You are on page 1of 12

Metamorphism of Pelitic (Al-Rich) Rocks

Richard M Palin, Department of Earth Sciences, University of Oxford, Oxford, United Kingdom
Brendan Dyck, Department of Earth Sciences, Simon Fraser University, Burnaby, BC, Canada
© 2020 Elsevier Inc. All rights reserved.

Introduction 2
Protolith Characteristics 2
Mineralogy and Bulk Chemical Composition 2
Pre- and Very Low-Grade Metamorphic Changes in Pelitic Rocks 3
Graphical Representation of Phase Equilibria in Metapelites 3
Metamorphism of Pelitic Rocks 4
Low P/T Series (Contact) Metamorphism 5
Intermediate P/T Series (Regional/Orogenic) Metamorphism 7
Barrovian Metamorphism 8
Partial Melting and Ultrahigh-Temperature Metamorphism 9
High P/T Series (Subduction Zone) Metamorphism 10
Summary and Future Research Directions 11
References 11
Further Reading 12

Glossary
Accretionary prism A volume of sedimentary material scraped off subducting oceanic crust at a convergent plate margin that
accumulates at the edge of the overlying arc.
Authigenic a mineral that forms in situ, with this term being used almost exclusively in reference to sedimentary rocks.
Clastic A family of sedimentary rock composed of fragments (clasts) of other rocks.
Contact metamorphism Metamorphism driven by heat provided by igneous intrusions into pre-existing, colder crust.
Diagenesis The physical and chemical changes associated with transformation of a sediment to a sedimentary rock.
Facies (metamorphic) A set of metamorphic rocks for which their bulk chemical compositions produce the same (or closely
similar) mineral assemblages at given pressure and temperature conditions.
Fissile The physical characteristic of a rock meaning that it is easily split along planes of weakness.
Gneiss A high-grade, coarse-crystalline, foliated metamorphic rock with distinct mineralogical banding (gneissosity), typically
defined by feldspar.
Granofels A coarse-crystalline, unfoliated (isotropic) metamorphic rock.
Graywacke A variety of immature sandstone dominated by angular, poorly sorted clasts, including feldspar and lithic
fragments.
Hornfels A fine-crystalline, unfoliated (isotropic) metamorphic rock.
Laccolith A lens-shaped igneous intrusion that causes uplift in overlying strata in the shape of a dome.
Matrix (metamorphic) The fine-crystalline mass of minerals in a metamorphic rock that larger porphyroblasts lie within, if
they are present at all.
Metasomatism Change in the composition of a metamorphic rock as a result of the introduction or removal of chemical
constituents, typically via aqueous fluids.
Migmatite A high-grade metamorphic rock that IS comprised of a liquid (melted) component and a solid (unmelted)
component, at or near peak metamorphic conditions.
Orogenesis The process of mountain building driven by tectonic plate motion.
Paragenesis An assemblage of minerals in a metamorphic rock interpreted to be in chemical equilibrium.
Pelite A clay-rich sedimentary rock, such as shale.
Petrogenetic grid A metamorphic phase diagram that shows the pressure and temperature conditions at which discontinuous
reactions occur between minerals in a specific compositional system.
Phyllite A strongly foliated but low-grade metapelite dominated by chlorite, white mica, and quartz.
Phyllosilicate A common group of silicate minerals in which [SiO4]4− tetrahedra are arranged in sheets.
Porphyroblast A metamorphic mineral that has a significantly greater crystal size than the rest of the rock (i.e., matrix).
Prograde Related to an increase in the intensity (grade) of metamorphism due to increasing temperature and/or pressure,
typically during burial or subduction.
Protolith (metamorphic) The precursor lithology that eventually formed the metamorphic rock in question.
Pseudosection A metamorphic phase diagram that shows the equilibrium mineral, fluid, and/or melt assemblages that may
occur in a specific rock (i.e., for a fixed bulk composition), commonly as functions of pressure and temperature.

Encyclopedia of Geology, 2nd edition https://doi.org/10.1016/B978-0-08-102908-4.00081-3 1


2 Metamorphism of Pelitic (Al-Rich) Rocks

Regional metamorphism Metamorphism driven by folding, thrusting, crustal thickening, and burial of rocks over large length
scales, typically during collisional orogenesis.
Retrograde Related to a decrease in the intensity (grade) of metamorphism due to decreasing temperature and/or pressure,
typically during exhumation towards the Earth’s surface.
Schist A medium-grade, fine-crystalline, foliated metamorphic rock with distinct mineralogical banding (schistosity), typically
defined by mica.
Slate A very low-grade, very fine-crystalline, foliated metamorphic rock with a distinct fissility.
Subduction The tectonic process of transporting crust or lithosphere into the Earth’s interior at convergent plate margins.
Thermobarometry A petrological technique that determines the temperature (thermo-) and/or pressure (baro-) of
equilibration of a mineral assemblage.
Wilson Cycle A cyclical plate tectonic model of continental break-up and reassembly associated with the formation and
closure of ocean basins.

Introduction

Pelites—from the Greek pelos, meaning clay—are Al- and K-rich clastic sedimentary rocks that form in low-energy depositional
environments, such as lakes, the distal portions of continental shelves, and deep-ocean abyssal plains. During the closing stages of
the Wilson Cycle, these sediments may become entrained between accreting arcs or continental terranes, undergo significant heating
and burial, and so experience regional metamorphism along intermediate thermal gradients. Intrusion of pre-, syn- or post-
collisional magmas at shallow crustal levels may induce low-pressure/high-temperature contact metamorphism. Alternatively,
pelitic sediments may be transported to mantle depths via subduction and so be metamorphosed at high-pressure/low-temperature
conditions.
Metapelites—the metamorphosed equivalents of pelites—are a very important family of rocks, as they experience a wide variety
of mineralogical transformations that are sensitive to minor changes in pressure and temperature conditions. Systematic patterns of
metamorphic minerals that form as a function of metamorphic style and intensity (grade) thus occur in almost all orogenic belts
worldwide. As a result, many quantitative thermobarometers have been calibrated for mineral assemblages that occur in metapelitic
rocks, making them important lithologies for interpreting the tectonic evolution of geological terranes, both for the modern-day
and ancient Earth.

Protolith Characteristics

Metapelites are metamorphosed products of unconsolidated clay-rich sediments, such as mud and clay, and/or consolidated
sedimentary rocks, such as mudstone and shale. These precursor materials are mature clastic sediments that form via weathering
and erosion of continental crust. The term pelite refers in a general sense to sedimentary protoliths that are fine-grained and
dominated by clay. As the proportion of silt increases in the source rock, the metamorphosed equivalent is then commonly referred
to as a semi-pelite, although there is no formal definition of when this lithological transition occurs (Bucher and Frey, 2002). Such
rocks are common in offshore sedimentary sequences deposited on continental shelves, where mudstones and shales accumulate in
distal portions and laterally grade into more silty and sandy sediments towards the eroded source. Indeed, shales likely represent at
least 75 vol% of all sedimentary rock types that are forming on Earth today and may have been even more dominant in the
geological past (Garrels and Mackenzie, 1969).

