alteration in the interior of the oceanic crust from modern (e.g., East Pacific Rise) and fossil (e.g., Oman and Semail ophiolites)
tectonic settings have led to the recognition of three distinct temperature-dependent metamorphic alteration zones, all in the
gabbroic crust, namely (a) very high-temperature (VHT), (b) high-temperature (HT) and (c) low-temperature (LT) in a sequence
from deeper to shallower crustal levels (Bosch et al., 2004). Ocean floor metamorphism is taken to be restricted to P < 0.3 GPa and
records extremely high-T/P ratios (1500  C/GPa).

Metamorphism in Plate Interiors


Two broad types can be distinguished: contact metamorphism on a local scale and its extreme thermal manifestation, pyrometa-
morphism, and burial metamorphism on a regional scale, as defined originally. We also discuss burial metamorphism in other
tectonic settings in this section.

Contact metamorphism
This special variety of thermal metamorphism, occurring in an anorogenic tectonic setting, is found around magmatic intrusions in
sedimentary and metamorphic country rocks. The zone of mineralogical and textural changes in the country rocks, which are
brought about by the release of heat, fluids and gases by the cooling intrusions, is termed a contact aureole. Since deformation is
absent or is very limited, rocks in the aureole are typically hornfels or granofels. One characteristic feature of contact metamorphism
is that mineral assemblages in the aureole show a concentric zonation around the intrusion with the grade of metamorphism and
intensity of metamorphic recrystallization decreasing radially outward. The width and the TMax reached in the thermal aureole are a
function of the size, composition and the intrusion depth of the magmatic body and the temperature difference between the
intrusion and the country rock (Spear, 1993).
Heat flow modeling (e.g., Spear, 1993; Gerya, 2010) provides a theoretical basis to interpret contact metamorphism. These
calculations predict the varying widths of contact aureoles to be positively correlated with the sizes of intrusions. Classical contact
metamorphism around shallow granite intrusions (P  0.3 GPa) is associated with unusually high thermal gradient (>1500  C/
GPa along a horizontal direction) and produces low-P/high-T (andalusite-sillimanite type in metapelites) of metamorphism and a
TMax 600  C. Contact metamorphism of impure limestone and dolostone due to granite intrusion is generally associated with
mass transfer of fluids and elements. As silica-rich and hydrous fluids are expelled from granite intrusions, the carbonate rocks are
silicified and hydrated, producing a metasomatic calc-silicate rock, termed skarn.
The extreme limit of this variety of contact metamorphism that reached UHT granulite facies condition (T/P ratio between 1800
and 2250  C/GPa) is shown by an aureole around the troctolite-anorthosite—charnockite intrusion in Labrador (McFarlane et al.,
Types of Metamorphism 9

2003). The intrusion led to a static heating (from 700  C to 900  C) of regionally metamorphosed Tasiuyak Gneiss mid-crust
(P ¼ 4–5 kbar, Emslie and Stirling, 1993).
Emery marks another variety of extreme thermal condition of contact metamorphism. These rocks occur as fine-grained
xenoliths or screens in deeper levels of mafic plutons. Silica-poor and aluminous emeries are dominated by Fe-Ti oxides, corundum,
spinel and with minor amounts of alumino-silicates. An example of well-studied emery deposits is the Cortland Complex,
New York, for which peak P-T condition has been estimated at 7.5 kbar, 1000  C (Tracy and McLellan, 1985).

Pyrometamorphism
Pyrometamorphism (Brauns, 1911) is a special variety of contact metamorphism distinguished by extremely high thermal gradient
(T/P! infinity) at near surface conditions and marks subsolidus mineralogical transformations, often leading to partial melting of
protoliths and subsequent very rapid crystallization of anatectic melts (Grapes, 2006; Del Moro et al., 2011). Pyrometamorphism of
metasediments in the sandinite facies produces buchites, which are light-colored glassy rocks. Pyrometamorphic assemblages are
found to occur in a variety of geological settings: as ejecta, paralavas and wall-rocks of a variety of compositions from calcareous,
metapelitic, arenitic to marly sediments (Grapes, 1986; Del Moro et al., 2011), volcanic aerosols [e.g., buchite-bearing ash particles
in Popocatépetl (Mexico) volcano-plume, Obenholzner et al. (2003)] and in metapelitic to marly sediments around burning coals
[combustion metamorphism (e.g., Sokol and Volkova, 2007)].

