You are on page 1of 241

Journal Pre-proof

Crustal melting and suprasolidus phase equilibria: From first


principles to the state-of-the-art

Tim Johnson, Chris Yakymchuk, Michael Brown

PII: S0012-8252(21)00279-8
DOI: https://doi.org/10.1016/j.earscirev.2021.103778
Reference: EARTH 103778

To appear in: Earth-Science Reviews

Received date: 25 January 2021


Revised date: 27 July 2021
Accepted date: 18 August 2021

Please cite this article as: T. Johnson, C. Yakymchuk and M. Brown, Crustal melting and
suprasolidus phase equilibria: From first principles to the state-of-the-art, Earth-Science
Reviews (2018), https://doi.org/10.1016/j.earscirev.2021.103778

This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. Please note that, during the production
process, errors may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

© 2018 © 2021 Published by Elsevier B.V.


Journal Pre-proof

Crustal melting and suprasolidus phase equilibria: From first principles to the

state-of-the-art

Tim Johnsona,* tim.johnson@curtin.edu.au, Chris Yakymchukb, Michael Brownc

a
School of Earth and Planetary Sciences, The Institute for Geoscience Research, Curtin University, GPO Box

U1987, Perth, WA 6845, Australia

f
b
Department of Earth and Environmental Sciences, University of Waterloo, Waterloo, ON, Canada N2L 3G1

oo
c
Laboratory for Crustal Petrology, Department of Geology, University of Maryland, College Park, Maryland

20742, USA
*
Corresponding author.
pr
e-
Abstract
Pr

Partial melting is the fundamental process by which juvenile crust was produced from the

mantle and subsequently reworked to become the stable, compositionally-differentiated


al

continents on which we live and which host most of the elemental resources required by our
n

modern technological society. Irreversible differentiation of the continental crust occurs


ur

principally through the production, segregation and migration of silica-rich (felsic) melts
Jo

from deeper source rocks to shallower sinks where they erupt or, more commonly, crystallise

as granite sensu lato. Here we provide for both novices and professionals a comprehensive,

but accessible, account of the processes involved in crustal melting and suprasolidus phase

equilibria from first principles to the forefront of modern research. To begin we introduce the

tectono-metamorphic context for crustal melting before considering the evidence at outcrop

and in thin section for partial melting, and briefly reviewing melt extraction from crustal

rocks. As a prerequisite to understanding the physicochemical basis for crustal melting, we

summarize the essential thermodynamics that underpin calculated phase equilibria,

distinguish different types of melting reaction, and review the requirements for, methodology
Journal Pre-proof

behind, and limitations of a phase equilibrium modelling approach based on equilibrium

thermodynamics. We explain the various types of phase diagram used to investigate partial

melting and assess open versus closed system processes, including internal and external

buffering of H2O. Those crustal sources that partially melt to produce granite are considered

in detail, namely basic rocks such as basalts and gabbros, clastic sedimentary rocks such as

greywackes, siltstones and mudstones (pelites), and granites themselves. We concentrate

mainly on intracrustal partial melting in convergent-margin settings, and on anatexis during

exhumation of deeply-subducted continental crust from mantle depths. We discuss the

f
oo
behaviour of trace elements and accessory minerals during melting, and consider the

implications for isotope geochemistry. To close we include a brief summary of some of the
pr
important points and offer a few suggestions for future lines of research on crustal melting.
e-
Pr

Keywords: Granite, Melting, Migmatite, Open vs closed systems, Suprasolidus phase

equilibria, including P–T–X pseudosections, Trace elements and accessory minerals


al

1. Introduction
n
ur

In this article we review the current understanding of crustal melting (also known as
Jo

anatexis) within the context of phase equilibria. Melting is the process of transforming a solid

substance into a liquid, usually, but not exclusively, by the addition of heat. Melting is

generally partial rather than complete, producing coexisting liquid (melt) and one or more

solids (minerals). Partial melting is the principal mechanism driving irreversible

differentiation of the crust, mostly through the production, segregation and migration of

silica-rich (felsic) melts from their source rocks to shallower levels, where they may erupt or,

more commonly, crystallise as granite sensu lato (hereafter simply granite). Thus, at one

extreme of scale, partial melting is responsible for the formation of the continents that we live
Journal Pre-proof

on, and at the other, for the generation of rocks such as obsidian that our earliest ancestors

used for tools. Furthermore, partial melting and crustal differentiation has resulted in the

concentration of the many elements and compounds that have fundamentally shaped the

development of our society, from the discovery and smelting of tin and copper ores that

ushered in the Bronze Age to the present-day demand for the rare metals without which our

smart phones could not function.

Our intention is to provide a comprehensive but accessible reference for those who

wish to understand, or teach, a phase equilibrium approach to crustal melting from first

f
oo
principles to the state-of-the-art. Following this introduction, we provide a context for crustal

melting by briefly reviewing: (i) the realm of crustal metamorphism, particularly in relation
pr
to metamorphic facies, types of metamorphism in terms of varying thermobaric ratio (T/P),
e-
and sources of heat; (ii) the nomenclature and characteristics of partially-melted rocks at
Pr

outcrop and under the microscope; and, (iii) the process of melt extraction from a partially

molten crustal source. For reasons of space, we restrict our review to those crustal rock types
al

that, through partial melting, are responsible for the production of granitic rocks, which we
n

subdivide into: (i) igneous protoliths, with an emphasis on metabasic rocks, and; (ii) clastic
ur

sedimentary protoliths, with an emphasis on metapelites and metagreywackes. Although their


Jo

importance is undoubtedly underappreciated, we do not consider partial melting of

carbonates (Ferrero et al., 2016; Lentz, 1999), sulphides (Frost et al., 2002; Stevens et al.,

2005; Tomkins et al., 2007), or ultramafic rocks such as peridotites, pyroxenites and

komatiites.

We use 'fluid' or ‘aqueous fluid’ to refer to a H2O-dominated vapour phase, and ‘melt’ and

‘liquid’ interchangeably to mean a polymerised silicate melt phase. In the appendix we

provide a glossary of some of the more important terms used in section 2 that relate to crustal
Journal Pre-proof

melting––key terms used elsewhere in the paper are explained in the main text. Phase

abbreviations follow Whitney and Evans (2010).

2. Metamorphism and crustal melting

2.1. Crustal metamorphism

f
oo
During prograde regional metamorphism, as temperature, T (and commonly

pressure, P), increase, the mineral assemblage in rocks generally equilibrates continuously
pr
as fluid or melt is generated. However, fluid and melt are less viscous and have a lower
e-
density than their metamorphic source rocks, so that as they become over pressured they
Pr

generally segregate and migrate to higher crustal levels. Thus, during the retrograde

evolution, as T (and commonly P) decrease, mineral assemblages generally undergo much


al

less change because fluid or melt is scarce or absent. An equilibrium mineral assemblage
n

will tend to be preserved once fluid is lost (before or around the peak T of metamorphism) or
ur

as the last dregs of trapped melt crystallize (after the peak T, during cooling and crossing of
Jo

the solidus), reflecting the P–T conditions at that point. This is the basic logic underlying the

concepts of the metamorphic peak and metamorphic facies.

Metamorphic facies are defined by commonly-occurring sets of mineral assemblages

developed in rocks of different compositions that recur in space and time (Eskola, 1920; Fyfe

et al., 1958); that is, metamorphic facies represent the overlapping stability fields in P–T

space for these sets of mineral assemblages. The common metamorphic facies, which are

shown in Fig. 1a, provide useful shorthand, for example in classifying the P–T ranges of

different sets of mineral assemblages and in a comparison of processes under different ranges
Journal Pre-proof

of P–T conditions, such as anatexis under granulite facies versus eclogite facies conditions

(Fig. 1a). In addition, they form the basis for the metamorphic facies series and the

classification of metamorphic belts by Miyashiro (1961).

On the contemporary Earth, high-grade metamorphism and crustal melting are

predominantly associated with convergent plate boundary processes. Therefore, with the

increase in quantitative P–T data now available, it is convenient to use the thermobaric

ratios (T/P) retrieved from the peak metamorphic mineral assemblages, since these vary both

spatially and temporally during orogenesis. Using T/P, metamorphism has been classified

f
oo
into three types, each with tectonic significance (Fig. 1b; (Brown, 2007; Brown and Johnson,

2019b; Brown and Johnson, 2019c): high T/P metamorphism with thermobaric ratios >775
pr
°C/GPa, intermediate T/P metamorphism with thermobaric ratios between 775 and 375
e-
°C/GPa, and low T/P metamorphism with thermobaric ratios <375 °C/GPa. In a simple one-
Pr

sided subduction-to-collision orogen, each type of metamorphism is typically associated with

a particular tectonic setting (Brown, 2010b). Low T/P metamorphism, which is characterised
al

by blueschists, low-temperature (quartz) eclogites and ultrahigh pressure (coesite/diamond)


n

eclogites, forms in subduction channels associated with accretionary orogens and during
ur

partial subduction of passive continental margins. Intermediate T/P metamorphism, which is


Jo

characterised by high-pressure granulites and medium-to-high temperature eclogites, is

associated with mountain building, particularly during collision where continental crust is

thickened. High T/P metamorphism, which is characterised by upper amphibolite and

granulite facies rocks, and ultrahigh temperature (UHT) granulites, occurs in orogenic

hinterlands (backarcs and plateaus) and in regions of lithosphere extension (Cipar et al.,

2020).