Mineralogy and Bulk Chemical Composition

Pelitic sediments are dominated by detrital or authigenic Al- and K-rich phyllosilicates, including clay (e.g., kaolinite and smectite)
and dioctahedral mica (e.g., muscovite and paragonite). Chlorite is another common constituent that provides significant Mg and
Fe to the protolith bulk composition. Together, these sheet silicates typically comprise around 60 vol% of shales, and fine-grained
quartz constitutes a further 30 vol% (Shaw and Weaver, 1965). Minor amounts of feldspar, iron oxide, carbonate minerals, sulfide
minerals, or organic matter may be present depending on the geology of the source region that supplies these grains to the
depositional environment.
Representative bulk-rock compositions of low-grade metapelites (shale/slate) are given in Table 1 alongside greenschist- and
amphibolite-facies metamorphosed equivalents. A less argillaceous and more feldspathic graywacke bulk composition is provided
for comparison. Pelites are distinguished from other common siliciclastic rocks by their high Al2O3 and K2O and low CaO contents,
which directly reflects the predominance of sheet silicates. They are also fluid-rich, containing 5 wt% of H2O + CO2 structurally
bound within clay minerals, mica, chlorite, and carbonate, although these volatiles are lost progressively during metamorphism as
their host minerals break down via prograde reactions (Table 1).
Metamorphism of Pelitic (Al-Rich) Rocks 3

Table 1 Globally averaged bulk compositions of shale/slate, and greenschist- and amphibolite-facies metapelites (wt%) reported by Ague (1991).

Shale/slate Greenschist-facies metapelite Amphibolite-facies metapelite Graywacke

SiO2 60.34 58.33 56.25 68.85


TiO2 0.76 0.90 1.05 0.74
Al2O3 17.05 18.63 20.18 12.05
Fe2O3 7.37 8.06 9.31 4.75
MgO 2.69 3.01 3.23 2.96
MnO 0.09 0.29 0.18 0.05
CaO 1.45 1.50 1.54 0.50
Na2O 1.55 1.59 1.80 4.87
K2O 3.64 3.98 4.02 1.81
P2O5 0.14 0.18 0.19 0.06
H2O 4.25 2.94 3.02 3.07
CO2 1.05 0.32 0.18 0.08
Total 100.38 99.73 100.95 99.77

Graywacke (sample E, Table 6: Pettijohn, 1963) is shown for comparison. All iron is reported as Fe2O3.
Ague JJ (1991) Evidence for major mass transfer and volume strain during regional metamorphism of pelites. Geology 19: 855–858; Pettijohn FJ (1963) Chemical Composition of
Sandstones Excluding Carbonate and Volcanic Sands. US Geological Survey, Professional Paper 440-S, Reston, VA.

Semi-pelites and feldspar-rich sandstones, such as graywacke, have higher SiO2/Al2O3 ratios and lower total iron contents
compared to true pelites. Lime-rich mudstones occur on marine continental shelves at latitudes that allow formation of carbonate
ooze, although these are not volumetrically important in orogenic systems and their metamorphosed products show more complex
sensitivity to changes in P-T conditions owing to the control that coexisting fluid composition (H2O–CO2) has on mineral
equilibria (Livi et al., 2002). As such, this article only considers metamorphic changes in CaO-poor pelites (sensu stricto), such
as mudstone and shale.

Pre- and Very Low-Grade Metamorphic Changes in Pelitic Rocks

The transition between the sedimentary and metamorphic environment is ill-defined, as some diagenetic processes closely resemble
those that characterize metamorphic reactions at deeper levels in the crust. In metabasic rocks, metamorphism is first identified by
the stabilization of zeolite minerals, which occurs at 100–200  C in water-rich environments. However, many physical and
chemical changes in pelitic sediments that take place at this temperature range during burial are transitional, such as progressive
increases in illite “crystallinity” and vitrinite reflectance (Meunier, 2005). This makes discrimination between compacted sedimen-
tary shales and very low grade metapelites difficult without in-depth analysis of the crystallographic properties of constituent grains.
Nonetheless, pelitic sediments and sedimentary rocks undergo several physical and chemical changes during the early stages of
burial and heating, including a significant reduction in porosity, replacement of clay minerals by mica and chlorite, and
vaporization or graphitization of organic carbon. The lowest grade metapelite that can be reliably identified in the field is slate,
which is mostly comprised of quartz, muscovite, and chlorite, with or without minor biotite, hematite, pyrite, and graphite. Slate is
characteristic of the prehnite-pumpellyite metamorphic facies (Fig. 1) and has a diagnostic fissility that allows separation of
individual ‘plates’ along continuous and closely spaced cleavage planes; hence its common use as a roofing stone.

Graphical Representation of Phase Equilibria in Metapelites

The bulk chemistry of most metamorphic rocks can be almost completely described by the 12 oxides given in Table 1: however,
effective analysis and graphical representation of phase equilibria and reactions between aqueous fluid, minerals, and melt can be
difficult in such complex chemical systems. Previous studies have shown that most major reactions and mineral assemblages in
metapelites can be represented within the chemical subsystem K2O–FeO–MgO–Al2O3–SiO2–H2O (KFMASH), although this is
often extended by addition of Na2O, CaO, TiO2, Fe2O3, and/or MnO to examine the sensitivity of particular minerals to minor
components or oxidation state (e.g., White et al., 2007). Minor and trace elements do not impose noticeable controls on the
mineralogy of metapelites, except for determining which accessory minerals stabilize (e.g., monazite, zircon, allanite, apatite,
xenotime etc.) or in rare cases where an element may permit a major mineral to form far outside of its normal stability range, such as
Zn in staurolite.
A simple and reliable method of examining mineral reactions and parageneses during metamorphism is construction of ternary
(three-component) compatibility diagrams, which show equilibria that may stabilize at a given pressure-temperature (P-T )
condition in different bulk-rock compositions. Metapelite phase equilibria in KFMASH can be reduced into the AFM ternary
4 Metamorphism of Pelitic (Al-Rich) Rocks