Burial metamorphism
Burial metamorphism (Coombs, 1961) is a type of very low-grade metamorphism, which results from partial to complete
recrystallization of deeply buried sedimentary and volcanic rocks at depths and temperatures in excess of 10 km and 200–300  C
respectively and along low to medium-P/T ratios (Spear, 1993; Smulikowski et al., 2007; Bucher and Grapes, 2011). Although of
regional extent, the geological setting of burial metamorphism characteristically lacks orogenesis, synchronous magmatism and
deviatoric stress. The resultant rocks largely preserve original rock fabrics and display incomplete mineral transformations. The
metamorphic minerals are commonly restricted to veins, vesicles, interstices and alteration zones. At deeper levels of burial
metamorphism, there is circulation of hot hydrous fluids at an elevated geothermal gradient and the metamorphism overlaps
with hydrothermal metamorphism.
Although, it is complicated to accurately constrain where diagenesis ceases and burial metamorphism initiates, the increasing
intensity of metamorphism progressively stabilizes zeolites, prehnite and/or pumpellyite and sub-greenschist facies mineral
assemblages in rocks of appropriate composition. At the northern end of the Bay of Bengal, Curray (1991) predicted a greenschist
facies metamorphic condition (P  0.6 GPa, T  395–480  C) at the base of the 22 km thick Bengal fan sediments.
However, categorization of burial metamorphism in any scheme of classification is problematic. The type area of Coombs
(1961) in South Island, New Zealand was later interpreted to be an accretionary complex related to convergent plate margin, and
not anorogenic (Robinson and Merriman, 1999). This type is included in convergent plate margin metamorphism in Table 3 as
A.2.1 under the heading, burial metamorphism (accretionary). On the other hand, there is ample evidence elsewhere of anorogenic
burial of sediments and low grade metamorphism, following the original attributes of burial metamorphism (Michigan basin,
present day passive margin sediments in the Bay of Bengal and Gulf of Mexico). In Table 3, burial metamorphism is kept under the
plate interior setting in line with the original description, albeit not from the type area. This implies that burial metamorphism in
other tectonic settings needs to be distinguished to avoid confusion.
Burial metamorphism under extensional tectonic settings, such as those in the Welsh basin, southwest England basin, and
English Lake district with characteristic anticlockwise P-T-t paths (Merriman and Frey, 1999) was given the term diastathermal
metamorphism (Robinson, 1987). This term is retained in Table 3 under the category, continental extension (-
Section “Metamorphism in continental extension zone”).

Other Types of Metamorphism


Shock/Impact Metamorphism
Unrelated to tectonic setting, shock or impact metamorphism is commonly produced by the impact of meteorites and comets. These
are very short-lived events characterized by extreme temperatures (tens of thousands of degree Celsius) and pressures (several
hundred GPa), and extreme rates of strain due to the passage of shock waves. The resultant products can be called impactites. Some
of the characteristic features of impact metamorphism are shatter cones, high pressure silica polymorphs, such as stishovite and
coesite, planar deformation features in quartz and zircon, impact melt breccias and abundant pseudotachylytes, and the presence of
spherules (Glikson, 2013).

