Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Controlling Factors of Metamorphism

Kurt Bucher, Mineralogy and Petrology, University of Freiburg, Freiburg, Germany


© 2020 Elsevier Inc. All rights reserved.

Metamorphic Rocks and Metamorphism 1


Supply of Thermal Energy and Increasing Temperature 2
Combined Temperature and Chemical Potential Gradients 4
Fluid Gradients, Fluid Flow, Fluid-Rock Interaction 4
Fluid-Consuming Reactions, Fluid-Rock Interaction 6
Pressure Dependence of Metamorphic Processes 6
Anisotropic Pressure, Stress and Deformation 7
Acknowledgments 8
References 8
Further Reading 9

Abbreviations (from Whitney and Evans, 2010)


An Anorthite
Bt Biotite
Cal Calcite
Chl Chlorite
Chu Clinohumite
Cld Chloritoid
Crd Cordierite
Di Diopside
Dol Dolomite
En Enstatite
Fo Forsterite
Fsp Feldspar
Gln Glaucophane
Grt Garnet
Hbl Hornblende
Jd Jadeite
Kfs K-feldspar
Ky Kyanite
Mgs Magnesite
Ms. Muscovite
Ol Olivine
Opx Orthopyroxene
Pg Paragonite
Phl Phlogopite
Qz Quartz
Sil Sillimanite
Spl Spinel
Tlc Talc
Tr Tremolite
Wo Wollastonite
Zo Zoisite

Metamorphic Rocks and Metamorphism

Metamorphic rocks are defined as rocks derived from a primary rock (protolith) by chemical and physical processes that change the
minerals and the structure of rock. The protolith can be of sedimentary, metamorphic or igneous origin. The type, modal abundance
and composition of the minerals may change. The processes typically involve a low-density fluid phase but no silicate melt. The
collective name for all processes is metamorphism. The similar processes of sediment diagenesis are normally separated from

Encyclopedia of Geology, 2nd edition https://doi.org/10.1016/B978-0-08-102908-4.00079-5 1


2 Controlling Factors of Metamorphism

Fig. 1 Garnet and hornblende grow from chlorite (and some additional minerals) in a metamorphic marlstone. Chlorite-bearing matrix has a greenish tint. Grt and
Hbl growth consumes Chl and removes it from the matrix leaving a bright Chl-free Qz + Fsp rock. With increasing progress of the Grt and Hbl forming reactions the
transport distance for the required components from the dissolving Chl increases and slows the reaction down. Diameter of 0.5 SFr coin 17 mm.

metamorphism by an artificial boundary at 200  C. Rock deformation such as brittle fracturing or ductile folding is typically
associated with metamorphism.
Fig. 1 shows a slab of a representative metamorphic rock. It would be named garnet-hornblende schist or gneiss. It has a fine-
grained greenish matrix consisting of chlorite, white mica, amphibole, plagioclase and quartz. The texture suggests that coarse garnet
and hornblende formed from matrix minerals notably from the chlorite. The green matrix has been discolored near Grt and Hbl
because they grow on the expense of chlorite. Consequently the “food” for the Grt and Hbl producing reactions must have been
transported across the discolored white zone increasing in thickness with the progress of the reaction. The reaction has apparently
been “frozen in,” which means that it did not run to completion because some regions of the rock still contain the green chlorite.
The structure of the Grt-Hbl schist shows that metamorphism produces new minerals from old minerals. The structure shown in
Fig. 1 suggests that the transformation processes may slow down and stop because the forces or imbalances driving the reactions
fade away. If the reactions would have continued, all chlorite would have disappeared and a homogeneously coarse-grained rock
with the minerals Grt, Hbl, Pl and Qz would have resulted. This group of minerals is the expected solid phase assemblage
corresponding to the suggested equilibrium assemblage of the rock. However, the Grt-Hbl schist is a typical metamorphic rock
with more than one successive assemblage preserved, with stalled reactions linking the assemblages, disequilibrium textures,
transport limited reaction progress and preserved general chemical and mechanical imbalances. Fortunately the common meta-
morphic rock is thus far more interesting than the rare rocks that may have reached complete mechanical and chemical equilibrium.
Chemical reactions in rocks follow the rules of chemical thermodynamics. The Gibbs free energy is a thermodynamic potential
and a function that depends on temperature and pressure in closed systems. The rules of thermodynamics request that any rock
arranges itself into a group of minerals that represent the lowest possible Gibbs free energy. Chemical reactions in rocks are driven
by the effort to reach a state of minimum Gibbs free energy. Therefore, changing T and P acting on a volume of rock necessarily
requires an adjustment of the rock to the new T-P conditions. T changes from adding or removing thermal energy from the volume
of rock by variations in heat flow. Changes of isotropic P result from vertical movements of the volume of rock in the rock column.
If the rock exchanges material with its surroundings (open systems) or if anisotropic pressure is acting on the rock additional
imbalances may alter the mineral assemblage and structure of metamorphic rocks. In the following sections, various causes of
metamorphism are further illustrated using specific field examples.