In crustal rocks containing hydrous minerals or a pore fluid, melting takes place at

the solidus in the temperature range 650–700 °C and higher, conditions that may be achieved
Journal Pre-proof

in all three types of metamorphism (Brown and Johnson, 2018, 2019a, b). Thus, we may refer

to subsolidus rocks, meaning protoliths that record peak P–T conditions below the solidus,

and suprasolidus rocks, referring to formerly melt-bearing lithologies that record peak P–T

conditions above the solidus.

The irreversible process of crustal differentiation by partial melting involves

feedback between heat flow, chemical reaction, melt flow and elastic–plastic deformation in

orogens (Brown, 2010c; Clark et al., 2011). In particular, partial melting requires high

temperatures and a large amount of energy due to the latent heat of fusion of crustal rocks

f
oo
(Bea, 2012; Moyen, 2020b; Schorn et al., 2018; Stüwe, 1995). Indeed, since the rates of melt

segregation, ascent and emplacement in the upper crust are all relatively fast, occurring in
pr
<<1 Myr, but the rate of prograde heating during regional metamorphism is generally
e-
relatively slow, taking >>1 Myr, it is heat flow that determines the overall rate of the crustal
Pr

melting process (Brown, 2013; Harris et al., 2000).

During intermediate and high T/P metamorphism, regional-scale partial melting may
al

be driven by thickening of continental crust, particularly if enhanced by a high concentration


n

of heat producing elements (principally K, Th and U), as well as by magmatic underplating


ur

and by the upwelling of hot asthenosphere during lithospheric extension (Clark et al., 2011;
Jo

Harley, 2016; Jamieson et al., 1998; Kelsey and Hand, 2015). At present, crustal melting is

occurring beneath orogenic plateaux associated with continental arcs, such the Andes (Yuan

et al., 2000), and in subduction-to-collision orogens and their hinterlands, in particular the

Himalaya and the Tibetan plateau (Nelson et al., 1996; Unsworth et al., 2005). During low

T/P metamorphism, regional-scale partial melting has been linked with the subduction and

exhumation of continental crust to and from mantle depths (ultrahigh-pressure

metamorphism) (Labrousse et al., 2011) (see section 8).


Journal Pre-proof

At a more localized scale, partial melting may occur in contact metamorphic aureoles

associated with large mafic/ultramafic igneous complexes, such as the Bushveld (Johnson et

al., 2003b, 2010) and the Duluth (Ripley and Alawi, 1988) complexes, and around granite

plutons (Marchildon and Brown, 2002) and volcanic feeders (Holness, 1999). Melting may

also be associated with: (1) the influx of hydrous fluid into crust at temperatures above, but

close to, the H2O-saturated (wet) solidus (Collins et al., 2020b; Pourteau et al., 2020; Tafur

and Diener, 2020; Weinberg and Hasalová, 2015); (2) frictional heating (viscous dissipation)

associated with shear zones (Nabelek et al., 2010) and faults (Zhong et al., 2021); (3)

f
oo
subducting lithosphere younger than 25 Myr (Defant and Drummond, 1990); and, (4)

meteorite impacts (Gibson, 2002), particularly during the Hadean and Archean eons (Johnson

et al., 2018).
pr
e-
Pr

2.2. Outcrop-scale features and nomenclature for partially-melted rocks


al

If the temperature due to metamorphism becomes sufficiently high, most crustal rocks
n

begin to melt (a process called partial melting or anatexis). Before partial melting occurs,
ur

source rocks are referred to as protoliths, some of which are more fertile than others,
Jo

producing more melt at the same P–T conditions. Source rocks from which melt has drained

are referred to as residual. Criteria for the recognition of partial melting are discussed by

Sawyer (1999).

At outcrop, the presence of irregular patches, lenses and/or veins of

quartzofeldspathic material with a granitic composition and texture commonly (but not

always) provides evidence for partial melting (Brown et al., 1995; Sawyer, 1999; Sawyer,

2008; Sawyer and Barnes, 1988; Sawyer and Robin, 1986). Since melting is commonly

associated with deformation, these patches, lenses and veins routinely coalesce to form
Journal Pre-proof

interconnected networks in response to the imposed stress field. Rocks that show such

features are called migmatites or are described as migmatitic. Commonly, migmatites

comprise two distinct components: the neosome—that portion affected by partial melting—

and the paleosome—that portion unaffected by partial melting (Sawyer, 2008). Compared to

the protolith or nearby paleosome, the neosome is generally coarser grained (Fig. 2a, b) and,

if not subsequently deformed in a solid state, lacks structural elements, particularly any pre-

anatectic foliation and/or lineation (Fig. 2b, e, f). These characteristics are especially helpful

in identifying partial melting in leucocratic protoliths, such as granites (Johnson et al., 2013;

f
oo
Sawyer, 2010).

Partial melting changes a rock into a two-phase material generally composed of a


pr
lower-viscosity, lower-density liquid (melt) and a higher-viscosity, higher-density solid
e-
residue (aka residuum), such that they may separate if there is sufficient driving force (Fig.
Pr

2a). Thus, the residuum is the mostly solid part of the neosome from which melt drained. It is

commonly dominated by ferromagnesian minerals, such that it is darker coloured and


al

sometimes called melanosome; for example, this term could be applied to the areas of
n

residue in Fig. 2a. The complementary lighter-coloured part of the neosome, which represents
ur

material crystallized mostly from segregated melt, is called the leucosome. It is commonly
Jo

dominated by quartz and feldspar, and it may or may not correspond to a liquid composition

according to whether fractionated melt was lost during partial crystallisation. In the case of

incongruent melting (discussed in detail below), melt and a peritectic mineral or minerals

are the liquid and solid products of the melting reaction, respectively; they are commonly co-

located, for example the garnet in leucosome in Fig. 2b.

Migmatites are heterogeneous at various scales (Fig. 2) with a morphology controlled

by the melt fraction formerly present in the rock and the amount of strain that accompanied

partial melting (Sawyer, 2008). A first-order classification of migmatites, based on the


Journal Pre-proof

inferred former melt fraction in the rock (Sawyer, 2008), separates metatexite (lower former

melt fraction, pre-anatectic structures generally preserved) and diatexite (higher former melt

fraction, pre-anatectic structures disrupted) (Fig. 2c). Nonetheless, both metatexite and

diatexite may preserve coherent syn-anatectic and/or syn-crystallisation structural elements,

such as foliation and lineation (Solar and Brown, 2001a). The second-order classification of

migmatites relates to the amount of syn-anatectic strain, where various terms are used to

describe the geometry of the leucosome network in metatexites, such as patch (Fig. 2d),

stromatic (Fig. 2e), and net-structured (Fig. 2f), and the leucosome–residuum–paleosome

f
oo
relations in diatexites, such as raft, schollen (bottom-right of Fig. 2c) and schlieren (lower

half of Fig. 2c) (Johnson et al., 2001b; Sawyer, 2008). In source rocks with S > L fabrics,
pr
stromatic metatexite migmatites form in which the leucosome is generally planar and parallel
e-
to foliation, whereas in source rocks with L > S fabrics, the leucosome is generally spindle-
Pr

shaped and parallel to lineation (Brown et al., 1999).

During partial melting, melt is more mobile than the residue, and the latter may be
al

thought of as the part that largely remains in situ. Therefore, the location of the leucosome
n

records whether the melt has remained in place or tracks the pathway(s) by which it drained
ur

from the source during the anatectic event.


Jo

Thus, we may distinguish in situ leucosome, representing material remaining at its

source, in source leucosome, representing material that migrated away, but remained

confined within its source layer, and leucosome veins or dykes, representing material that

migrated out of its source layer, but which remained in the region affected by the anatectic

event (Sawyer, 2008). The hierarchy of leucosome types describes segments of the channel

network through which melt/magma migrates out of a migmatite (Brown, 2010a). In this

regard, an important concept to note is the melt connectivity transition (MCT), the melt

fraction (around 7 vol.% melt) above which the melt forms an interconnected network and
Journal Pre-proof

the strength of a partially-molten rock drops by an order of magnitude or more (Rosenberg

and Handy, 2005). If the melt volume produced during reaction was below the MCT (i.e. <7

vol.%) then, in the absence of significant syn-anatectic deformation, melt may be retained.