Lightning Metamorphism
Caused by lightning strike on a very local scale, this type of metamorphism is characterized by temperatures in excess of 2000  C
that result in partial melting, vaporization or chemical reduction of the affected material. If loose sand or soil is affected, lightning
produces tube-like structures of glassy (melted) material, called fulgurite (Essene and Fisher, 1986; Grapes, 2006).
10 Types of Metamorphism

Hydrothermal Metamorphism
Mineralogical and chemical changes in rocks brought about by circulating hydrothermal solutions along fractures and other
openings come under this class of metamorphism (Coombs, 1961). The source of the hot solution could be (a) circulating seawater
descending into the fractures of ocean ridge basalts (see further discussion under section “Ocean floor/ocean ridge metamor-
phism”), (b) meteoric water heated up due to geothermal gradient or in contact with shallow level intrusive, as in active geothermal
areas, and (c) fluid coming out of cooling shallow level intrusive plutons in continental areas and percolating through adjacent
rocks. The mineralogical and chemical changes may thus occur in the intrusive itself (as in ocean ridges) or in adjoining rocks (as in
geothermal areas or continental shallow intrusives). Obviously, hydrothermal metamorphism is in an open-system (metasoma-
tism) characterized by exchange of constituents and addition of heat, and produces a greater variety of mineral associations than
simple contact metamorphism, depending on the nature of the host rocks. Fluid-rock interaction processes have been studied by
many workers (Giggenbach, 1981; references cited in Bucher and Grapes, 2011), and assumed further importance because of close
genetic relationships with some ore deposits, such as those containing sulfides (Burnham, 1979).
Hydrothermal metamorphism in active geothermal fields is of special interest as it offers an opportunity to study the
transformation process in deep boreholes (Emmermann and Lauterjung 1997; Bucher and Stober, 2010). Such active cases of
hydrothermal metamorphism have been described from Japan, New Zealand, Iceland, California and elsewhere.

Asteroidal Metamorphism
This is a special variety of thermal metamorphism recorded in chondritic meteorites. The meteorites show variations of composi-
tional and textural characteristics of minerals that suggest thermal metamorphism on the primitive materials that condensed from
the solar nebula. Examples of these characteristics are devitrification and crystallization of glass, progressive attainment of
compositional homogeneity of minerals like olivine and pyroxene, and progressive blurring of the boundary between chondrules
and matrix. Van Schmus and Wood (1967) thus proposed a classification of chondrites into petrologic types from 1 to 6 in
ascending order of the degree of thermal metamorphism.
The primary cause of heating of an asteroid from which the meteorites were derived is the heat produced primarily by the decay
of the short-lived radionuclides 26Al–26Mg (Urey, 1955) and to a lesser extent of 60Fe–60Ni (Shukolyukov and Lugmair, 1993).
On the basis of cooling rate calculations and thermochronological data in meteorites, Ganguly et al. (2013, 2016) presented a
complex model of asteroid evolution. This involves break-up of the parent asteroids of the chondrites in the very early history of the
solar system and subsequent re-accretion of the disrupted materials on to other asteroids as well as to the still intact portion of a
parent asteroid.

Acknowledgments

We thank Professor G. Zhou for invitation to write this article. We are grateful to Profs. J. Ganguly and S. Chakraborty for offering
valuable criticisms and suggestions. SDG thanks INSA for support as Senior Scientist. SKB acknowledges financial support from IIT
Kharagpur in the form of a Cumulative Professional Development Allowance.