Supply of Thermal Energy and Increasing Temperature

First the effect of T-changes is illustrated using a serpentinite outcrop in the Alps (Trommsdorff and Evans, 1972). The serpentinite is
a hydrated version of mantle peridotite. It formed from harzburgite and mainly consists of two different Mg-silicate minerals, for
Controlling Factors of Metamorphism 3

example the serpentine mineral antigorite and forsterite. The sheet silicate antigorite contains about 12 wt% H2O. This rock may
receive thermal energy from a geological process such that its temperature increases. It is evident that the hydrous mineral antigorite
may not be heated to very high temperatures but is rather forced to give off its H2O fixed in the sheet structure. The chemical reaction
that describes the thermal decomposition of antigorite can be written as:

Antigorite ) 18Forsterite + 4Talc + 27H2 O (1)


The dehydration reaction (1) is a very characteristic type of chemical reaction in metamorphic rocks and describes the transforma-
tion of one metamorphic rock consisting of the mineral pair Atg + Fo to a second metamorphic rock consisting of Tlc + Fo (Fig. 2).
This kind of rock transformation is the essence of the processes collectively known as metamorphism. The outcrop of Fig. 2 shows
the reactant and the products of reaction (1) together indicating that the reaction did not run to completion; it has been “frozen in.”
The field example shows that chemical reactions in rocks change the type and the modal proportion of minerals making up the
rock. In the example the reaction antigorite is consumed and eventually disappears from the rock being replaced by a new mineral
talc. Forsterite has been present already in the serpentinite but more of it is produced by the reaction. The reaction involves not only
minerals but releases H2O as a fluid phase.
At the T-P acting on the rock shown in Fig. 2 the mineral pair Fo + Tlc (+H2O) has a lower Grock than the pair Fo + Atg. This can
P P
be expressed by the free energy change of the reaction △Gr for reaction (1) ( niGi of right hand side − niGi left hand side). If △Gr
is negative the reaction may progress to the right hand side and replace Atg by Tlc. △Gr is the force driving the reaction. If the force
vanishes and ultimately △Gr ¼ 0 the reaction has reached its equilibrium conditions and does not progress any further. In the
serpentinite example all three minerals can coexist with an aqueous fluid at a unique equilibrium temperature Teq at a given pressure
P. The metamorphic reaction progresses only if △Gr 6¼ 0. It requires this finite driving force implying a non-equilibrium situation to
proceed.
The antigorite breakdown reaction consumes thermal energy like any other metamorphic dehydration reaction. This energy is
typically provided by geological large scale processes external to the considered rock. In the case of the serpentinite of Fig. 2 the heat
has been supplied by an igneous intrusion. Its contact with the serpentinite is a few hundred meters away from the serpentinite
outcrop shown. The geological setting suggests that the serpentinite has first been heated above and later cooled below the
equilibrium temperature of reaction (1).

Fig. 2 Metamorphic ultramafic rock showing the breakdown of the serpentine mineral antigorite (green patches) by a chemical reaction producing talc (white
spots) and olivine (brownish elongated forsterite crystals) and releasing H2O.
4 Controlling Factors of Metamorphism

Combined Temperature and Chemical Potential Gradients

Another example of thermal metamorphism is shown in Fig. 3. The outcrop shows two different types of hornfels. The black,
extremely fine-grained hornfels was derived from shale rich in mica (clay) and chlorite. The minerals making up the hornfels
include Sil, Bt, Kfs, Crd and Qz. The black rock encloses calc-silicate hornfels with the minerals Wo, An and Di. The hornfelses
formed from an Ordovician shale - nodular limestone formation shown as an inset in Fig. 3. The limestone bands and nodules
contain calcite only. Adding heat to the formation from a cooling syenite pluton transformed the shale to the black hornfels by a
series of different dehydration reactions (Jamtveit et al. 1992a,b). The limestone nodules were converted to calcite marble but also
reacted along the contacts with the black metapelitic hornfels. For instance quartz of the black rock reacted with the calcite making
up the marble and formed wollasonite by the reaction:

Quartz + Calcite ) Wollastonite + CO2 (2)


This decarbonation reaction produces CO2 as a volatile fluid phase and consumes thermal energy for its advancement. Because the
reactants Qz and Cal reside in two different rocks the reaction involves transport of chemical components. The product wollastonite
forms a shell around the limestone nodules, which increases in thickness as the reaction progresses.
Thus the transport distance for the reactants increases during the reaction. The chemically zoned structure of the nodules is
clearly visible on Fig. 3. Reaction (2) produces wollastonite as long as △Gr < 0 and Qz and Cal can chemically interact with each
other. Transporting the two reactants to the site where Wo forms requires chemical potential gradients as driving force for
diffusional transport. With increasing transport distance, equivalent to the thickness of the wollastonite zone, the reaction proceeds
at a decreasing rate. Finally, all calcite has been used up in the nodules, which implies that the reaction stopped because one of the
reactants has been exhausted. The second example shows that chemical potential gradients causing diffusional transport are an
important cause of metamorphic changes in rocks in addition to temperature gradients causing heat flow.

Fluid Gradients, Fluid Flow, Fluid-Rock Interaction

Both reactions presented above produce a fluid component. Because hydrate minerals such as chlorite, mica and amphibole are
abundant in many sedimentary and metamorphic rocks they release H2O if thermal energy is added to the rocks as explained above
(e.g. reaction 1). The produced H2O may accumulate as a volatile low-density phase in the newly formed pore space of the rock.
Most commonly the H2O is a liquid or at T above the critical point a supercritical fluid phase. The new porosity forms because the
volume change △Vs of the solids (minerals) in for example reaction (1) is negative. The H2O fluid is internally produced and
controlled by the rock. It is a member of the phase assemblage of the rock. However the H2O fluid phase usually separates and is lost
from the host rock because of its low density and low viscosity. Consequently the reverse or back-reaction of reaction (1) becomes

Fig. 3 Banded outcrop of metapelitic hornfels with visibly zoned layers and nodules of calc-silicate rock. The rocks have been metamorphosed by the heat
released from a Permian syenite pluton. The inset shows unmetamorphosed Ordovician fossiliferous nodular limestone in a shale matrix, which together form the
source rock (protolith) of the hornfels.
Controlling Factors of Metamorphism 5

impossible and antigorite cannot form if the rock cools below the equilibrium temperature of reaction (1). The rock becomes fluid
absent.
The separated H2O fluid phase migrates through a volume of crust by different mechanisms, including Darcy flow in the fracture
network. During its migration the fluid may react with the minerals of the rocks it passes through. It may hydrate rocks consisting of
anhydrous minerals, forming, for example, micas and chlorite. The fluid is an external fluid to these rocks and the rocks can become
completely controlled by the imported fluid as in the following example (Fig. 4). The outcrop shows two different kind of marble
rocks (small frames to the left). One rock is a pure coarse grained dolomite marble, the other rock is a tremolite-calcite marble. The
Tr-marble formed from the Dol-marble by reaction (3):

5 Dolomite + 1 H2 O + 8 SiO2aq ) 1Tremolite + 3 Calcite + 7 CO2 (3)


The reaction converts pure dolomite marble to a rock containing 71 vol% tremolite and 29 vol% calcite matching precisely the
observed modal composition of the Tr-marble (Bucher-Nurminen, 1981). The reaction progresses along a reaction front that
migrates from an initial fracture through the Dol-marble to its last position about 12 cm from the central fracture. The central part of
the fracture was the conduit for an external fluid carrying dissolved SiO2aq (aq signifies aqueous silica complexes in a hydrous fluid).
The central extension fracture has later been sealed by coarse grained tremolite and calcite directly precipitated from the fluid.
Reaction (3) is often referred to as a mixed volatile reaction because both H2O and CO2 are reacting species in addition to dissolved
silica.
The conditions are such that the concentration of SiO2aq in the fracture fluid is incompatible with the dolomite of the marble.
Silica in the external fluid has been provided by Qz- and Fsp-rich rocks cut by the same fracture outside the marble outcrop. The
silica needed at the reaction front must be transported through a zone of Tr-Cal marble of increasing thickness. The advancement of
the reaction front thus depends on the diffusion velocity of SiO2aq. However, it also depends on the reaction velocity or reaction
kinetics of reaction (3). In the example shown here the reaction kinetics control the growth of the Tr-Cal marble and not the
transport of silica (Bucher, 1998). In fact, the textural details at the reaction front (inset on Fig. 3) suggest fluid absent conditions in
the Dol-marble and fluid-present conditions in the Tr-Cal marble. The Dol-Dol grain boundaries in the marble are not acting as
fluid transport locations. The reaction cuts right across the dolomite grains.
The Tr-Cal vein transects Dol-marble with some primary sedimentary beds that have been transformed to Phl-Fo-Spl marble by
thermal metamorphism. The veins are a typical example of fluid-rock interaction. The vein growth stopped because fluid flow in the
central fracture ended as a result of Tr-Cal precipitation. The process of fracture sealing is analogous to scaling and clogging of
geothermal wells or reservoirs.