For further information about migmatites, particularly field descriptions and the

associated (evolving) nomenclature, the reader is referred to the books by Mehnert (1968),

Ashworth (1985), Ashworth and Brown (1990), Brown and Rushmer (2006a) and Sawyer

(2008).

f
oo
2.3. Thin section-scale features

pr
2.3.1. Microstructural features of partially-melted rocks
e-
Pr

Recognition of partial melting at the grain scale may not always be clear cut because

the microstructures associated with prograde melting and retrograde melt crystallization are
al

commonly modified during exhumation and/or cooling. For example, if melt is retained at
n

the grain scale, retrograde reaction between melt and the peritectic products may
ur

progressively destroy any prograde reaction microstructures (Brown, 2002). Fortunately, melt
Jo

drainage limits such retrograde effects (Brown, 2002; White and Powell, 2002b) and allows

inferences about melting to be made. After the metamorphic peak, microstructures in an

anatectic rock evolve during cooling to reduce the internal energy, as discussed in detail by

Holness and Sawyer (2008), Holness et al. (2011) and Vernon (2018).

Notwithstanding the potential limitations imposed by modifications during cooling, at

the grain scale, the former presence of melt may be inferred based on various

microstructures, in particular pseudomorphs of melt pores and grain boundary melt films.

Furthermore, in rocks that record P–T conditions consistent with partial melting, the presence
Journal Pre-proof

of mineral assemblages that correspond to a melt-producing (or, during cooling, a melt-

consuming) reaction is sufficient to demonstrate anatexis.

To illustrate these principles in a simple way, we refer to the experimental melting of

a natural rock—a quartz–muscovite schist—under H2O-saturated conditions at 0.1 GPa

(Rubie and Brearley, 1990). In this experiment (Fig. 3a), quartz and muscovite react to

produce a layer of melt along the quartz–muscovite grain boundary, together with the

localised nucleation of the solid reaction product mullite (due to excess Al2O3). This is an

example of a fluid-present incongruent melting reaction. For comparison, a natural example

f
oo
of muscovite and quartz breakdown to form a film of melt, now pseudomorphed by K-

feldspar, along the common grain boundary is shown in Fig. 3b.


pr
More mineralogically complex, but otherwise similar microstructures are produced by
e-
fluid-absent incongruent melting reactions involving, for example, the breakdown of biotite
Pr

and sillimanite/kyanite in metapelites to produce peritectic garnet and/or cordierite with melt

(Fig. 3c). In these cases, it is common to find inclusions of relict reactant minerals, such as
al

biotite and sillimanite, in the peritectic product that itself is surrounded by leucosome, as
n

shown in Figs 2b, d and 3c, and discussed by Jones and Brown (1990) and White et al.
ur

(2005). Drainage of melt may leave behind a compacted residue in which trapped melt
Jo

pockets (in 3D) may now be pseudomorphed by quartz or feldspar patches (in 2D) that have

cuspate edges with low dihedral angles against host minerals (Fig. 3c; Johnson et al.,

2003b). In some cases, prograde melting reactions may be inferred from the modal

proportions and/or microstructural relations between reactant and product phases (Guernina

and Sawyer, 2003; discussed further below). By contrast, fluid-present congruent melting

reactions, for example between quartz and feldspars, that produce only melt are better

identified using diagnostic microstructures related to either partial melting, such as corroded

relict grains of reactant minerals, or resulting from crystallisation of retained melt pockets,
Journal Pre-proof

such as pseudomorphs of quartz or feldspar after melt in pores and along grain boundaries.

Overall, a wide variety of microstructures related to partial melting and melt crystallisation is

commonly preserved in residual (migmatitic) source rocks, many of which are illustrated in

section F of the Atlas of Migmatites (Sawyer, 2008).

Microstructures related to the beginning of melt segregation in migmatites in orogenic

belts are rarely preserved due to the effects of syn-anatectic deformation and melt flow.

However, the grain-scale process of melt segregation into micro-leucosomes has been studied

in metatexites from the contact aureole of the Duluth complex in Minnesota (Sawyer, 2014b).

f
oo
The leucosomes are small and display a systematic progression in morphology from equant

micropores to elliptical micro-leucosomes (~1 mm in length), to en echelon arrays of


pr
elliptical micro-leucosomes, to zigzag-shaped micro-leucosomes (~4 mm in length), and, as
e-
the zigzag arrays connect, to high aspect ratio micro-leucosomes up to 19 mm in length. This
Pr

process is reminiscent of the way ductile fractures grow, and consistent with earlier

observations and interpretations at outcrop scale (Brown, 2004).


al

One microstructural feature largely ignored until relatively recently is the presence in
n

migmatites and residual granulites of tiny glass or nanogranite inclusions in peritectic


ur

minerals, most commonly garnet. These features have been interpreted as aliquots of
Jo

anatectic melt that were captured during growth of the peritectic mineral (Cesare et al., 2015).

Melt inclusions are particularly important because, although difficult to measure accurately,

they provide some of the best natural examples of largely unmodified anatectic melt

compositions from a wide range of P–T conditions (Bartoli et al., 2016; Cesare et al., 2011;

Ferrero et al., 2019).

2.3.2. Inference of melting reactions


Journal Pre-proof

To illustrate the methodology by which we may infer melting reactions, we briefly

review two examples from the literature. In the first example, a residual granulite terrane,

although most of the melt had drained from the rocks, the melting reaction and volume of

melt produced can be estimated from the mode of peritectic orthopyroxene in the granulites.

The second study concerns an eclogite sample in which the melt produced during exhumation

mostly crystallised in situ within the reaction microstructure itself.

In the case of the residual granulites, the prograde melting reaction was deduced from

the decrease in modal biotite and quartz and the increase in orthopyroxene and plagioclase

f
oo
contents, together with microstructures that indicated melting through breakdown of biotite +

quartz + plagioclase to melt + orthopyroxene + Fe–Ti oxides (Guernina and Sawyer, 2003).
pr
Based on the proportion of melt to peritectic orthopyroxene in experimental studies of this
e-
reaction, Guernina and Sawyer (2003) estimated the volume of melt produced on average to
Pr

have been ~30 vol.%. However, only 1–3 vol.% melt was retained in the melt-depleted

granulites, largely as quartz, plagioclase and K-feldspar pseudomorphs along what were once
al

melt-filled grain boundaries, leading to the conclusion that melt drainage in this case was
n

highly efficient. The counterintuitive relative increase in plagioclase in the mode of the
ur

residual granulites, despite it being a reactant phase, was due to breakdown of dominantly
Jo

biotite and quartz during melting, and efficient loss of melt, which was the dominant product

of the reaction relative to orthopyroxene.

In circumstances where the melt volume produced is low, for example below the

MCT, it may crystallise in situ preserving the reaction microstructure. Such an example has

been described by Feng et al. (2021) from eclogite; representative microstructures from their

study are shown in Fig. 4. Here the evidence for incongruent partial melting is clear. First,

there are optically continuous networks of plagioclase that span the width of the field of view

in Fig. 4a, b and are interpreted as pseudomorphs of former melt films along grain
Journal Pre-proof

boundaries. The plagioclase has low relief and is colourless in Fig. 4a but is pale orange (due

to the use of the 1 plate to emphasise continuity) in Fig. 4b. Second, this network of

plagioclase is associated with patches that show reaction microstructures, as seen on the left-

hand side and at the bottom right in Fig. 4a and b. Similarly, in the centre of Fig. 4c and d

there is a large (1000–1500 m) grain of Na–Al-rich pyroxene (omphacite) that is partially

replaced by optically-continuous plagioclase, which has pseudomorphed former melt. This is

a two-dimensional cut through a small pocket of leucosome that is a typical feature formed

f
during the early stage of crustal anatexis (Sawyer, 2008). Small grains of peritectic omphacite

oo
within the leucosome pockets have lower concentrations of the Na–Al pyroxene (jadeite)

pr
component than the large rock-forming grains. Phengite shows some evidence for minor

corrosion, but zoisite appears to be largely stable together with garnet. Major and trace
e-
element compositions of minerals support the microstructural interpretation that the melting
Pr

reaction in this case involved mostly omphacite with only minor contributions from phengite

+ zoisite + quartz to produce melt and the small grains of less sodic omphacite. Subsequently,
al

the melt has crystallised to plagioclase, which now pseudomorphs the former melt, both in
n

pockets and along grain boundaries, together with subhedral-to-euhedral amphibole grains in
ur

the leucosome pockets (Feng et al., 2021).