References
Agard P, Yamato P, Jolivet L, and Burov JE (2009) Exhumation of oceanic blueschists and eclogites in subduction zones: Timing and mechanisms. Earth Science Reviews 92: 53–79.
Agard P, Yamato P, Soret M, Prigent C, Guillot S, Plunder A, Dubacq B, Chauvet A, and Monié P (2016) Plate interface rheological switches during subduction infancy: Control on slab
penetration and metamorphic sole formation. Earth and Planetary Science Letters 451: 208–220. https://doi.org/10.1016/j.epsl.2016.06.054.
Angiboust S, Agard P, Jolivet L, and Beyssac O (2009) The Zermatt-Saas ophiolite: The largest (60 km wide) and deepest (c. 70–80 km) continuous slice of oceanic lithosphere
detached from a subduction zone? Terra Nova 21(3): 171–180.
Ao A and Bhowmik SK (2014) Cold subduction of the Neotethys: The metamorphic record from finely banded lawsonite and epidote blueschists and associated metabasalts of the
Nagaland Ophiolite complex, India. Journal of Metamorphic Geology 32(8): 829–860.
Barrow G (1912) On the geology of the lower Dee-side and the southern Highland border. Proceedings of the Geologist’s Association 23: 268–284.
Beaumont C, Jamieson RA, and Nguyen MH (2010) Models of large, hot orogens containing a collage of reworked and accreted terranes. Canadian Journal of Earth Sciences
47: 485–515. https://doi.org/10.1139/E10-002.
Best MG (2002) Igneous and Metamorphic Petrology, 2nd edn Wiley-Blackwell, 752.
Bhowmik SK and Ao A (2016) Subduction initiation in the Neo-Tethys: Constraints from counterclockwise P–T paths in amphibolite rocks of the Nagaland Ophiolite complex, India.
Journal of Metamorphic Geology 34(1): 17–44.
Bhowmik SK and Chakraborty S (2017) Sequential kinetic modelling: A new tool decodes pulsed tectonic patterns in early hot orogens of Earth. Earth and Planetary Science Letters
460: 171–179.
Bosch D, Jamais M, Boudier F, Nicolas A, Dautria J-M, and Agrinier P (2004) Deep and high-temperature hydrothermal circulation in the Oman ophiolite—Petrological and isotopic
evidence. Journal of Petrology 45: 1181–1208.
Brauns R (1911) Die Kristallinenschiefer des Laacher See Gebietes und ihre Umbildung zu Sanidinit. Stuttgart: Schweizerbart, 61.
Brown M (2009) Metamorphic patterns in orogenic systems and the geological record. In: Cawood PA and Kröner A (eds.) Accretionary Orogens in Space and Time. vol. 318,
pp. 37–74. London: Geological Society. Special Publications.
Brown M and Johnson T (2019a) The 51st Hallimond Lecture: Time’s arrow, time’s cycle: Granulite metamorphism and geodynamics. Mineralogical Magazine 83: 323–338. https://
doi.org/10.1180/mgm.2019.19.
Types of Metamorphism 11