Fig. 4 Outcrop of dolomite marble with Phl-Fo-Chu-Spl-Cal in originally sedimentary layers thermally metamorphosed by a cooling tonalite pluton. The marble is
cut by a fracture that served as a fluid conducting structure. The aqueous fluid contained dissolved silica from the surrounding Qz-rich rocks. The Dol-marble was
replaced with a tremolite + calcite marble by the fluid-rock interaction process. Length of pencil 15 cm.
6 Controlling Factors of Metamorphism

Fluid-Consuming Reactions, Fluid-Rock Interaction

The examples of chemical reactions in rocks (1) and (2) produce a volatile reaction product that is eventually lost by the rock.
Example (4) showed that metamorphic rocks may form from reaction of a rock with dissolved solids in externally derived fluids
(fluid-rock interaction). Reaction with externally derived volatiles is the key process in the next example (Fig. 5).
The photomicrograph shows an ultramafic mantle-derived rock originally a dunite or harzburgite. The primary olivine displays
extreme resorption textures and is replaced by euhedral enstatite und magnesite. In addition there are some small opaque grains of
Cr-spinel also belonging to the mantle stage of the rock. The texture of the rock can be readily related to the progress of the
reaction (4):

2Forsterite + 2CO2 ) Enstatite + 2Magnesite (4)


The reaction sequesters CO2 from an externally derived fluid into the carbonate mineral magnesite (Note that this natural CO2
sequestration reaction could be highly relevant for fixing antropogenic CO2 gas in environmental protection measures). However,
the dissolution of primary olivine and the reprecipitation of its components as enstatite and magnesite occur in an aqueous fluid.
The transformation mechanism is coupled to a dissolution-precipitation mechanism in an H2O fluid. H2O is the essential solvent
and polar reaction medium making reaction (4) possible although H2O does not formally appear in the stoichiometric reaction
Eq. (4).
The En + Mgs rock comes from a lens-shaped outcrop about 100 m in length, with predominantly very coarse-grained enstatite
and very little magnesite. The main reaction transforming the mantle peridotite into a pure metamorphic enstatite rock thus can be
written as: Mg2SiO4 (Fo) ) Mg2Si2O6 (En). The reaction can be balanced by Si (2 Fo ) En + MgO) or Mg (Fo + SiO2 ) En). MgO
and SiO2 can be understood as dissolved components in an aqueous fluid. High-T reactive fluids are often saline and the two
reactions can be written as follows: 2Fo + 2HCl ) En + MgCl2 + H2O and Fo + SiO2aq ) En respectively. Thus enstatite may form
by interaction of forsterite (olivine) with an externally derived saline aqueous fluid either by addition of Si or removal of Mg or a
combination of the two processes (Bucher and Stober, 2019). The two processes are accompanied by different volume changes of
the solids, one produces volume and one loses volume. Metamorphic bulk processes tend to conserve the volume of a rock body so
that the two reactions in the example above tend to progress in proportions that conserve the volume of the original dunite lens.

Pressure Dependence of Metamorphic Processes

Since the Gibbs free energy depends on T and P in closed systems, rocks can also be forced to readjust their mineral assemblage if
large scale geological processes change the pressure imposed on the rock. Simple examples for the pressure dependence of
metamorphic assemblages are isothermal phase transitions. At constant temperature and high pressure the lowest Gibbs free energy
of the composition Al2SiO5 is realized by the mineral kyanite. At low pressure the mineral sillimanite has a lower Gibbs free energy
than kyanite and thus is the more stable phase. If the pressure acting on a Ky-bearing rock is decreased Sil becomes more stable and
should form from Ky. However, Ky typically does not directly transform to Sil. Ky persists as a metastable phase. Also in this