Jo

2.4. Melt extraction from a melting crustal source

Crustal melting and segregation, ascent and emplacement of melt are complex

processes. Melt is generated as micrometre-scale droplets at multiple grain boundaries in a

deep crustal source, but ultimately accumulates to form plutons in the upper crust that

comprise on the order of 103 to 104 km3 of magma aggregated from countless smaller batches

of melt (each perhaps 10–1 to 102 km3) (Brown, 2010a; Brown, 2010c). Examples of this style
Journal Pre-proof

of magma emplacement from different tectonic settings include the Manaslu granite in the

Himalayas (Deniel et al., 1987), the Phillips pluton in Maine (Pressley and Brown, 1999), and

the Tuolumne intrusive suite in California (Coleman et al., 2004). Thus, in the extreme case

of a lower crustal source and an upper crustal sink, melt extraction is a process with a length

scale that spans more than seven orders of magnitude or a volume concentration factor that

exceeds 1021 (Brown, 2010a; Brown, 2010c). The mechanisms by which this might have

been achieved, from the inception and growth of leucosomes to the formation of plutons,

have been addressed in many articles (e.g. Brown, 2013; Clemens and Stevens, 2016;

f
oo
Sawyer, 2014a; Soesoo and Bons, 2015; Weinberg et al., 2013), and we provide only a

succinct summary here.


pr
In residual migmatites and granulites, evidence of melt-extraction pathways at
e-
outcrop scale is recorded by crystallized products of melt (leucosome) and residual material
Pr

from which melt has drained (melanosome). These features form networks or arrays that

potentially demonstrate the temporal and spatial relations between deformation and melting.
al

Multiple studies have shown that syn-anatectic shear-enhanced compaction drove melt from
n

dynamically created porosity, recorded by pseudomorphs after melt along grain boundaries,
ur

into melt-induced deformation-band networks, represented by leucosome-filled veins, and


Jo

from these to granite-filled ascent conduits, evidenced by anastomosing ductile or brittle

dykes (Brown, 2004, 2005; Etheridge et al., 2020; Weinberg and Regenauer-Lieb, 2010;

Weinberg et al., 2015). In these cases, melt may have accumulated locally, forming a

connected microfracture network (Etheridge et al., 2020), until the volume reached then

exceeded the MCT throughout a large volume of the source (Rosenberg and Handy, 2005).

This multi-scale connectivity allowed the melt to drain down hydraulic potential gradients to

crack-like, ductile opening-mode fractures propagating from dilation or shear bands during

one or several melt extraction events (Brown, 2013). Thus, any outcrop surface though a
Journal Pre-proof

migmatite terrain likely represents a level of melt generation, but also records evidence of

influx of melt, mostly from deeper in the crust, and drainage of melt to shallower crustal

levels (Fig. 5). In these circumstances, the volume of leucosome frozen in the anatectic crust

may exceed that produced at that level, even though some of the melt produced locally has

migrated (Marchildon and Brown, 2003; Yakymchuk et al., 2013).

In granulite facies migmatite terrains where syn-anatectic deformation drove efficient

melt drainage, for example along the foliation, even though the melt volume remained below

the MCT (Guernina and Sawyer, 2003), melt flow may have been continuous for as long as

f
oo
melt was generated or fluxed through the anatectic crust. If a sufficient volume of melt

accumulates in the anatectic crust, diatexite magma may form, which may lead to multiple
pr
phases of partial crystallization and melt extraction under amphibolite and granulite facies
e-
conditions (Sawyer, 2020; Solar and Brown, 2001b). Deep crustal migmatite complexes may
Pr

link to shallower granite plutons (Brown et al., 2016; Brown and Solar, 1999; Johnson et al.,

2001b) or they may give way upwards to injection complexes that include sheeted plutons
al

(Morfin et al., 2013, 2014).


n

Leucosomes in metatexite migmatites have thickness and spacing distributions


ur

generally consistent with being sampled from power-law (scale-invariant) distributions


Jo

(Bonamici and Duebendorfer, 2010; Brown, 2003; Marchildon and Brown, 2003; Saukko et

al., 2020; Tanner, 1999; Yakymchuk et al., 2013). Furthermore, the spatial distribution of

ascent conduits in residual crustal sources and intrusive complexes at higher structural levels

indicates that these features are also scale-invariant (Brown, 2005; Cruden, 2006; Cruden and

McCaffrey, 2001; Hall and Kisters, 2012; Koukouvelas et al., 2006; McCaffrey and Petford,

1997). These conclusions from outcrop and map measurements are supported by numerical

modelling of stepwise accumulation and transport of batches of melt by propagation of

hydrofractures (Bons et al., 2004). In this study, Bons et al. (2004) showed that the model
Journal Pre-proof

system developed spontaneously into a self-organised critical state, in which the melt

distribution was organised in a self-similar way that allowed melt drainage above a system-

determined volume threshold. In nature, in the absence of significant deformation, this

threshold could be the MCT, but during regional-scale crustal anatexis, deformation will

lower the threshold below the MCT allowing smaller melt batches to migrate. Thus, it is

likely that anatectic systems in the crust are strongly self-organized at a range of scales from

the bottom up, becoming more ordered by decreasing the number and increasing the width of

melt-bearing conduits from the anatectic zone through the overlying subsolidus crust to the

f
oo
ductile-to-brittle transition zone, where melt accumulates in plutons (Brown, 2010c).

pr
e-
3. The thermodynamic basis of calculated phase equilibria
Pr

Any quantitative treatment of crustal melting is founded on equilibrium


al

thermodynamics. As the name implies, thermodynamics is the branch of physics concerned


n

with the relationship between heat and work, and how heat relates to other forms of energy
ur

and the properties of matter. As a serious scientific discipline, thermodynamics emerged from
Jo

the industrial revolution, principally through a desire to understand, and thereby increase the

efficiency of, steam engines (Carnot, 1824). Formalisation of the principles and equations

governing thermodynamics involved some of the best-known names in science, including

Ludwig Boltzmann, Robert Boyle, Rudolf Clausius, Robert Hooke, James Clerk Maxwell,

Max Planck, William Rankine and William Thomson (Lord Kelvin). To this (non-exhaustive)

list can be added the name of J. Willard Gibbs, who elegantly showed how thermodynamic

processes could be visualised graphically in terms of changing pressure, temperature, volume


Journal Pre-proof

and other thermodynamic variables (Gibbs, 1879), an approach that forms the foundation of

much of what follows.

Before considering melt-bearing (suprasolidus) phase equilibria, we will discuss those

aspects of equilibrium thermodynamics that are a pre-requisite for understanding what is to

follow, with reference to some straightforward examples involving simple minerals (solids).

For those seeking a deeper understanding, the HPx-eos and THERMOCALC

(https://hpxeosandthermocalc.org/) and Science Education Resource Center at Carleton

College (https://serc.carleton.edu/research_education/equilibria/index.html) websites are

f
oo
good places to start. For a comprehensive treatment of the application of equilibrium

thermodynamics to petrological problems, a number of excellent books are available,


pr
including those by Wood and Fraser (1976), Powell (1978), Spear (1995) and Ganguly
e-
(2008).
Pr

3.1. Some simple definitions


n al

A system is that part of the universe that we wish to consider. In many cases that
ur

system might be an individual rock sample or a thin section. However, it might be an entire
Jo

planet or a single atomic nucleus. A system may be completely isolated, meaning it cannot

exchange energy or matter with its surroundings, closed, such that it can exchange energy but

not matter with its surroundings, or open, meaning it can exchange both energy and matter. In

nature, crustal rocks undergoing metamorphism are open systems, in which the heat sources

may be internal (e.g. radioactive decay or viscous dissipation) or external (e.g. mantle heat,

heat from nearby plutons), and fluids (principally H2O and/or CO2) and/or melt may be

produced and lost. However, when modelling phase equilibria, we commonly make the

(simplistic but necessary) assumption that the system is closed, as discussed later in this
Journal Pre-proof

section. A more comprehensive treatment of closed versus open behaviour in partially melted

systems is given in section 5.

Of fundamental importance to calculated phase equilibria is the bulk chemical

composition (hereafter bulk composition) of a system. The bulk composition consists of

variable quantities of the different independent chemical components that comprise the

system. Crustal rocks mostly consist of the elements oxygen, silicon, aluminium, iron,

calcium, sodium, magnesium, potassium and titanium, which together account for ~99% of

the crust by mass (Goldschmidt, 1954; Hofmann, 1988; Rudnick and Gao, 2003). By

f
oo
convention, concentrations of these major elements, along with some minor elements such as

hydrogen, phosphorus, manganese and chromium, are generally expressed as the weight
pr
percent of their oxides (SiO2, Al2O3, FeO/Fe2O3, CaO, Na2O, MgO, K2O, TiO2, H2O, P2O5,
e-
MnO, Cr2O3), whereas concentrations of the remaining trace elements are reported as weight
Pr

fractions in parts per million (ppm; equivalent to grams per tonne and more formally as

mg.kg–1). As we shall see, petrological experiments and phase equilibria are mostly founded
al

on systems considering the major element oxides as the components. However, how the trace
n

elements behave is key to deciphering the petrogenesis of many rocks (see section 9), in
ur

constraining the behaviour of accessory minerals (section 10) and in understanding the
Jo

formation of many key mineral deposits.

A phase is a part of the system within which all physical and chemical properties are

essentially uniform, and which is (commonly) mechanically separable. For our purposes,

phases are the different minerals (solid phases) that coexist with a melt (liquid phase) and/or

fluid (liquid or gas phase). Importantly, the number of components in a system is the

minimum required to define the compositions of all the phases within that system. For

example, when considering reactions between the minerals quartz (SiO2), corundum (Al2O3)

and sillimanite (Al2SiO5), two components (SiO2 and Al2O3) are required to define the
Journal Pre-proof

system. However, when considering reactions between the aluminosilicate polymorphs

sillimanite, andalusite and kyanite, only one component (Al2SiO5) is required.