Fig. 1 Metamorphic facies diagram showing idealized thermal gradients for contact metamorphism (low P/T ), regional metamorphism (medium P/T ), and
subduction zone (high P/T ) environments discussed in the main text, although variation on absolute P-T conditions in each scenario is expected in nature. Ab, albite;
Coe, coesite; Dia, diamond; Gr, graphite; Jd, jadeite; pre-pump, prehnite-pumpellyite; Qtz, quartz; UHT, ultrahigh temperature; WGS, wet granite solidus; Zeo,
zeolite. Modified after Palin RM, Santosh M, Cao W, Li S, Hernandez-Uribe D and Parsons AJ (2020) Secular change and the onset of plate tectonics on Earth. Earth-
Science Reviews, doi: 10.1016/j.earscirev.2020.103172. From Elsevier.

compositional system (Thompson, 1957) for display in 2-D by projection from quartz, water, and muscovite, which are assumed to
occur in the protolith and remain stable throughout metamorphism (Fig. 2A). In this case, apices of the AFM plot have the following
definitions (molar proportions of oxides): A ¼ Al2O3–3K2O, F ¼ FeO, and M ¼ MgO. For example, almandine (Fe3Al2Si3O12)
would have an A-value ¼ 1 − (3  0) ¼ 1, F-value ¼ 3, and M-value ¼ 0. At high-grade conditions, where K-feldspar is typically
present in place of muscovite, the A coordinate for the AFM projection is modified to Al2O3–K2O due to the different Al2O3:K2O
ratios in mica and feldspar (Fig. 2B). In this modified projection, minerals that do not contain potassium, such as almandine, have
the same A-F-M coordinates as with the projection from muscovite. Although not formally included in this compositional system,
ternary AFM diagrams often consider Na2O in pelitic protoliths to be hosted in the albitic component of feldspar and/or paragonite
during metamorphism, and CaO to be hosted in the grossular component of garnet and/or the anorthite component of plagioclase,
effectively extending the system to NCKFMASH (Spear, 1993).
Minerals that can stably coexist at the specified P and T condition are connected by tie lines on a compatibility diagram, which
divide compositional space into smaller domains. Bulk-rock compositions defined by proportions of the ternary diagram’s apex
components will develop the mineral assemblage defining the domain boundaries, plus any minerals that are in projection.
Example AFM diagrams showing phase equilibria in metapelites at lower amphibolite facies and granulite facies conditions are
shown in Fig. 2A and B, respectively, with the typical bulk compositional range of pelitic sediments and granitoids highlighted in
gray. More complex graphical representations of phase equilibria can be constructed if specific mineral and rock compositions can
be defined, including petrogenetic grids showing the P-T dependence of discontinuous reactions (Fig. 2C), or pseudosections, which
show the relative stability of different mineral, fluid, and melt phases as independent functions of P and T conditions (Fig. 2D). For
detailed discussion of the use of different graphical representation tools in metamorphic petrology, the reader is referred to
references provided in the “Further Reading” section of this article.

Metamorphism of Pelitic Rocks

It has been recognized for over 100 years that pelitic rocks show significant mineralogical variation during progressive changes in
metamorphic grade (e.g., Barrow, 1912). Given that different tectonic environments on Earth are characterized by distinct P/T
gradients, it is convenient to consider the major changes that occur in metapelites along the low-, medium-, and high-P/T series of
Miyashiro (1973), which broadly represent contact metamorphism around shallow to mid-crustal igneous intrusions, regional
(orogenic) metamorphism in mountain belts, and metamorphism in subduction zones, respectively (Fig. 1: Palin et al., 2020).
Metamorphism of Pelitic (Al-Rich) Rocks 5

(A) (B)
A A
Ky Sil

+ Ms + Kfs
St + Qtz + Qtz
+ H 2O + H 2O

Grt Crd
Chl

F M
pelites and Grt
granitoids pelites and
granitoids
Bt
F Bt M

(C) (D)
KFMASH + ms, qz, H2O
1.0
1.4 13 14
6 g ep pa ab
5
ep bi ilm hem 12
7
Ctd Bt
Grt

Grt

pa
0.9

mt
ilm ru hem
ab bi
Bt a
als

St

1.2

i ilm
ep ilm 10 15
Bt als l Bt als
St

ls
Ctd

chl hem

ab b
ab 8 g pl pa ab
Ch
Grt St

0.8 4

g pl
Ctd

bi ru

ep chl ab bi
±Tlc

bi ilm g st
hem 9
Pressure (GPa)

1.0 mt
Grt t

pl bi
St C ls

Pressure (GPa)
St Chl
hl
a

3 ilm
als
S

Ky
Ctd

mt bi ilm
0.7 11 mt
Sil
2

l ab
0.8 ep ab pl pa ab

tp
chl bi bi ilm

gs
0.6 17
Prl

Bt als ilm mt mt
Ky

mt

Crd Ms

t
1

m
rt

16
St Bt

ls

0.6
Bt C Chl
Chl G

ru bi ilm
rt
td

ilm
Chl G

rd a
Bt C

0.5 pl sil

bi
mt

ab
bi ilm
Ct hl

ru bi mt

st
ilm
al tz
dB

mt
l

ep chl ab
Kf s Q

pl
h
C

St C

s
Grt Ctd

als pl sil

bi
t

0.4 Bt rd
ab
s

0.4 ab bi ilm
tC
ep chl ab

Ky Gr ab bi pl mt
And chl pl
t
i ilm m

ilm mt 19
0.2 0.3 pl and 18
An rl
P
d

Sil ab bi ilm
b
l

pl bi
pl ch

And 20
ab ru

mt
ksp ilm
0.2 mt cd (–mu)
400 500 600 700 450 475 500 525 550 575 600 625 650 675 700
Temperature (°C) Temperature (°C)
Fig. 2 Example phase diagrams showing mineral equilibria for metapelites. Classical AFM compatibility diagrams calculated using a (A) muscovite projection at
amphibolite facies conditions and (B) a K-feldspar projection at granulite facies conditions showing selected minerals and representative tie lines. Dark gray region
represents typical bulk compositions for pelitic and granitoid protoliths. Tie lines connect minerals that may stably coexist at these metamorphic conditions. For
definitions of apices, see the main text. (C) A petrogenetic grid in the KFMASH compositional system and (D) a pseudosection in the MnNCKFMASHTO compositional
system for an average Dalradian metapelite from the Barrow Zones, Scotland (Palin, unpublished). (C) After Spear FS (1993) Metamorphic Phase Equilibria and
Pressure-Temperature-Time Paths, 799p. Washington, DC: Mineralogical Society of America Monograph 1.