Fig. 5 Photomicrograph of an ultramafic rock with strongly resorbed forsterite in a state of being replaced by enstatite and magnesite.
Controlling Factors of Metamorphism 7

example the Ky ) Sil transformation needs a polar reaction medium, that is the presence of an aqueous fluid to progress at a
sufficient rate.
The phase-transition quartz ) coesite in the SiO2-system is another important example of a metamorphic reaction mostly
controlled by pressure. At 600  C and about 2.6 GPa corresponding to a depth of 85 km below surface, the SiO2 polymorphs have
the same Gibbs free energy and are stable simultaneously. If a metamorphic reaction produces a separate SiO2 phase at e.g. 3 GPa
the phase has the structure of coesite. Coesite has a lower molar volume than quartz at this P-T and is thus the favored polymorph.
Coesite-bearing rocks are the type rocks of ultra-high pressure metamorphism (UHP) typically related to tectonic plate subduction.
Coesite has only been found as small inclusions mostly in garnet where it is protected from complete recrystallization to more stable
quartz during uplift and exhumation of the UHP rocks.
An important example of pressure controlled mineral assemblages can be found in mantle peridotite (Fig. 6). The garnet-
peridotite consists of olivine, orthopyroxene and coarse pyrope-rich garnet. The Ol-Opx matrix is covered by a characteristic thin
yellow weathering rind (Bucher et al., 2015). The Al-content of the mantle rock is predominantly stored in pyrope and a small
Tschermak component in Opx. Alternatively the Al can also be stored in spinel instead of garnet. Garnet-peridotite and spinel-
peridotite are linked by the heterogeneous fluid-absent mineral reaction (5):

2ðMg; FeÞ2 Si2 O6 ðOpx Þ + ðMg; FeÞAl2 O4 ðSplÞ ) ðMg; FeÞ2 SiO4 ðOlÞ + ðMg; FeÞ3 Al2 Si3 O12 ðGrtÞ (5)
The equilibrium conditions of reaction (5) depend strongly on pressure and separate high-pressure Grt-peridotite from low-
pressure Spl-peridotite. The garnet–spinel boundary is at about 1.7 GPa (at 900  C) for a typical mantle peridotite Fe-Mg
composition (1 GPa at 900  C for the pure Mg-endmember reaction 2En + Spl ) Fo + Prp). Thus spinel peridotite is the Stable
mantle rock down to depths of about 60 km. At greater depths, garnet peridotite dominates the mantle.

Anisotropic Pressure, Stress and Deformation

Pressure in the examples above refers to isotropic pressure, meaning the pressure exerted on the rock by the rock column above; this
is commonly referred to as lithostatic pressure. Rock is solid material with a shear strength resisting forces imposed by anisotropic
pressure. A geologically active crust or lithosphere is characteristically subject to a stress field. Thus metamorphism progresses
normally in the presence of stress over and above the lithostatic baseline pressure. The resulting shear stress or tensile stress acts on
the rocks undergoing chemical reaction. This has consequences mostly for the structure of the resulting metamorphic rock. The rocks
are strained and deformed, as witnessed by fractures, faults, folds, shears, foliation, and other structural features resulting from this
deformation. However, metamorphism is usually predominantly controlled by heat flow, with little associated deformation such as
in contact aureoles around igneous intrusions (Figs. 2 and 3).
Stress and associated strain may also control the rate at which the mineral transformations and chemical reactions in rocks occur.
Reaction kinetics depend strongly on the conductivity of the material for fluids, and this often increases during shearing or
extension, thus making metamorphism of rocks possible.

Fig. 6 Garnet-peridotite with coarse pyrope garnet in an olivine + orthopyroxene matrix. The peridotite outcrop is covered with a characteristic yellow weathering
rind (Bucher et al. 2015).
8 Controlling Factors of Metamorphism

Fig. 7 Three samples from one gabbro outcrop: (A) Rock sample with an igneous mineral assemblage (labradorite—augite—olivine) and structure. (B) Polished
rock slab of the same gabbro showing eclogite facies assemblages (coarse green omphacite, whitish patches with Jd - Zo - Ky + Qz pseudomorphing labradorite,
Tlc (+Cld) replacing Ol). The primary magmatic structure is preserved. (C) Polished rock slab of the above gabbro showing a structure caused by ductile deformation.
Six different hydrate minerals are present in the dark patches (Gln, Tlc, Cld, Pg, Ms., Chl). Preserved plagioclase pseudomorphs.