The changeable physicochemical properties of a system can be subdivided into

intensive and extensive variables. An intensive variable is one that is independent of the size

of that system, or the amount of matter within it. By definition, at equilibrium (from the Latin

aequi = equal and libra = balance), the state at which the energy of the system is minimised,

the values of the intensive variables are identical everywhere within the system. For our

purposes, the main intensive variables are temperature (T; in ºC or K) and pressure (P; in

f
oo
kbar or GPa), but also include chemical potential (). In response to changes in these

pr
variables, the system will attempt to equilibrate through conduction/convection (for T),

deformation (for P) and diffusion (for ). By contrast, extensive variables do depend on the
e-
size of the system and the quantity of matter within it. These include entropy (S), volume (V),
Pr

and chemical composition (X), which can be considered as the conjugate pairs to T, P and ,

respectively, as discussed in detail by Powell et al. (2005). We will return to many of these
al

terms later.
n
ur

3.2. Gibbs’ phase rule


Jo

Stemming from his seminal paper (Gibbs, 1879), Gibbs derived the phase rule that

bears his name and which, in a system at equilibrium, relates the number of components (C)

and phases (P) to the degrees of freedom (F; also known as variance), the number of

intensive variables that can be changed independently. The Gibbs phase rule is F = C – P + 2,

and is explained with reference to a simple example, the aluminosilicate phase diagram (Fig.

6).
Journal Pre-proof

As discussed previously, the stability of the aluminosilicate polymorphs can be

described in the one-component (C = 1) chemical system Al2SiO5. Let us consider a P–T

condition, for example a point at 0.8 GPa and 550 ºC, where a rock (the system) with a

composition of pure Al2SiO5 contains grains of only one phase, kyanite. With P = 1, the

Gibbs phase rule tells us that F = 2 (1 – 1 + 2), meaning that, within limits, we are free to

change pressure (to a’) or temperature (to a’’) or both (to a’’’) without ‘upsetting’ the system

(changing the stable assemblage). With two degrees of freedom, the primary phase field of

kyanite (or sillimanite or andalusite) is divariant. Now consider a P–T condition at which the

f
oo
rock contains two phases, for example point b, where both kyanite and sillimanite are stable

(i.e. P = 2) at ~0.65 GPa, 645 ºC. Here, F = 1 (2 – 1 + 2), meaning that if we change one of
pr
the intensive variables, the other must change in a dependent way. For example, we can
e-
change the pressure (e.g. to Pb’) but, if we are to maintain the same assemblage (kyanite +
Pr

sillimanite), the temperature must change to Tb’. Similarly, we may change the temperature

(to Tb’’), but then pressure must also change (to Pb’’). The lines along which F = 1 are
al

univariant equilibria. Finally, point c is the unique P–T condition (~0.42 GPa, 540 ºC) at
n

which all three polymorphs may stably coexist, the so-called aluminosilicate triple point.
ur

Here, P = 3, so F = 0 (3 – 1 + 2)––we can change neither P nor T without changing the


Jo

assemblage. The triple point is an example of an invariant equilibrium.

3.3. Reactions, heat and temperature

Heat is the thermal energy supplied to (or extracted from) a system. The addition or

subtraction of heat may lead to a reaction, the process by which a phase or combination of

phases transforms into a different phase or combination of phases. A change in the supply of

heat can lead to a change in temperature, which is a measure of the average vibrational
Journal Pre-proof

energy of particles within the system, but in some cases does not. Let us return to the triple

point (Fig. 6, point c), where the rock contains coexisting kyanite, sillimanite and andalusite.

If heat is supplied to the system at constant pressure, the tendency is for the system to move

into the primary phase field of sillimanite at higher temperature. However, it cannot do so

until all the kyanite and andalusite have been consumed. Initially, the additional heat (energy)

does the work of transforming the kyanite and andalusite to sillimanite, during which the

temperature remains at a fixed value (that of the triple point). Only when all kyanite and

andalusite have been consumed, and the rock consists only of sillimanite, does heating lead to

f
oo
a rise in temperature. The transformation described here is a reaction of the form:

pr
kyanite + andalusite → sillimanite. (1)
e-
Pr

On the left-hand-side of this reaction (kyanite + andalusite) are the reactants (the phase or

phases that are consumed in the reaction), whereas on the right-hand-side (sillimanite) are the
al

products (the phase or phases produced by the reaction).


n
ur

3.4. Gibbs’ free energy


Jo

The stability fields shown in Fig. 6 are a manifestation of a key thermodynamic

property that also derives its name from J. Willard Gibbs, the Gibbs free energy (G) of the

three aluminosilicate polymorphs as a function of pressure and temperature. By definition, in

a system at equilibrium, G is minimised. At points a, a’, a’’ and a’’’ in Fig. 6, the Gibbs free

energy of kyanite (GKy) is less than that of sillimanite or andalusite, or of any combination of

the three minerals. At points b, b’ and b’’’, the Gibbs free energies of kyanite and sillimanite
Journal Pre-proof

are equal, and both are lower than that of andalusite (GKy = GSill < GAnd). At the triple point,

the Gibbs free energy of all phases is the same (GKy = GSill = GAnd).

The Gibbs free energy of a substance, which changes with pressure and temperature,

can be broken down into its component parts that include some of the variables we

introduced earlier:

G = E + PV – TS, (2)

f
oo
where E is the internal energy of the phase, V is the volume and S is the entropy. The internal

energy can be thought of as the total energy a substance possesses, whereas entropy is a
pr
measure of the randomness or disorder of a substance, which increases with increasing
e-
temperature. By definition, at absolute zero (0 K or –273.15 °C), entropy is also zero,
Pr

meaning the atoms within a substance are perfectly stationary such that it is perfectly ordered.

As heat is supplied, so the vibrational energy (temperature) and entropy both increase.
al

An undoubtedly simplistic but potentially useful way to think about how the different
n

terms in equation 2 relate to each other is as follows: the Gibbs free energy, G, can be
ur

thought of as the amount of energy a substance possesses that is available (free) for it to react
Jo

with other substances. This free energy is not simply the total energy (E) the substance

possesses, but must account for the energy that the substance requires to maintain its

coherence. This comprises an entropy contribution that increases with temperature and which

must be subtracted from the energy available for reaction (G)––temperature makes the

substance vibrate more vigorously and become less ordered (S), such that it needs to use

more energy to maintain its coherence––and a volume term (V) that is pressure dependent,

but which is added to the energy available for reaction––a higher pressure helps the substance

maintain its coherence. The first two terms on the right-hand-side of equation 2 together
Journal Pre-proof

define another thermodynamic variable, enthalpy (H = E + PV), the amount of energy

consumed or released when a substance expands or contracts. This leads to an alternative

definition of the Gibbs free energy of a substance:

G = H – TS. (3)

When thinking about reactions between substances, equations 2 and 3 need to be

recast in terms of the difference in the energies of the reactants and products, such that:

f
oo
G = E + PV – TS

and
pr (4)
e-
G = H – TS, (5)
Pr

in which the delta () notation refers to the total energy of the products minus the total
al

energy of the reactants. If G > 0, the products have a higher Gibbs free energy than, and are
n

unstable with respect to, the reactants. Conversely, if G < 0 the products are more stable
ur

than the reactants. If G = 0, the products and reactants are in stable coexistence and the
Jo

system is at equilibrium.

In detail, the entropy change (S) of a system includes for each phase of interest a

heat capacity (Cp) term, and the volume change (V) includes terms accounting for thermal

expansivity () and compressibility (), all of which have a dependence on temperature

and/or pressure. Those wishing to delve deeper might begin with the web pages and books

cited in the introductory paragraph of this section.

3.5. Reactions and the equilibrium relationship


Journal Pre-proof

Thus far we have considered only the thermodynamic properties of pure phases with

a fixed chemical composition, such as kyanite, sillimanite and andalusite. However, most

phases in rocks are solid solutions, whose composition is a mixture of two or more end

members. For example, olivine is a solid solution mineral that consists essentially of a

mixture of the end members forsterite (Mg2SiO4) and fayalite (Fe2SiO4), in which

composition changes by the substitution of Mg for Fe2+. The energetics of mixing in these

impure phases must be accounted for, which brings us to the fundamental equation

f
oo
underlying phase diagram calculations.

For a system at equilibrium:


pr
e-
G = –RT ln K, (6)
Pr

where R is the gas constant (= 8.314 J.K–1.mol–1). We have already discussed the left-hand-
al

side of this equation (G, but added a degree sign to remind us that it refers to the pure end
n

members), which can be broken down into a combination of the end-member thermodynamic
ur

variables given in equations 4 and 5. However, the right-hand-side of equation 6 includes a


Jo

new term, the equilibrium constant (K), which permits consideration of the energies

associated with impure (solid solution) phases.