Brown M and Johnson T (2019b) Metamorphism and the evolution of subduction on Earth. American Mineralogist 104: 1065–1082.
Bucher K and Grapes R (2011) Petrogenesis of Metamorphic Rocks. Heidelberg: Springer, 426.
Bucher K and Stober I (2010) Fluids in the upper continental crust. Geofluids 10: 241–253.
Burnham CW (1979) Magmas and hydrothermal fluids. In: Barnes HL (ed.) Geochemistry of Hydrothermal Ore Deposits, pp. 71–136. New York: Wiley.
Carswell DA, Brueckner HK, Cuthbert SJ, Mehta K, and O’Brien PJ (2003) The timing of stabilisation and the exhumation rate for ultra-high pressure rocks in the Western Gneiss
Region of Norway. Journal of Metamorphic Geology 21(6): 601–612. https://doi.org/10.1046/j.1525-1314.2003.00467.x.
Chopin C (1984) Coesite and pure pyrope in high-grade blueschists of the western Alps: A first record and some consequences. Contributions to Mineralogy and Petrology
86: 107–118.
Clark C, Fitzsimons ICW, Healy D, and Harley SL (2011) How does the continental crust get really hot? Elements 7: 235–240.
Collins AS, Reddy SM, Buchan C, and Mruma A (2004) Temporal constraints on Palaeoproterozoic eclogite formation and exhumation. Earth and Planetary Science Letters
224: 177–194.
Coombs DS (1961) Some recent work on the lower grades of metamorphism. The Australian Journal of Science 24: 203–215.
Curray JR (1991) Possible greenschist metamorphism at the base of a 22-km sedimentary section, Bay of Bengal. Geology 19(11): 1097–1100.
Dallwitz WB (1968) Coexisting sapphirine and quartz in granulite from Enderby Land, Antarctica. Nature 219: 476–477.
Dasgupta S, Chakraborty S, and Neogi S (2009) Petrology of an inverted Barrovian sequence of metapelites in Sikkim Himalaya: Constraints on the tectonics of inversion. American
Journal of Science 309: 43–84.
Davis PB and Whitney DL (2006) Petrogenesis of lawsonite and epidote eclogite and blueschist, Sivrihisar, Turkey. Journal of Metamorphic Geology 24: 823–849.
De Paoli MC, Clarke GL, Klepeis KA, Allibone AH, and Turnbull IM (2009) The eclogite-granulite transition: Mafic and intermediate assemblages at Breaksea Sound, New Zealand.
Journal of Petrology 50(12): 2307–2343. https://doi.org/10.1093/petrology/egp078.
de Sigyoer J, Gulliot S, and Dick P (2004) Exhumation processes of the high-pressure low-temperature Tso Morari dome in a convergent context (eastern Ladakh, NW Himalaya).
Tectonics 23: TC3003.
Del Moro S, Renzulli A, and Tribaudino M (2011) Pyrometamorphic processes at the magma–hydrothermal system interface of active volcanoes: Evidence from buchite ejecta of
Stromboli (Aeolian Islands, Italy). Journal of Petrology 52(3): 541–564.
DeLong SE, Schwarz WM, and Anderson RN (1979) Thermal effects of ridge subduction. Earth and Planetary Science Letters 44: 239–246.
Emmermann R and Lauterjung J (1997) The German continental deep drilling program KTB. Journal of Geophysical Research 102: 18179–18201.
Emslie R and Stirling J (1993) Rapakivi and related granitoids of the Nain Plutonic Suite: Geochemistry, mineral assemblages and fluid equilibria. Canadian Mineralogist 31: 821–847.
Erdman ME and Lee CTA (2014) Oceanic-and continental-type metamorphic terranes: Occurrence and exhumation mechanisms. Earth Science Reviews 139: 33–46.
Ernst WG (1988) Tectonic history of subduction zones inferred from retrograde blueschist P–T paths. Geology 16: 1081–1084.
Eskola P (1915) On the relations between the chemical and mineralogical composition in the metamorphic rocks of the Orijarvi region. Commission Géologique de Finlande Bulletin
44: 1–107 (in Finnish); 109–145 (in English).
Eskola P (1920) The mineral facies of rocks. Norsk Geolologisk Tidsskrift 6: 143–194.
Eskola P (1939) Die metamorphen Gesteine. In: Barth TFW, Correns CW, and Eskola P (eds.) Die Entstehung der Gesteine, p. 422. Berlin: Springer.
Essene EJ and Fisher DC (1986) Lightning strike fusion: Extreme reduction and metal-silicate liquid immiscibility. Science 234: 189–193.
Faryad W and Cuthbert SJ (2020) High-temperature overprint in (U)HPM rocks exhumed from subduction zones: A product of isothermal decompression or a consequence of slab
break-off (slab rollback)? Earth-Science Reviews 202: 103108. https://doi.org/10.1016/j.earscirev.2020.103108.
Ganguly J, Tirone M, Chakraborty S, and Domanik K (2013) H-chondrite parent asteroid: A multi-stage cooling, fragmentation and re-accretion history constrained by thermometric