Low P/T Series (Contact) Metamorphism

Contact metamorphism occurs due to heating, with or without burial, of rocks that lie close to a magma intrusion. It is characterized
by low P/T gradients, as strong thermal gradients between an intruding magma and adjacent country rock are best established at
shallow crustal levels. Given that the uppermost portions of the continental crust are dominantly sedimentary/metasedimentary,
metapelites are commonly found within thermal aureoles. If rocks lie structurally above or to the side of an intrusion, no change in
pressure is expected during metamorphism; temperature simply increases with proximity to the magma-country rock interface and
prograde metamorphism over time would be characterized by isobaric heating. However, if magma spreads laterally, for example
forming a thick sill or laccolith, the added load on top of pre-existing crust will cause an increase in pressure and temperature, which
may produce a counter-clockwise P-T path (Spear, 1993).
Contact metamorphism is often described using subdivisions of the low-pressure hornfels facies (Fig. 1), although the
eponymous minerals used to define this sequence, such as hornblende, do not always form in metapelites due to their Al-rich
6 Metamorphism of Pelitic (Al-Rich) Rocks

and Ca-poor bulk composition (Table 1). Nonetheless, these sub-facies have use in describing the relative intensity of metamor-
phism. A type locality for contact metamorphism in pelitic rocks is the contact aureole surrounding the Ballachulish Granite,
western Scotland (cf. Pattison, 1992). Additionally, unlike in metamafic rocks, various pressure-sensitive assemblages in metapelitic
rocks allow the approximate depth of magma emplacement to be made relatively quickly in the field, simply by identification of key
parageneses. This technique was pioneered by Carmichael (1978), who used invariant points in the KFMASH compositional system
to divide P-T space into multiple mappable pressure-sensitive facies series, or “bathozones” (cf. Fig. 3). As such there is no universal
series of isograds that appear in metapelites in contact aureoles due to the dependence of prograde mineral stability on both
pressure and bulk-rock composition.
Distal from an igneous intrusion in the shallow continental crust, pelitic sediments may have already experienced low-grade
burial metamorphism, such that slates and phyllites are common as “unmodified” country rocks. These lithologies are characterized
by fine-crystalline quartz, albite, epidote, muscovite, and chlorite and/or biotite. They also often lack a prominent metamorphic
foliation, although may preserve mineralogical banding that represent relics of a pre-existing sedimentary fabric (Fig. 4A). The first
recognizable isograd associated with contact metamorphism occurs within the upper albite–epidote facies (300–500  C; Fig. 1)
where biotite porphyroblasts form due to breakdown of chlorite and detrital K-feldspar:

Chlorite + K − feldspar ¼ biotite + muscovite + quartz + H2 O


Chloritoid may also form at similar grades in particularly Fe-rich pelites, such as in the Buchan Series of northeast Scotland.
Two diagnostic minerals that stabilize in contact metamorphic environments, but are absent from higher-pressure metamorphic
series, are cordierite and andalusite; the latter being the low-pressure polymorph of Al2SiO5. While both may be introduced at
450–500  C by the breakdown of chloritoid, typical metapelites are too Al- and Mg-poor to “see” this potential reaction. Instead,
at low-pressure (<0.25 GPa) chlorite terminally destabilizes in the presence of muscovite and quartz at hornblende-hornfels facies
conditions (530–560  C; Fig. 3) to produce both minerals simultaneously:

Chlorite + muscovite + quartz ¼ cordierite + andalusite + biotite + H2 O


In most metamorphic aureoles, both andalusite and cordierite form ovoid porphyroblasts, and if the pelite protolith is carbon-rich,
andalusite may instead form as the graphite-bearing variety chiastolite (Fig. 4B). At slightly deeper levels in the crust (P > 2.5 GPa),
chlorite breakdown forms staurolite over a narrow temperature range (560–580  C; Fig. 4C) instead of cordierite, thus defining a
useful bathograd for field investigation (Fig. 3). Prograde heating through the amphibolite-hornfels facies is also associated with an
increase in the K/Na content of white mica and the Ca content of plagioclase, and sillimanite replaces andalusite.
At higher-grade conditions (>650–700  C) of the pyroxene-hornfels facies, muscovite destabilizes to form K-feldspar and
incipient partial melting may occur at the wet granite solidus (Fig. 3). In Fe-rich bulk compositions, annite-rich biotite may
break down to produce minor garnet and additional cordierite at this temperature range (Fig. 4D), indicating how rapidly mineral
assemblages may change over short length scales in metamorphic aureoles. In more Mg-rich pelitic bulk compositions and at
shallow crustal depths (P < 2.5 kbar), phlogopite-rich biotite may break down incongruently to form orthopyroxene plus partial
melt:

Biotite + plagioclase + quartz ¼ K − feldspar + cordierite + orthopyroxene + melt

Fig. 3 Schematic pseudosection showing the pressure-dependence of various prograde facies series on the index minerals that form in pelitic bulk-rock
compositions in contact metamorphic aureoles. Modified after Pattison DRM and Vogl JJ (2005) Contrasting sequences of metapelitic mineral-assemblages in the
aureole of the tilted Nelson batholith, British Columbia: Implications for phase equilibria and pressure determination in andalusite–sillimanite-type settings. The
Canadian Mineralogist 43: 51–88.
Metamorphism of Pelitic (Al-Rich) Rocks 7

(A) (B)

Bt

Crd
Bt
20 cm 5 mm
And

(C) (D)

St St

St (Bt)

Bt Grt
Bt
And Bt
And (Bt)
5 mm 1 mm

Fig. 4 Field photographs and photomicrographs of metapelites formed during low P/T series (contact) metamorphism. All photomicrographs shown in plane
polarized light (PPL). (A) Outcrop of metapelitic hornfels in the Timeball Hill Formation, Bushveld Complex, South Africa, showing mineralogical banding formed due
to millimeter-scale variation in protolith composition. (B) Pelitic (lower) and semi-pelitic (upper) sedimentary strata metamorphosed to amphibolite-hornfels facies
conditions in the Bushveld Complex, South Africa. Andalusite (var. chiastolite), cordierite, and biotite porphyroblasts are prominent in the pelitic layer, which also
contains graphite. (C) Staurolite, andalusite, and biotite hornfels with a matrix of very fine muscovite, plagioclase, and quartz from the Bushveld Complex, South
Africa. Clear patches in the matrix are pseudomorphs after chloritoid. (D) Garnet-cordierite granofels from the innermost metamorphic aureole of the Ballachulish
Granite, western Scotland, equivalent to pyroxene-hornfels facies. Thin section shows part of a large garnet poikiloblast with an adjacent matrix of cordierite, biotite,
K-feldspar, and quartz. From OESIS/D. J. Waters.