Consider a univariant (F = 1) reaction in the three-component (C = 3) CAS (CaO–

Al2O3–SiO2) system involving the mineral end members anorthite (CaAl2Si2O8), grossular

garnet (Ca3Al2Si3O12), kyanite (Al2SiO5) and quartz (SiO2) (P = 4). In this and any chemical

reaction, mass is conserved, requiring that the reaction be balanced. This mass balance is

achieved by adding stoichiometric coefficients, the number of atoms or molecules of each

reactant and product phase involved in a reaction. In the example given, three moles of
Journal Pre-proof

reactant anorthite are consumed to produce one mole each of grossular and quartz and two of

kyanite (if a single mole is involved, the ‘1’ is by convention omitted):

3 anorthite → grossular + 2 kyanite + quartz (7)

3 CaAl2Si2O8 → Ca3Al2Si3O12 + 2 Al2SiO5 + SiO2

In natural rocks in which these minerals occur, such as metapelites, both kyanite and quartz

may be considered as effectively pure phases with a fixed composition, but plagioclase and

f
oo
garnet are seldom (or ever) pure anorthite and grossular. In particular, plagioclase

incorporates variable amounts of Na (as the albite component) and garnet commonly contains
pr
significant quantities of Fe2+, Mg and Mn (as the almandine, pyrope and spessartine
e-
components, respectively). These ‘diluting’ species reduce the energetic contribution of the
Pr

Ca end members. This contribution is K, which consists of the product of the activity (a) of

each new phase (product) raised to the power of its respective stoichiometric coefficient,
al

divided by the product of the activity of each reactant phase raised to the power of its
n

respective stoichiometric coefficient. In the example given:


ur
Jo

K = aGr. aKy2. aQtz / aAn3. (8)

The activity can be thought of as the ‘effective concentration’ of an end member within the

solid solution mineral in question. In essence, the energy associated with the end member is

not simply a linear function of the concentration (or mole fraction, X) of that end member,

which we can easily measure using an electron probe microanalyser, but involves additional

energy related to how different atoms (or ions) within the substance interact, which we can

estimate based on theory and/or experiments. In pure phases, a = 1, but in solid solution
Journal Pre-proof

phases (including melts and supercritical fluids), a < 1. Thus, in equation 7, if we consider

quartz and kyanite to be pure, the associated terms can be disregarded to leave aAn3, the

activity of anorthite in plagioclase raised to the power of three (there are three moles of

plagioclase in the balanced reaction 8) and aGr, the activity of the grossular component in

garnet, of which only one mole participates in the reaction. Thus, equation 8 reduces to K =

aGr / aAn3, which forms the basis for the so-called GASP (garnet–aluminosilicate–quartz

(silica)–plagioclase) barometer (Ghent, 1976; Holdaway, 2001).

Therefore, to perform quantitative calculations to constrain the (P, T) conditions of

f
oo
equilibration of a system comprising a mixture of pure and impure phases (i.e. a rock), we

pr
must solve both sides of equation 6. Solving the left-hand-side (G) requires a knowledge of

the thermodynamic properties of the pure end members, whereas solving the right-hand-side
e-
–RT ln K) requires a knowledge of the energies associated with mixing (the interaction
Pr

energy, W) between the various end members of solid solution phases, collectively termed

activity–composition (a–X) relationships (also known as solution models) for the phases of
al

interest. Providing, expanding and refining these data represents a very significant proportion
n

of the progress that has been made in quantitative petrology over the past few decades. We
ur

will address the various sources of thermodynamic data and existing a–X models required to
Jo

solve equation 6, and the different types of software that are most commonly used, after

describing the types of suprasolidus phase diagrams a user might encounter or want to

calculate.

3.6. Slopes of reactions


Journal Pre-proof

The Clapeyron equation (also called the Clausius–Clapeyron equation or relationship)

relates the slope of a reaction line in P–T space (P/T) to the entropy and volume terms,

such that:

P/T = S/V. (9)

As we will see, this relationship is useful in qualitatively orienting reaction lines in P–T space

f
and in establishing which phases occur on which side of the reaction, in particular when

oo
constructing petrogenetic grids (see section 4.2). For example, reaction 7 involves only solid

pr
phases (minerals), for which the differences in the degree of disorder (S) between the

products and reactants is small, and the energetics of the reaction relate mainly to changes in
e-
volume (V). Consequently, the slope of the reaction 7 is shallow (small P/T) in P–T
Pr

space, such that it provides a useful means of constraining pressure in rocks of an appropriate

composition (i.e. it is a good (geo)barometer).


al

Of particular importance here is that, during heating, most crustal rocks first undergo
n

dehydration reactions then, as the solidus is crossed, melting reactions. As the products
ur

include H2O and melt, respectively, both types of reactions involve a positive volume change
Jo

and a very significant increase in the disorder of the system associated with solid to fluid

and/or liquid transitions. Consequently, dehydration and melting reactions generally have

steep positive slopes in P–T space, with H2O or melt on the high T (high S) side of the

reaction, as intuition tells us.

3.7. Equilibrium volume


Journal Pre-proof

To end, an important but difficult variable to consider is the equilibrium (or

equilibration) volume, that is, the volume of a system that is considered to have attained

(chemical/diffusive) equilibrium (Powell et al., 2005, 2019). In high-grade crustal rocks, such

as migmatites and granulites, equilibrium is generally assumed to have been attained on a

hand-specimen or larger scale, although in many cases this may not be justified. For example,

where mineral grains preserve compositional zonation, such as Ca in garnet, the core of those

grains cannot be in equilibrium with garnet rims or the matrix of the rock. In addition, in

crustal rocks that reached the highest metamorphic temperatures (900–1000 °C), reaction

f
oo
microstructures such as symplectites and coronae are common, indicating gradients in

chemical potential and that equilibrium was attained at a grain scale or smaller (Mitchell et
pr
al., 2020; Powell et al., 2019; White and Powell, 2011; White et al., 2008). The development
e-
of strongly-domainal equilibrium volumes generally reflects the absence of melt and/or fluid,
Pr

inhibiting diffusion, which in the case of high-T migmatites and granulites requires effective

melt loss (Johnson et al., 2010; Mitchell et al., 2020).


al

In nature, constraining a single equilibration volume is impossible for several reasons,


n

including that different elements diffuse at different rates that are dependent on changes in P
ur

and T, the minerals involved and whether any melt or fluid was present at equilibrium, and
Jo

many other factors (Powell et al., 2019). However, in practice, an appropriate equilibrium

volume can generally be estimated, at least for the major elements, that is constrained by the

slowest diffusing species, commonly thought to be Al (Carmichael, 1969).

4. Suprasolidus phase equilibria


Journal Pre-proof

As described briefly in the introduction, melting is the process by which a solid or

solids are transformed into a liquid, usually by the addition of heat. By assuming that liquid is

not allowed to migrate, this case can be considered as closed system behaviour and modelled

as such. However, melting of rocks may also occur due to a reduction in pressure

(decompression melting, which also can be modelled as a closed system) or the addition of a

fluid, generally H2O (H2O- or fluid-fluxed melting), which is an example of open system

behaviour. To begin this section, we summarise some key aspects of melting experiments

using various starting compositions. Then we describe the fundamentals of suprasolidus

f
oo
phase equilibria and associated terminology with reference to the simple case by which heat

is supplied to a closed and compositionally simple system. We examine open system


pr
behaviour, including fluid-fluxed melting, in section 5.
e-
Pr

4.1. Experiments
al

Experimental petrology fundamentally underpins the thermodynamic data on which


n

quantitative phase equilibrium calculations depend. For example, the end-member data used
ur

in the construction of many of the calculated phase diagrams that follow (Holland and Powell
Jo

(2011) come from >650 articles, most of which report experimental studies (although not

necessarily partial melting experiments). We cannot do justice to the numerous melting

experiments that have been undertaken, but we provide some background information for

context. Detailed summaries are provided elsewhere (Castro, 2013; Gao et al., 2016b;

Johannes and Holtz, 1996; Patiño Douce, 1999). In particular, table 1 in the paper by Gao et

al. (2016b) provides a synthesis of many of the experimental studies that have considered

partial melting of a broad range of crustal rocks or their synthetic analogues, including both

(meta)igneous and (meta)sedimentary compositions, over a wide range of P–T conditions.


Journal Pre-proof

Table 1. Bulk compositions (mol.% normalised to 100%) used in the construction of

pseudosections.