studies, diffusion kinetic modeling and geochronological data. Geochimica et Cosmochimica Acta 105: 206–220.
Ganguly J, Tirone M, and Domanik K (2016) Cooling rates of LL, L and H chondrites and constraints on the duration of peak thermal conditions: Diffusion kinetic modeling and
implications for fragmentation of asteroids and impact resetting of petrologic types. Geochimica et Cosmochimica Acta 192: 135–148.
Gerya T (2010) Introduction to Numerical Geodynamic Modelling. Cambridge University Press, 345.
Giggenbach WF (1981) Geothermal mineral equilibria. Geochimica et Cosmochimica Acta 45: 393–410.
Glikson AY (2013) The Asteroid Impact Connection of Planetary Evolution—With Special Reference to Large Precambrian and Australian Impacts. New York: Springer, 149.
Grapes RH (1986) Melting and thermal reconstitution of politic xenoliths, Wehr volcano, East Eifel, Germany. Journal of Petrology 27: 343–396.
Grapes R (2006) Pyrometamorphism. Berlin, Heidelberg: Springer, 275.
Hacker BR (1990) Simulation of the metamorphic and deformational history of the metamorphic sole of the Oman ophiolite. Journal of Geophysical Research 95: 4895–4907.
Hacker BR (2006) Pressures and temperatures of ultrahigh-pressure metamorphism: Implications for UHP tectonics and H2O in subducting slabs. International Geology Review 48(12):
1053–1066. https://doi.org/10.2747/0020-6814.48.12.1053.
Harker A (1932) Metamorphism: A Study of Transformations of Rock Masses. Methuen & Co. Ltd, 360.
Harley SL (1998) On the occurrence and characterization of ultrahigh temperature crustal metamorphism. In: Treloar PJ and O’Brien PJ (eds.) What Drives Metamorphism and
Metamorphic Relations? pp. 81–107. London: Geological Society. Special Publications.
Harley SL (2008) Refining the P-T records of UHT crustal metamorphism. Journal of Metamorphic Geology 26: 125–154.
Johnson MRW and Harley SL (2012) Orogenesis: The Making of Mountains. Cambridge University Press, 398.
Kelsey DE (2008) On ultrahigh-temperature crustal metamorphism. Gondwana Research 13: 1–29.
Liou JG, Zhang R, and Ernst WG (1994) An introduction to ultrahigh-pressure metamorphism. Island Arc 3: 1–24.
McFarlane CRM, Carlson WD, and Connelly JN (2003) Prograde, peak, and retrograde P–T paths from aluminium in orthopyroxene: High-temperature contact metamorphism in the
aureole of the Makhavinekh Lake Pluton, Nain Plutonic Suite, Labrador. Journal of Metamorphic Geology 21: 405–423.
Merriman RJ and Frey M (1999) Patterns of very low-grade metamorphism in metapelitic rocks. In: Frey M and Robinson D (eds.) Low-Grade Metamorphism, pp. 61–107. Oxford:
Blackwell Science.
Merriman RJ and Peacor DR (1999) Very low-grade metapelite: Mineralogy, microfabric and measuring reaction progress. In: Frey M and Robinson D (eds.) Low-Grade Metamorphism,
pp. 10–60. Oxford: Blackwell Science.
Mitchell R and Harley SL (2017) Zr-in-rutile resetting in aluminosilicate ultra-high temperature granulites: Refining the record of cooling and hydration in the Napier Complex,
Antarctica. Lithos 272–273: 128–146. https://doi.org/10.1016/j.lithos.2016.11.027.
Miyashiro A (1961) Evolution of metamorphic belts. Journal of Petrology 2: 277–311.
Miyashiro A (1971) Pressure and temperature conditions and tectonic significance of regional and ocean floor metamorphism. Tectonophysics 13: 141–159.
Miyashiro A (1972) Paired and unpaired metamorphic belts. Tectonophysics 17: 241–254.
Miyashiro A (1973) Metamorphism and metamorphic belts. London: Allen & Unwin, 492.
Obenholzner JH, Schroettner H, Golob P, and Delgado H (2003) Particles from the plume of Popocatépetl volcano, Mexico—The FESEM/EDS approach. In: Oppenheimer C, Pyle D, and
Barclay J (eds.) Volcanic Degassing. vol. 213, pp. 123–148. London: Geological Society. Special Publications.
O’Brien PJ (2008) Challenges in high-pressure granulite metamorphism in the era of pseudosections: Reaction textures, compositional zoning and tectonic interpretation with examples
from the bohemian massif. Journal of Metamorphic Geology 26(2): 235–251.
O’Brien PJ (2018) Eclogites and other high-pressure rocks in the Himalaya: A review. Geological Society of London, Special Publication 483. https://doi.org/10.1144/SP483.13.
Platt JP, Argles TW, Carter A, Kelley SP, Whitehouse MJ, and Lonergan L (2003) Exhumation of the Ronda peridotite and its crustal envelope: Constraints from thermal modelling of a
P-T-time array. Journal of the Geological Society 160(5): 655–676.
12 Types of Metamorphism