Increasingly efficient diffusion-driven cation transport at these metamorphic grades leads to extensive recrystallization and
coarsening of fabrics, causing hornfelsic textures to trend into granofelsic textures with proximity to the intrusion, although the
duration of metamorphism represents a limiting factor on the time available for fabric development. Small intrusions such as sills
and dikes will experience short-lived thermal perturbations and produce narrow thermal (contact) metamorphic aureoles, whereas
larger intrusions such as plutons or layered intrusions will produce long-lived and wide aureoles. Quantitative thermobarometry
performed on metapelitic minerals in these contact metamorphic zones can constrain the pressure—and so depth—of magma
emplacement, and so identify post-emplacement deformation and tilting of the crust.
Extremely high-grade (>800  C) conditions in the shallow crust may occur at the very interface of a pluton with a host country
rock, although this typically requires a dry, silica-poor magma with a high emplacement temperature. This extreme variety of the
hornfels series is termed the sanidinite facies (Fig. 1) and is best recorded in metasedimentary xenoliths that have been subject to
prolonged thermal metamorphism while resident in a magma. Metapelites formed at these P-T conditions would be characterized
by cordierite, K-feldspar (sanidine), tridymite, corundum, spinel, or orthopyroxene, and commonly show microstructural evidence
for extensive partial melting.

Intermediate P/T Series (Regional/Orogenic) Metamorphism

Metamorphism caused by burial and heating of sedimentary sequences during orogenesis is characterized by clockwise P-T paths
that broadly follow intermediate geothermal gradients of 450–750  C/GPa. Such metamorphism occurs during arc-continent and
continent-continent collision (Palin et al., 2020). The classical example of regionally metamorphosed pelites is found within the
Dalradian Series, Scotland, which hosts the type locality of “Barrovian” metamorphism. This region exhibits a zonal sequence of
index minerals that form in metapelites during progressive metamorphism along an intermediate crustal geotherm from
greenschist- to amphibolite-facies P-T conditions: chlorite ! biotite ! garnet ! staurolite ! kyanite ! sillimanite (Barrow,
1912). While these isograds are named based on their diagnostic mineral’s first appearance, many of the indicator minerals remain
8 Metamorphism of Pelitic (Al-Rich) Rocks

stable at higher metamorphic grade, such as biotite. Barrovian metamorphic assemblages have been documented in many other
orogenic systems worldwide, including the Southern Appalachians, United States; the Sikkim Himalaya, India; and the Songpan-
Garzê terrane, eastern Tibet (e.g., Weller et al., 2013).

Barrovian Metamorphism
Many distinctive textural and mineralogical changes occur during Barrovian metamorphism of pelitic rocks. At sub-greenschist to
lower-greenschist facies temperatures (200–300  C), slates and phyllites dominate, and are characterized by the assemblage
muscovite + chlorite + quartz, with small amounts of accessory minerals (Fig. 5A). Phyllite exhibits a millimeter-scale foliation.
In the biotite zone (300–400  C), biotite nucleates and grows as widely spaced, dark-colored porphyroblasts that give such schists
a “spotted” appearance in hand specimen. This occurs via the same reaction noted previously for initial biotite porphyroblast

(A) (B)

Bt–Ms–Qz
Ms–Ab–Qz

Grt

Bt–Chl–Ilm
Bt Bt

Ms–Ab–Qtz
1 mm 1 mm

(C) (D)

Ms
Grt

Grt Ky Grt

Bt

St
Bt Ky
1 mm Ms–Bt–Qz 1 mm

(E) (F) Qz–Pl–Kfs


Ms–Bt
Sil

Bt
Sil
Sil
Sil
Qz–Pl–Kfs

1 mm 1 mm

Fig. 5 Photomicrographs of metapelites formed during intermediate P/T series (regional/orogenic) metamorphism shown in both plane polarized light (PPL). All
except the kyanite-zone and upper sillimanite zone examples are from the Glen Esk exposures within type locality Barrovian succession, northeast Scotland.
(A) Biotite-zone metasediment showing pelitic layers (central band), dominated by biotite, chlorite, and ilmenite, and semi-pelitic layers (top and bottom) dominated
by muscovite, albite, and quartz. (B) Garnet-zone metapelite showing preferential garnet porphyroblast growth in Al-rich domains. Note also that biotite has a
porphyroblastic form too. (C) Staurolite-zone metapelite showing a large garnet porhyroblast and several smaller staurolite porphyroblasts in a biotite and muscovite
mica-rich matrix, which shows a prominent foliation. (D) Kyanite-zone metapelite from Langtang, Nepalese Himalaya, showing porphyroblasts of kyanite and garnet
in a biotite, quartz and plagioclase matrix. (E) Lower sillimanite zone metapelite showing a knot of fibrolitic sillimanite. (F) Upper sillimanite zone metapelitic
migmatite from the Great Slave Lake shear zone, Northwest Territories, Canada, with prismatic sillimanite and quartzo-feldspathic leucosome.
Metamorphism of Pelitic (Al-Rich) Rocks 9

growth in contact metamorphic aureoles. In the garnet zone, porphyroblasts of Fe-rich garnet form due to the breakdown of chlorite
(Fig. 5B) via the following continuous reaction:

Fe − rich chlorite + muscovite + quartz ¼ Fe − rich garnet + Mg − rich chlorite + biotite + H2 O


Continued burial and heating may recrystallize early-formed foliations and lower-grade biotite is distributed throughout the matrix.
The temperature of initial garnet nucleation is around 450  C for typical metapelites but may decrease significantly in Ca- or
Mn-rich bulk composition (Tinkham et al., 2001), as recorded in some samples by grossular- or spessartine-rich
porphyroblast cores.
Both garnet and biotite persist into the staurolite zone (500–550  C), with chlorite and muscovite consumed to form staurolite
(and other products). The Al-rich composition of staurolite and relatively low mobility of Al in the metamorphic environment often
lead to grains showing poikiloblastic textures, unless they grow by direct replacement of phyllosilicate-rich domains. Staurolite-
zone metapelites typically have schistose fabrics (Fig. 5C). Continued metamorphism and burial lead to kyanite porphyroblast
growth due to staurolite breakdown at 600–620  C (Fig. 5D), thus defining the kyanite zone:

Staurolite + chlorite + muscovite + quartz ¼ kyanite + biotite + H2 O


Barrovian metamorphism requires substantial burial and crustal thickening for the prograde P-T path to enter the kyanite stability
field following staurolite breakdown (P  7 kbar); otherwise, sillimanite forms instead. Metapelites formed along hotter geother-
mal gradients, but still at the regional scale, are often referred to as Buchan Series rocks, and lack kyanite altogether. These may be
considered an intermediate series between Barrovian metamorphism, sensu stricto, and low-pressure contact metamorphism.
The highest-grade section of the Barrow Zones is marked by the appearance of sillimanite and is often divided into lower and
upper portions based on the presence of muscovite or K-feldspar. These are occasionally referred to in the literature as the first-
sillimanite zone and second-sillimanite zone, respectively. Sillimanite first stabilizes at 650  C along Barrovian geotherms, albeit
sensitive to the shape of the prograde P-T path and shallow slope of the kyanite-sillimanite transition. While sillimanite ostensibly
forms via polymorphic transformation of kyanite, Carmichael (1969) showed that this reaction proceeds non-linearly, with kyanite
directly replaced by muscovite and fibrous sillimanite overgrowing biotite elsewhere within the rock (Fig. 5E). At higher temper-
atures (680  C), muscovite and quartz break down to form K-feldspar, silicate melt (see below), and a second generation of
sillimanite, which may exhibit the prismatic habit (Fig. 5F). Upper sillimanite zone metapelites acquire a gneissose fabric and often
display evidence for in-situ partial melting, thus being best described as migmatites.