An example of deformation-enhanced metamorphism is shown on Fig. 7 (Bucher and Grapes 2009). The gabbroic rock of an
ophiolite complex has been exposed to about 2.5 GPa pressure and a temperature near 600  C during the subduction of oceanic
lithosphere. At this P-T point the gabbro detached from the slab and returned to the surface where it is exposed in a 2 km outcrop
today. The gabbroic structure is well-preserved, primary igneous plagioclase, augite and olivine are intact (Fig. 7A). The only
indication of metamorphism is the beginning of the augite to omphacite transformation seen from the green patches in the black
magmatic augite. The outcrop also shows the same gabbro transformed to an eclogite facies metagabbro (Fig. 7B). Green omphacite
replaced black augite; olivine changed to talc and the magmatic labradorite has been substituted by an ultrafine mixture of jadeite,
zoisite, kyanite and quartz. The metamorphic response to the high pressure seen in one rock (Fig. 7B) but not in the other (Fig. 7A)
was made possible by brittle deformation of the feldspar during subduction that increased the conductivity for H2O. Note however
that the macro-structure of the gabbro did not change. At the outcrop the same meta-gabbro also shows complete transformation to
a hydrated rock containing glaucophane, talc, chloritoid, paragonite, muscovite and chlorite in the dark patches (Fig. 7C). The fine-
grained light plagioclase pseudomorphs from Fig. 1B are preserved and did not recrystallize. The structure of the rock is
characterized by ductile deformation. Deformation created access paths for H2O during subduction zone hydration and the
production of the six different hydrate minerals listed above.
Summing up: Metamorphism results from and is controlled by imbalances in temperature, isotropic and anisotropic pressure,
chemical potentials, and fluid pressure. The local imbalances causing chemical reactions in rocks are produced by large scale
geologic processes including igneous intrusions, basin subsidence and plate movements (lithospheric subduction, nappe stacking).

Acknowledgments

Comments and suggestions by Lennart Fischer and Ingrid Stober to earlier versions of this text are gratefully acknowledged. Reto
Giere provided the field photograph of Fig. 2.

References
Bucher K (1998) Growth mechanisms of metasomatic reaction veins in dolomite marbles from the Bergell Alps. Mineralogy and Petrology 63: 151–171.
Bucher K and Grapes R (2009) The Eclogite-facies Allalin Gabbro of the Zermatt-Saas Ophiolite, Western Alps: A record of subduction zone hydration. Journal of Petrology
50: 1405–1442.
Bucher K and Stober I (2019) Interaction of mantle rocks with crustal fluids: The Sagvandites of the scandinavian caledonides. Journal of Earth Science 30(6): 1084–1094.
Bucher K, Stober I, and Müller-Sigmund H (2015) Weathering crusts on peridotite. Contributions to Mineralogy and Petrology 169/52: 1–15.
Bucher-Nurminen K (1981) The formation of metasomatic reaction veins in dolomitic marble roof pendants in the Bergell intrusion (Province Sondrio, Northern Italy). American Journal
of Science 281: 1197–1222.
Jamtveit B, Bucher-Nurminen K, and Stijfhoorn DE (1992a) Contact metamorphism of layered metasediments in the Oslo rift: I. Buffering, infiltration and the mechanisms of mass
transport. Journal of Petrology 33: 377–422.
Jamtveit B, Grorud H-F, and Bucher-Nurminen K (1992b) Contact metamorphism of layered carbonate-shale sequences in the Oslo rift. II: Migration of isotopic and reaction fronts
around cooling plutons. Earth and Planetary Science Letters 114: 131–148.
Trommsdorff V and Evans BW (1972) Progressive metamorphism of antigorite schist in the Bergell tonalite aureole (Italy). American Journal of Science 272: 487–509.
Whitney DL and Evans BW (2010) Abbreviations for names of rock-forming minerals. American Mineralogist 95: 185–187.
Controlling Factors of Metamorphism 9

Further Reading
Bucher K and Grapes R (2011) Petrogenesis of Metamorphic Rocks, 8 edn. vol. 428. Berlin, Heidelberg: Springer-Verlag. ISBN: 978-3-540-74168-8.
Hobbs BE and Ord A (2018) Structural Geology: The Mechanics of Deforming Metamorphic Rocks. Amsterdam, Netherlands: Elsevier, 680 pp.
Hollocher KT (2014) A Pictorial Guide to Metamorphic Rocks in the Field. Taylor and Francis, 326 pp.
Winter JD (2013) Principles of Igneous and Metamorphic Petrology. Harlow, UK: Pearson New International Edition, 752 pp.

You might also like