Rajkakati M, Bhowmik SK, Ao A, Ireland TR, Avila J, Clarke GL, Bhandari A, and Aitchison JC (2019) Thermal history of early Jurassic eclogite facies metamorphism in the Nagaland
Ophiolite Complex, NE India: New insights into pre-cretaceous subduction channel tectonics within the Neo-Tethys. Lithos 346–347: 105166.
Rivers T (2009) The Grenville Province as a large hot long-duration collisional orogen—Insights from the spatial and thermal evolution of its orogenic fronts. In: Murphy JB, Keppie JD,
and Hynes AJ (eds.) Ancient Orogens and Modern Analogues. vol. 327, pp. 405–444. London: Geological Society. Special Publications.
Robinson D (1987) Transition from diagenesis to metamorphism in extensional and collision settings. Geology 15: 866–869.
Robinson D and Merriman RJ (1999) Low-temperature metamorphism: An overview. In: Frey M and Robinson D (eds.) Low-Grade Metamorphism, pp. 1–9. Oxford: Blackwell Science.
Schorn S, Diener JFA, Powell R, and Stüwe K (2018) Thermal buffering in the orogenic crust. Geology 46: 643–646.
Sengupta P, Dasgupta S, Bhattacharya PK, Fukuoka M, Chakraborti S, and Bhowmick S (1990) Petrotectonic imprintsin the sapphirine granulites from Anantagiri, Eastern Ghats mobile
belt, India. Journal of Petrology 31: 971–996.
Shukolyukov A and Lugmair GW (1993) Live iron-60 in the solar system. Science 259: 1138–1142.
Sizova E, Gerya T, and Brown M (2014) Contrasting styles of Phanerozoic and Precambrian continental collision. Gondwana Research 25: 522–545. https://doi.org/10.1016/j.
gr.2012.12.011.
Sizova E, Hauzenberger C, Fritz H, Faryad SW, and Gerya T (2019) Late orogenic heating of (ultra)high pressure rocks: Slab rollback vs. Slab Breakoff. Geosciences 9: 499.
Skrzypek E, Kawakami T, Hirajima T, Sakata S, Hirata T, and Ikeda T (2016) Revisiting the high temperature metamorphic field gradient of the Ryoke Belt (SW Japan): New constraints
from the Iwakuni-Yanaiarea. Lithos 260: 9–27.
Smulikowski W, Desmons J, Harte B, Sassi FP and Schmid R (2007) A Systematic Nomenclature for Metamorphic Rocks: 2. Types, Grade and Facies of Metamorphism.
Recommendations by the IUGS Subcommission on the Systematics of Metamorphic Rocks. SCMR website: Hwww.bgs.ac.uk/SCMRH.
Sobolev NV and Shatsky VS (1990) Diamond inclusions in garnets from metamorphic rocks. Nature 343: 742–746.
Sokol EV and Volkova NI (2007) Combustion metamorphic events resulting from natural coal fires. Reviews in Engineering Geology 18: 97–115. https://doi.org/10.1130/2007.4118(07).
Spear FS (1993) Metamorphic Phase Equilibria and Pressure–Temperature—Time Paths. Monograph Series, vol. 1. Washington, DC: Mineralogical Society of America.
Thompson AB, Schulmann K, Jezek J, and Tolar V (2001) Thermally softened continental extensional zones (arcs and rifts) as precursors to thickened orogenic belts. Tectonophysics
332: 115–141.
Tracy RJ and McLellan EL (1985) A natural example of the kinetic controls of compositional and textural equilibration. In: Thompson AB and Rubie DC (eds.) Metamorphic Reactions:
Kinetics, Textures, and Deformation. Advances in Physical Geochemistry, vol. 4, pp. 118–137. Springer.
Urey HC (1955) The cosmic abundance of potassium, uranium and thorium and the heat balances of the earth, the moon and Mars. Proceedings of the National Academy of Sciences
of the United States of America 42: 889–891.
Van Schmus WR and Wood JA (1967) A chemical-petrologic classification of chondritic meteorites. Geochimica et Cosmochimica Acta 31: 747–765.
Wei CJ and Clarke GL (2011) Calculated phase equilibria for MORB compositions: A reappraisal of the metamorphic evolution of lawsonite eclogite. Journal of Metamorphic Geology
29: 939–952.
Whitney DL, Mechum TA, Kuehner SM, and Dilek YR (1996) Progressive metamorphism of pelitic rocks from protolith to granulite facies, Dutchess County, New York, USA: Constraints
on the timing of fluid infiltration during regional metamorphism. Journal of Metamorphic Geology 14: 163–181. https://doi.org/10.1046/j.1525-1314.1996.05836.x.
Winter JD (2010) Principles of Igneous and Metamorphic Petrology. New York: Prentice Hall, 702.
Wintsch RP, Yang HJ, Li XH, and Tung KA (2011) Geochronologic evidence for a cold arc–continent collision: The Taiwan orogeny. Lithos 125: 236–248.
Wise DU, Dunn DE, Engelder JT, Geiser PA, Hatcher RD, Kish SA, Odon AL, and Schamel S (1984) Fault-related rocks: Suggestions for terminology. Geology 12: 391–394.
Zhang RY, Liou JG, and Ernst WG (2009) The Dabie–Sulu continental collision zone: a comprehensive review. Gondwana Research 16: 1–26.