Partial Melting and Ultrahigh-Temperature Metamorphism


Upper-amphibolite facies metapelites begin to partially melt at mid-crustal pressures in the presence of water due to destabilization
of muscovite at the wet granite solidus via either of the following reactions (Patiño Douce and Harris, 1998):

Muscovite + quartz + plagioclase + H2 O ¼ aluminosilicate + melt


Muscovite + quartz + plagioclase + K − feldspar + H2 O ¼ melt
Free fluid in this case may be present along grain boundaries, causing anatexis to begin at temperatures as low as 625  C. However,
the amount of melt produced from these vapor-present reactions is very small, and so volumetrically significant melting does not
occur during prograde metamorphism until fluid-absent dehydration melting reactions begin, which liberate water that is
structurally bound within hydrous minerals. The first such reaction to be encountered in metapelites during Barrovian-type regional
metamorphism is muscovite breakdown in the upper sillimanite zone:

Muscovite + quartz + plagioclase ¼ K − feldspar + aluminosilicate + melt


At low pressures (e.g., Fig. 3), muscovite may decompose to form K-feldspar in the subsolidus regime instead. The amount of melt
produced by muscovite-dehydration is roughly equivalent to the amount of muscovite that was present below the solidus (Dyck
et al., 2020). At higher metamorphic grade, biotite is consumed via the reaction (Spear et al., 1999):

Biotite + sillimanite ¼ garnet + K − feldspar + melt


which can produce up to 40 vol% melt. This reaction is thought to be responsible for generating significant quantities of granite in
collisional orogens (e.g., Yakymchuk, 2019). This partial melt may migrate away upwards through the crust once it reaches a critical
volume threshold that permits inter-connectivity along grain boundaries, allowing it to flow (Vigneresse et al., 1996).
The major-element compositions of melts generated during high-grade metamorphism of pelites vary significantly according to
P-T conditions. Experimental petrology has shown that melts generated during relatively low-temperature, vapor-present melting
reactions along Barrovian-type geotherms are trondhjemitic (Na-rich) in composition and lack the K-rich characteristics of most
orogenic S-type magmas (Patiño Douce and Harris, 1998). Such geochemical observations agree with thermodynamic calculations
showing that only small volumes of partial melt can be generated at wet solidi in natural rocks before the reacting fluid is exhausted
(Dyck et al., 2020). Fluid-absent hydrate-breakdown melting reactions involving muscovite and biotite that accompany high-
temperature metamorphism both produce monzogranitic melt with associated peritectic ferromagnesian minerals, such as garnet.
10 Metamorphism of Pelitic (Al-Rich) Rocks

Due to the clockwise nature of P-T paths associated with collisional orogenesis, additional melt generation is expected to occur
during the decompression stage, soon after peak temperature is reached.
Ultrahigh temperature metamorphism is defined by exceeding 900  C at pressures below the sillimanite-kyanite transition, and
represents an extreme thermal event in the middle and/or lower continental crust outside of conditions expected to occur during
typical continent–continent collision. Metapelites are critical rock types used to identify this metamorphic regime, as diagnostic
mineral assemblages of such P-T conditions readily form in high-Al bulk compositions. These parageneses include orthopyroxene
+ sillimanite + quartz, sapphirine + quartz, and spinel + quartz, and metamorphic pigeonite or ternary feldspar. These assemblages
typically do not stabilize in other key families of metamorphic rocks, such as metabasites.

High P/T Series (Subduction Zone) Metamorphism

Pelitic sedimentary rocks may be metamorphosed during subduction at convergent plate margins, with slab-top temperatures
commonly defined by high-P/T geothermal gradients of 150–350  C/GPa (Palin and White, 2016). Pelagic shales are common on
oceanic abyssal plains and may be scraped off before entering a trench, forming an accretionary prism, or else subducted with the
remainder of the downgoing plate. The shallow levels of accretionary prisms are thus typically characterized by pelitic schists
metamorphosed to prehnite-pumpellyite-facies conditions, and are comprised of quartz, phengite, paragonite, chlorite (sudoite),
albite, and/or titanite. In unusually thick accretionary prisms, carpholite stabilizes in place of sudoite at pressures greater than
7–8 kbar (Fig. 6A).
During the early stages of subduction, carpholite breaks down to form chloritoid, chlorite, and/or kyanite at lower blueschist
facies conditions of 12–16 kbar and 450–550  C, and talc may be present in Mg-rich pelitic bulk compositions. Notably, key
mineralogical indicators of blueschist-facies metamorphism in subducted basalts (glaucophane and lawsonite) do not form at
equivalent P-T conditions in deeply buried metapelites due to the latter having lower bulk-rock CaO and Na2O contents (cf. Palin
and White, 2016). Chloritoid and talc break down to form Mg-rich garnet (pyrope) and kyanite at higher pressures indicative of the
eclogite facies (>20 kbar and >600  C), and Mg-rich staurolite may form in particularly Al-rich protoliths (Fig. 6B). Potassic white
mica in subducted metapelites incorporates more Fe and Mg as pressure increases, allowing extensive solid solution towards the
celadonite end member, leading to the absence or rarity of biotite at blueschist or eclogite facies conditions. The coeval