SiO Ti Al2 Fe Mg Ca Na2 K2 H2 Mn XM Xferr K/N


Figure Sample 2 O2 O3 OT O O O O O O O g ic a
64.3 3.1 0.4
9 Taylor exp. 3 – 9.35 5.51 4.54 – 3.22 8 9.87 – – 5 – 0.99
54.9 1.2 11.4 0.3 0.5 0.3 0.1
12 C-F2 3 7 8.91 9 6.22 8.56 2.95 6 4.73 – 8 8 0 0.12
50.0 1.0 12.1 11.6 0.2 0.4 0.6 0.1
14a N-MORB 9 1 8.69 7.78 5 7 2.50 2 5.44 – 7 4 2 0.09
49.1 0.5 17.1 0.1 0.4 0.6 0.1
14b Isua basalt 5 5 6.11 9.55 8 9.95 1.49 8 5.37 – 8 7 0 0.12
15a and X = 0 in 53.4 1.0 11.9 0.6 1.1 0.5 0.3
15b 87S35a 8 1 0 7.68 6.80 9.47 4.33 8 3.48 – 6 6 0 0.16
45.1 0.8 10.0 0.5 18.5 0.9 0.5 0.3
X = 1 on 15b 87S35a 2 6 3 6.48 5.74 7.99 3.66 7 8 – 8 6 0 0.16
60.5 0.8 12.8 2.7 0.6 0.4 0.1

f
20 Metapelite 5 5 0 7.52 5.18 1.49 1.88 6 6.24 0.16 0 5 6 1.47

oo
Metagreywa 75.3 0.5 1.2 0.3 0.5 0.1
22 cke 5 1 7.96 4.09 3.74 1.12 2.77 2 2.61 0.34 1 2 5 0.44

4.1.1. Felsic (meta)igneous rocks pr


The long history of experimental studies of granitic rocks is documented in detail in
e-
the book by Johannes and Holtz (1996), so we highlight only a few key aspects here. The first
Pr

(fluid-absent) experiments melting a granite, by Hall (1805), date back almost to the

beginnings of geology as a scientific discipline, and were undertaken with the direct purpose
al

of testing Hutton’s postulate of an origin of granite through crystallisation of magma (Hutton,


n

1788). However, it was Norman L Bowen who established experimental petrology as a


ur

serious subdiscipline within geology. With Oliver Tuttle, he published the first detailed
Jo

investigations on melting in the synthetic haplogranite system (Ab–Or–Qz–H2O) (Tuttle and

Bowen, 1958) that resolved the granite controversy (Read, 1957), in which it was established

that granites may form at T < 700 °C. In the same study, these authors also performed

experiments on two natural plagioclase-bearing granites from the USA, the two-mica (‘S-

type’) Westerly granite, subsequently used in several key experimental studies, and an

amphibole-bearing (‘I-type’) granite from Quincy (the I–S typology of granites is discussed

in section 6).
Journal Pre-proof

Noting the close correspondence between the onset of melting in the synthetic and

natural systems, Tuttle and Bowen (1958) constrained the location in P–T space of the H2O-

saturated (‘wet’) solidus (the locus of P–T conditions at which, with increasing T, an H2O-

saturated composition will begin to melt) in quartz- and feldspar-bearing rocks that is

fundamental to understanding suprasolidus phase equilibria in common crustal lithologies.

The H2O-saturated solidus has a shallow negative slope in P–T space (P/T) at pressures

below a few tenths of a GPa, but steepens at higher P. Subsequently, this study was extended

f
to pressures from 0.4 to 1.0 GPa (Luth et al., 1964) and to investigate the influence of CO2

oo
(Wyllie and Tuttle, 1959), which was found to be much less soluble than H2O in granitic

pr
melt. These last authors went on to show the effects of other volatile species, notably that the

addition of Li and F reduced the temperature of the wet solidus by 50–90 °C (Wyllie and
e-
Tuttle, 1964; Wyllie and Tuttle, 1961).
Pr

Experiments considering the effects of H2O, ranging from anhydrous to H2O-

saturated, were key to understanding the processes of H2O-undersaturated (fluid-absent)


al

partial melting and the fundamental controls of H2O-content and temperature in the genesis
n

of granite (Clemens and Wall, 1981; Maaløe and Wyllie, 1975; Robertson and Wyllie, 1971a,
ur

b; Whitney, 1988; Wyllie, 1971). Experiments extending to much higher pressures (up to ~3–
Jo

4 GPa) showed that the H2O-saturated solidus had a positive P/T above ~2 GPa and that

aqueous solutions at high-P are significantly enriched in solute (Boettcher and Wyllie, 1968;

Huang and Wyllie, 1975; Stern and Wyllie, 1981). Although most of the melting experiments

on felsic igneous rocks have concentrated on ‘S-type’ (muscovite- and/or biotite-bearing)

granites, there have been detailed studies on tonalite (Patiño Douce, 1997, 2005; Singh and

Johannes, 1996a, b; Watkins et al., 2007) and trondhjemite (van der Laan and Wyllie, 1992)

starting compositions. These rocks are compositionally and mineralogically similar to

greywackes, and behave in an analogous manner (see below and section 7).
Journal Pre-proof

The compositions of melt derived from felsic igneous protoliths are highly variable,

and a function of P, T, aH2O and source composition. As a generalisation, muscovite-bearing

(‘S-type’) granites produce high K/Na peraluminous melts ((ASI = Al2O3/(Al2O3 – 3.33P2O5

+ Na2O + K2O) > 1; (Frost et al., 2001; Shand, 1943), biotite-bearing (‘S-type') granites

produce peraluminous to weakly metaluminous melts (ASI < 1), and rocks with lower K/Na

ratios and hornblende-bearing (‘I-type’) granites produce mostly metaluminous sodic melts

(Johannes and Holtz, 1996; Whitney, 1988). These compositions and the peritectic minerals

that form with melt correspond to those produced by partial melting of common

f
oo
metasedimentary (in the case of ‘S-type’) and metabasic (in the case of ‘I-type’) protoliths, as

detailed in the next two subsections and in sections 6 and 7. However, granites generally are
pr
less fertile under higher T fluid-absent conditions due to a lower abundance of hydrous
e-
reactant minerals. Nevertheless, as we show later, a wide variety of melt compositions may
Pr

be generated from a single source composition depending mostly on the P, T and aH2O

conditions of melting (Conrad et al., 1988; Holtz and Johannes, 1991).


n al

4.1.2. Basic (meta)igneous rocks


ur

The first detailed experiments on the anatexis of a hydrated basaltic composition


Jo

(amphibolite) were undertaken by Yoder and Tilley (1962), who noted that the addition of

H2O to basalt led to a drastic drop in the temperature of the solidus such that the onset of

melting closely corresponded with that of granite. They also noted the ‘anorthositic’ (likely

tonalitic or trondhjemitic) composition of the melts so formed. Building on this and later

studies, in particular that of Burnham (1979), many of the key experiments on the partial

melting of amphibolite, including both fluid-absent and fluid-present conditions, were

undertaken in the period from 1989 to 1995 (Beard and Lofgren, 1989, 1991; Patiño Douce

and Beard, 1995; Rapp and Watson, 1995; Rapp et al., 1991; Rushmer, 1991; Sen and Dunn,
Journal Pre-proof

1994; Wolf and Wyllie, 1994, 1995; Wyllie and Wolf, 1993). The results are generally

consistent, allowing some general statements to be made regarding the appropriate variables

and phase equilibria.

Fluid-absent melting of amphibolites proceeds by incongruent reactions mainly

consuming hornblende, along with any biotite and quartz; although some plagioclase may be

consumed, in most fluid-absent experiments it is not a major reactant (Beard and Lofgren,

1989, 1991; Patiño Douce and Beard, 1995). The onset of fluid-absent partial melting is

dependent on the starting bulk composition and hornblende (i.e. structurally-bound H2O)

f
oo
content, and for a fully-saturated amphibolite at plausible crustal pressures (~0.5–1.5 GPa)

the (wet) solidus occurs at temperatures of 600–700 °C (see summary in (Wyllie and Wolf,

1993).
pr
e-
The melt produced in fluid-absent experiments ranges in composition from
Pr

granodiorite to tonalite, but most commonly is low-K metaluminous (ASI < 1) to weakly

peraluminous (ASI > 1) and trondhjemitic to tonalitic (i.e. sodic); melt volumes show a
al

broadly linear increase with temperature, and are around 40–50 vol.% at 1000 °C, with melt
n

compositions becoming increasingly sodic (higher Na2O/K2O) and magnesian, and depleted
ur

in Si and Al (Beard and Lofgren, 1989, 1991; Patiño Douce and Beard, 1995; Rapp and
Jo

Watson, 1995; Rapp et al., 1991; Rushmer, 1991; Sen and Dunn, 1994; Wolf and Wyllie,

1994, 1995; Wyllie and Wolf, 1993). The main solid products are clinopyroxene with or

without orthopyroxene at lower pressures and include garnet (with or without epidote) at

higher pressures, with plagioclase generally remaining stable to temperatures in excess of

1000 °C (López and Castro, 2001; Patiño Douce and Beard, 1995; Wyllie and Wolf, 1993);

see section 6). The lower limit of suprasolidus garnet stability, which has a shallow P/T, is

strongly dependent on bulk composition, in particular XMg, and garnet may be stable at

pressures as low as ~0.7 GPa (Johnson et al., 2017; Wyllie and Wolf, 1993).
Journal Pre-proof

Importantly, the experiments predict total consumption of hornblende at temperatures

of around 1000 °C at ~1.0 GPa, but at lower temperatures at both lower and higher pressures,

and hornblende is not stable with melt at pressures much in excess of 2 GPa (Beard and

Lofgren, 1991; Burnham, 1979; Rushmer, 1991; Wyllie and Wolf, 1993). This has important

consequences for the partial melting of eclogite. Although experiments at P > or >> 2.0 GPa

predict the formation of large volumes of melt (Rapp and Watson, 1995; Rapp et al., 1991),

the starting materials used in these experiments contain ~1 vol.% H2O. As almost all of this

H2O must exist as a free fluid in natural rocks, it is likely to have been lost during burial,

f
oo
questioning the importance of high-pressure melting of eclogite in nature (see also section 8).