Further Reading
Beaumont C, Jamieson RA, Nguyen MH, and Medvedev S (2004) Crustal channel flows: 1 Numerical models with applications to the tectonics of the Himalayan-Tibetan orogen.
Journal of Geophysical Research 109: B06406. https://doi.org/10.1029/2003JB002809.
Brown M (2010) Paired metamorphic belts revisited. Gondwana Research 18: 46–59. https://doi.org/10.1016/j.gr.2009.11.004.
Chopin C (2003) Ultrahigh-pressure metamorphism: Tracing continental crust into the mantle. Earth and Planetary Science Letters 212: 1–14. https://doi.org/10.1016/S0012-
821X(03)00261-9.
England PC and Thompson AB (1984) Pressure-temperature-time paths of regional metamorphism: Part I—Heat transfer during the evolution of regions of thickened continental crust.
Journal of Petrology 25: 894–928. https://doi.org/10.1093/petrology/25.4.894.
Ferry JM (ed.) (1982) Characterization of Metamorphism Through Mineral Equilibria: Reviews in Mineralogy. In: vol. 10, p. 397. Washington, DC: Mineralogical Society of America..
Ribbe PH series ed.
Okay AI (1989) Alpine-Himalayan blueschists. Annual Review of Earth and Planetary Sciences 17: 55–87.
Kerrick DM (ed.) (1991) Contact Metamorphism: Reviews in Mineralogy. In: vol. 26. Washington, DC: Mineralogical Society of America.
Kelsey DE and Hand M (2015) On ultrahigh temperature crustal metamorphism: Phase equilibria, trace element thermometry, bulk composition, heat sources, timescales and tectonic
settings. Geoscience Frontiers 6: 311–356. https://doi.org/10.1016/j.gsf.2014.09.006.
Pattison DRM (1992) Stability of andalusite and sillimanite and the Al2SiO5 triple point: Constraints from the Ballachulish aureole, Scotland. The Journal of Geology 100: 423–446.
https://doi.org/10.1086/629596.
Spear FS and Peacock SM (1989) Metamorphic Pressure-Temperature-Time Paths. Washington, DC: American Geophysical Union102.. Short Course in Geology.

You might also like