(A) (B)
Car–Gr–Qz
Qz
Grt
Ky
Ky
Car
Ms (St)
Car
Ms

3 mm

(C) (D)
Ph Qz
Grt
Hbl–Pl
Zo
Grt

Ph
Grt
Ph–Tlc
Grt

3 mm 3 mm

Fig. 6 Photomicrographs of metapelites formed during high P/T series (subduction zone) metamorphism, all shown in plane polarized light (PPL). (A) Carpholite
porphyroblasts in a medium pressure pelitic schist from the Schists Lustre, Western Alps. (B) High-pressure (eclogite-facies) metapelitic schist from the Hyllestad
region of the Western Gneiss Complex, Norway. Photomicrograph shows an aligned and fine-crystalline muscovite matrix, with coarse muscovite pseudomorphs
after staurolite. Garnet, kyanite, and quartz are also present. Field of view 15 mm. (C) Metapelitic schist from the Drage, Stadlandet region of the Western Gneiss
Complex, Norway, showing a quartz and phengite rich matrix, with margins of the latter replaced by symplectite of biotite and plagioclase. Other minerals include
minor garnet, omphacite, and zoisite. (D) Whiteschist (high-Mg metapelite) from the type locality, Dora Maira massif, Western Alps, containing skeletal pyrope-rich
garnet porphyroblasts, phengite, talc, and minor kyanite. From OESIS/D. J. Waters.
Metamorphism of Pelitic (Al-Rich) Rocks 11

incorporation of Si into the muscovite structure via this coupled substitution forms phengite, which is the characteristic phyllosi-
licate in high-pressure/low-temperature rocks (Fig. 6C). The pressure-dependence of Si incorporation into white mica forms the
basis of several barometers used to determine the tectonic evolution of deeply subducted and exhumed metasediments (e.g.,
Massonne and Schreyer, 1987). High- to ultrahigh-pressure (HP-UHP) metapelites are thus typically characterized by Mg-rich
garnet, kyanite, phengite, rutile, and either quartz or coesite depending on the absolute conditions of peak metamorphism.
If graphite is present at lower grade conditions, diamond may also form in sediments transported to extreme depths within
the Earth.
A subset of metapelites with extremely Mg-rich bulk compositions produces diagnostic mineral assemblages during subduction.
While such unusual bulk compositions are not primary features of shales and mudrocks, high-Mg mica schists and gneisses with
molar Mg/(Mg + Fe) > 0.8 occur in many orogenic terranes worldwide. The formation of these rocks is debated, and it is likely that
some may not be derived from pelitic precursors at all (e.g., evaporitic sediments), although Mg-metasomatism of shales is likely in
some cases. If subducted, these rocks form so-called whiteschists comprised principally of the diagnostic assemblage talc + kyanite,
which forms due to the breakdown of Mg-rich chlorite + quartz. Almost pure pyrope may also be present, such as in the type locality
for such rocks—the Dora Maira massif, Western Alps (Fig. 6D).

Summary and Future Research Directions

Metapelites are metamorphosed clay-rich rocks, such as mudstone or shale, that show a great diversity in mineral assemblages and
microstructures at different metamorphic facies. Typical pelitic sedimentary protoliths form in low-energy depositional environ-
ments, such as deep-sea abyssal plains. These sediments are readily incorporated into the mountain-building process and occur in
almost all geological terranes worldwide, of almost any age. The sensitivity of metapelitic mineral assemblages and the composi-
tions of key mineral constituents to changes in P and/or T conditions makes these rocks highly valuable for thermobarometry.
Contact metamorphism of pelites typically leads to the index minerals biotite, andalusite, cordierite, K-feldspar, and sillimanite
appearing with increasing proximity to an intrusive contact, and potentially orthopyroxene at very high temperatures; however, this
order of appearance is strongly dependent on pressure and bulk-rock composition. Such rocks often lack foliations and so are
hornfelsic. Metamorphism due to burial in a collisional orogen tectonic setting is characterized by a well-known isograd sequence
described by the sequential appearance of chlorite, biotite garnet, kyanite, and sillimanite, although rock-specific compositional
variation may alter this sequence. At upper amphibolite- and granulite-facies conditions, muscovite and biotite mica break down to
produce granitic partial melt. Subduction of pelitic sediments leads to stabilization of minerals such as chlorite (sudoite),
carpholite, Mg-rich garnet, talc, kyanite, chloritoid, and Si-rich white mica (phengite).
Finally, it is worthy of note that metapelites are critical rock types that can be used to address important, ongoing debate within
the petrological community. Because of the acute sensitivity of metapelitic phase equilibria to small changes in P and T conditions,
Barrovian- and Buchan-type metapelites have been the target of focused study into reaction kinetics and quantifying the degree of
disequilibrium during crustal metamorphism (Waters and Lovegrove, 2002). Overstepping of reactions is necessary to nucleate new
minerals, although some sequences of index mineral appearance predicted by thermodynamic phase equilibrium modeling conflict
with parageneses observed in many metamorphic terranes. Many research groups are actively investigating how the texture and
chemistry of minerals in metapelites can be used to constrain this effect, and to decipher whether overstepping is more common
during contact or regional metamorphism, given differences in their heating and strain rates.

References
Ague JJ (1991) Evidence for major mass transfer and volume strain during regional metamorphism of pelites. Geology 19: 855–858.
Barrow G (1912) On the geology of lower Deeside and the southern Highland border. Proceedings of the Geologists Association 23: 268–273.
Bucher K and Frey M (2002) Petrogenesis of Metamorphic Rocks. Springer Science & Business Media.
Carmichael DM (1969) On the mechanism of prograde metamorphic reactions in quartz-bearing pelitic rocks. Contributions to Mineralogy and Petrology 20: 244–267.
Carmichael DM (1978) Metamorphic bathozones and bathograds; a measure of the depth of post-metamorphic uplift and erosion on the regional scale. American Journal of Science
278: 769–797.
Dyck B, Waters DJ, St-Onge MR, and Searle MP (2020) Muscovite dehydration melting: Reaction mechanisms, microstructures, and implications for anatexis. Journal of Metamorphic
Geology 38: 1–24.
Garrels RM and Mackenzie FT (1969) Sedimentary rock types: Relative proportions as a function of geological time. Science 163: 570–571.
Livi KJ, Ferry JM, Veblen DR, Frey M, and Connolly JA (2002) Reactions and physical conditions during metamorphism of Liassic aluminous black shales and marls in Central
Switzerland. European Journal of Mineralogy 14: 647–672.
Massonne HJ and Schreyer W (1987) Phengite geobarometry based on the limiting assemblage with K-feldspar, phlogopite, and quartz. Contributions to Mineralogy and Petrology
96: 212–224.
Meunier A (2005) Diagenesis and very low-grade metamorphism. In: Clays, pp. 329–377. Springer.
Miyashiro A (1973) Metamorphism and Metamorphic Belts. 492p. Springer.
Palin RM and White RW (2016) Emergence of blueschists on Earth linked to secular changes in oceanic crust composition. Nature Geoscience 9: 60–64.
Palin RM, Santosh M, Cao W, Li S, Hernandez-Uribe D, and Parsons AJ (2020) Secular change and the onset of plate tectonics on Earth. Earth-Science Reviews. https://doi.org/
10.1016/j.earscirev.2020.103172.
Patiño Douce AE and Harris N (1998) Experimental constraints on Himalayan anatexis. Journal of Petrology 39: 689–710.
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.
12 Metamorphism of Pelitic (Al-Rich) Rocks

Pettijohn FJ (1963) Chemical composition of sandstones excluding carbonate and Volcanic Sands. In: US Geological Survey, Professional Paper 440-S, Reston, VA.
Shaw DB and Weaver CE (1965) The mineralogical composition of shales. Journal of Sedimentary Research 35: 213–222.
Spear FS (1993) Metamorphic Phase Equilibria and Pressure-Temperature-Time Paths. Washington, DC: Mineralogical Society of America Monograph. 1. 799p.
Spear FS, Kohn MJ, and Cheney JT (1999) P-T paths from anatectic pelites. Contributions to Mineralogy and Petrology 134: 17–32.
Thompson JB Jr. (1957) The graphical analysis of mineral assemblages in pelitic schists. American Mineralogist 42: 842–858.
Tinkham DK, Zuluaga CA, and Stowell HH (2001) Metapelite phase equilibria modeling in MnNCKFMASH: The effect of variable Al2O3 and MgO/(MgO + FeO) on mineral stability.
Geological Materials Research 3: 1–42.
Vigneresse JL, Barbey P, and Cuney M (1996) Rheological transitions during partial melting and crystallization with application to felsic magma segregation and transfer. Journal of
Petrology 37: 1579–1600.
Waters DJ and Lovegrove DP (2002) Assessing the extent of disequilibrium and overstepping of prograde metamorphic reactions in metapelites from the Bushveld Complex aureole,
South Africa. Journal of Metamorphic Geology 20: 135–149.
Weller OM, St-Onge MR, Waters DJ, Rayner N, Searle MP, Chung SL, Palin RM, Lee YH, and Xu X (2013) Quantifying Barrovian metamorphism in the Danba structural culmination of
eastern Tibet. Journal of Metamorphic Geology 31: 909–935.
White RW, Powell R, and Holland TJB (2007) Progress relating to calculation of partial melting equilibria for metapelites. Journal of Metamorphic Geology 25: 511–527.
Yakymchuk C (2019) On granites. Journal of the Geological Society of India 94: 9–22.

Further Reading

Pre- and Low-Grade Metamorphism


Frey M and Robinson D (eds.) (2009) Low-Grade Metamorphism. John Wiley & Sons.
Nieto F (2002) Characterization of coexisting NH4-and K-micas in very low-grade metapelites. American Mineralogist 87: 205–216.

Graphical Representation of Phase Equilibria


Powell R, Guiraud M, and White RW (2005) Truth and beauty in metamorphic phase-equilibria: Conjugate variables and phase diagrams. The Canadian Mineralogist 43: 21–33.
White RW, Powell R, and Baldwin JA (2008) Calculated phase equilibria involving chemical potentials to investigate the textural evolution of metamorphic rocks. Journal of
Metamorphic Geology 26: 181–198.
Xu G, Will TM, and Powell R (1994) A calculated petrogenetic grid for the system K2O-FeO-MgO-Al2O3-SiO2-H2O, with particular reference to contact-metamorphosed pelites. Journal
of Metamorphic Geology 12: 99–119.

Low P/T Series (Contact) Metamorphism


Loomis TP (1972) Contact metamorphism of pelitic rock by the Ronda ultramafic intrusion, southern Spain. Geological Society of America Bulletin 83: 2449–2474.
Pattison DRM and Tinkham DK (2009) Interplay between equilibrium and kinetics in prograde metamorphism of pelites: An example from the Nelson aureole, British Columbia. Journal
of Metamorphic Geology 27: 249–279.
Symmes GH and Ferry JM (1995) Metamorphism, fluid flow and partial melting in pelitic rocks from the Onawa contact aureole, Central Maine, USA. Journal of Petrology 36: 587–612.

Intermediate P/T Series (Regional/Orogenic) Metamorphism


Le Breton N and Thompson AB (1988) Fluid-absent (dehydration) melting of biotite in metapelites in the early stages of crustal anatexis. Contributions to Mineralogy and Petrology
99: 226–237.
Palin RM, Searle MP, St-Onge MR, Waters DJ, Roberts NMW, Horstwood MSA, Parrish RR, Weller OM, Chen S, and Yang J (2014) Monazite geochronology and petrology of kyanite-
and sillimanite-grade migmatites from the northwestern flank of the eastern Himalayan syntaxis. Gondwana Research 26: 323–347.
Treloar PJ, Palin RM, and Searle MP (2019) Towards resolving the metamorphic enigma of the Indian plate in the NW Himalaya of Pakistan. Geological Society of London, Special
Publication 483: 255–279.

High P/T Series (Subduction Zone) Metamorphism


López-Carmona A, Pitra P, and Abati J (2013) Blueschist-facies metapelites from the Malpica–Tui Unit (NW Iberian Massif ): Phase equilibria modelling and H2O and Fe2O3 influence in
high-pressure assemblages. Journal of Metamorphic Geology 31: 263–280.
Maldonado R, Ortega-Gutiérrez F, and Hernández-Uribe D (2016) Garnet–chloritoid–paragonite metapelite from the Chuacús Complex (Central Guatemala): New evidence for
continental subduction in the North America–Caribbean plate boundary. European Journal of Mineralogy 28: 1169–1186.
Proyer A (2003) The preservation of high-pressure rocks during exhumation: Metagranites and metapelites. Lithos 70: 183–194.
Wei C and Powell R (2003) Phase relations in high-pressure metapelites in the system KFMASH (K2O–FeO–MgO–Al2O3–SiO2–H2O) with application to natural rocks. Contributions to
Mineralogy and Petrology 145: 301–315.

Relevant Websites
https://www.virtualmicroscope.org/—A virtual collection of high-resolution thin section scans of many rock types, including metapelites, that can be viewed online.
https://bgs.ac.uk/SCMR/—The homepage of the Subcommission on the Systematics of Metamorphic Rocks (SCMR); a branch of the IUGS Commission on the Systematics in Petrology
(CSP). The SCMR provides a unified nomenclature scheme and recommended definitions for metamorphic rocks to aid international understanding.
https://hpxeosandthermocalc.org/—The homepage for the petrological modeling program THERMOCALC and associated thermodynamic database, which are commonly used for
construction of phase diagrams and performing thermobarometry in metapelitic rocks.
http://www.earth.ox.ac.uk/oesis/teaching/metageol/index.html—A public-access repository of petrological images and information compiled by Prof. David Waters, University of
Oxford, UK. Many thin section images of metapelites from different geological environments are available.

You might also like