Partial melting of amphibolite in experiments with excess H2O produces results that
pr
differ significantly from fluid-absent experiments (Beard and Lofgren, 1989, 1991). As the
e-
H2O required for melting is already present, fluid-present melting proceeds by reactions
Pr

consuming plagioclase and any quartz along with the H2O. Although peritectic phases are

predicted in some of the experiments, melt is the major product, and hornblende is predicted
al

to remain stable in the residuum until very high temperature. Importantly, experimental H2O-
n

present melting produces melts that are generally low-K but may be strongly peraluminous
ur

(Beard and Lofgren, 1989, 1991).


Jo

4.1.3. Metasedimentary rocks

The majority of partial melting experiments on metasedimentary lithologies have used

(meta)pelitic or (meta)greywacke starting compositions, and thereby considered fluid-absent

melting by incongruent reactions consuming mica (Gao et al., 2016b; Johannes and Holtz,

1996). However, early experiments on compositions varying from greywacke to calcareous

pelite were H2O-saturated and showed that large melt volumes may be generated at near-

solidus conditions producing melts that range in composition from tonalite to granite
Journal Pre-proof

(Winkler, 1957; Winkler and Von Platen, 1958, 1960; Winkler and von Platen, 1961).

Experiments by Wyart and Sabatier (1959) at low P (~0.2 GPa) and 800 °C produced

cordierite and spinel as solid (peritectic) biproducts. Importantly, these authors recognised the

fundamental role of partial melting and melt drainage in the creation of the large-scale

differentiation of the crust into granitic upper and residual granulitic lower portions (Wyart

and Sabatier, 1959).

Due to their high fertility and sensitivity to metamorphic conditions, numerous

experimental studies have been undertaken to investigate the fluid-absent partial melting of

f
oo
metapelitic compositions across a wide range of P–T–X conditions (Carrington and Harley,

1995; Gardien et al., 1995; Grant, 2004; Le Breton and Thompson, 1988; Nair and Chacko,
pr
2002; Patiño Douce and Harris, 1998; Patiño Douce and Johnston, 1991; Pickering and
e-
Johnston, 1998; Spicer et al., 2004; Stevens et al., 1997; Vielzeuf and Holloway, 1988). At
Pr

pressures above 0.4–0.5 GPa, fluid-absent melting initially occurs by reactions consuming

mainly muscovite plus quartz to produce sillimanite (or kyanite), K-feldspar, minor biotite
al

and melt, the latter with strongly peraluminous compositions that show a close
n

correspondence with Himalayan leucogranites (Le Breton and Thompson, 1988; Patiño
ur

Douce and Harris, 1998; Vielzeuf and Holloway, 1988; see section 7). At higher temperature,
Jo

or at lower pressure, major melting occurs by the fluid-absent breakdown of biotite,

aluminosilicate and quartz to produce K-feldspar, melt and cordierite at low P and/or garnet

at higher P; orthopyroxene and/or aluminous spinel generally form at higher T depending on

pressure and bulk composition, in particularly XMg and XAl (= molar Al2O3/(Al2O3 + FeO +

MgO)) (Carrington and Harley, 1995; Gardien et al., 1995; Grant, 2004; Le Breton and

Thompson, 1988; Nair and Chacko, 2002; Patiño Douce and Harris, 1998; Patiño Douce and

Johnston, 1991; Pickering and Johnston, 1998; Spicer et al., 2004; Stevens et al., 1997;

Vielzeuf and Holloway, 1988; see section 7). The melts produced are mostly peraluminous.
Journal Pre-proof

Experimental studies on greywacke starting compositions have concentrated on

peraluminous (mostly continental-arc-derived) rather than metaluminous (mostly oceanic-

arc-derived) variants. Conrad et al. (1988) undertook H2O-saturated and -undersaturated

experiments at 1.0 GPa on a peraluminous greywacke and compared these with those on a

metaluminous dacite, both from the Taupo Volcanic Zone in New Zealand. Fluid-absent

melting of the greywacke consumed biotite, quartz and plagioclase to produce melt with

gedrite (orthoamphibole), garnet and/or orthopyroxene depending on temperature. Fluid-

present melting of the same composition preceded mostly through consumption of H2O,

f
oo
plagioclase and quartz. Experiments investigating the fluid-absent melting of greywacke over

a wide range of temperatures and pressures up to 2 GPa showed that reactions preceded by
pr
breakdown of biotite, plagioclase and quartz to produce mainly orthopyroxene with or
e-
without cordierite, spinel or garnet depending on composition (in particular XMg) and pressure
Pr

(Montel and Vielzeuf, 1997; Patiño Douce and Beard, 1996; Vielzeuf and Montel, 1994; see

section 7). Melt compositions evolved to lower K2O/Na2O and higher FeO + MgO + TiO2
al

contents with increasing P and lower ASI with increasing T (Montel and Vielzeuf, 1997;
n

Patiño Douce and Beard, 1996).


ur
Jo

4.2. Melting in unary (C = 1) and binary (C = 2) systems

At any given pressure in a one-component system, the reaction by which a pure solid

may melt (i.e. break down) to produce a liquid of the same composition is univariant (C = 1,

P = 2) and occurs at a fixed temperature known as the melting point. For example, at

atmospheric pressure (P = 0.1 MPa or 1 bar), solid H2O starts to breaks down (react) to liquid

H2O at 0 C, the melting point of ice. Adding heat drives the melting reaction, but the two-

component system (ice + water) remains at 0 C until all of the ice has been consumed (C =
Journal Pre-proof

1), after which temperature may rise. However, in multicomponent systems, such as rocks,

the transformation of the system from 100% solid to 100% liquid occurs over a range of

temperatures, the melting interval. The melting interval extends from the temperature at

which melt first forms, termed the solidus, to the temperature at which the system is entirely

molten, called the liquidus.

These different features are illustrated with reference to Fig. 7a, which shows phase

relations in a simple fictive binary system (C = 2) at fixed pressure. This is an example of a

T–X phase diagram, in which temperature increases vertically upwards and composition (X)

f
oo
changes along the horizontal axis between two end-member compositions, in this case A and

B. Melting relations for the pure phase A in the one-component system (C = 1 and X = A) are
pr
shown along the left-hand-edge of Fig. 7a. For example, at temperature TSS the system is
e-
entirely solid and consists of the pure phase A. On heating, T increases until the melting point
Pr

of A is reached (MPA), where the addition of heat leads to the breakdown of solid phase A to

a liquid with composition A by the univariant reaction A → Liq, which occurs at fixed T until
al

all solid A is consumed. A similar behaviour is shown for the one-component system with the
n

bulk composition B, although the melting point of B (MPB) is at higher T.


ur

Now consider an intermediate bulk composition, X1, in the two-component system A–


Jo

B. At temperature TSS, the system consists of two solid phases, A and B, whose proportions

are given by the lever rule. The proportion of A is the distance between X1 and B (= a0)

divided by the distance between A and B (AB), and the proportion of B is bo/AB, so the

subsolidus assemblage consists of 60% A and 40% B. With an increase in heat, temperature

rises and the assemblage and proportion of A and B remain the same until the solidus is

encountered at temperature TS. At this point, melt with a composition half way between A

and B (XE) begins to form at the eutectic point, EP, by the univariant reaction:
Journal Pre-proof

A + B → Liquid (10)

This first-formed melt is known as the eutectic melt or minimum melt. In reaction 10, the

solid reactants (A + B) form a melt of the same composition (AB), and the reaction is said to

be congruent. This reaction continues at fixed temperature, TE (meaning it is a discontinuous

reaction), until all B is consumed, at which point the assemblage consists of 20% A and 80%

melt ((a1/(a1 + l1) and l1/(a1 + l1), respectively) and temperature may rise again within the

field A + Liq. As temperature rises, the proportion of A decreases, and of melt increases by

f
oo
the divariant congruent reaction A → Liq, until all A is consumed at the liquidus at

temperature TL. As this reaction occurs over a range of temperatures, it is said to be a


pr
continuous reaction. In almost all cases, rocks melt via continuous (F > 1) reactions.
e-
For any intermediate bulk compositions more enriched in B than the composition of
Pr

XE (i.e. to the right of XE), for example X2, reaction 10 consumes all A, and melting continues

by the reaction B → Liq until the liquidus is reached. Although all intermediate compositions
al

begin to melt at the same temperature, that of the eutectic point, the temperature of the
n

liquidus, and consequently the melting interval, are a function of composition. For the unique
ur

bulk composition XE, the solidus and liquidus coincide.


Jo

Now consider a two-component system that includes an intermediate phase, AB (Fig.

7b). In Fig. 7b, the three phases, A, AB and B, each melt congruently, but at different melting

points (MPAB > MPA > MPB). The solidus and eutectic point (LE1) for intermediate

compositions in the subsystem A–AB (TE1) are at higher T than those in the subsystem AB–B

(TE2). Otherwise, this T–X diagram can be considered as two modified versions of Fig. 7a

joined together.

Another binary phase diagram with an intermediate phase AB is shown in Fig. 7c.

However, rather than melting congruently to a liquid of the composition of AB at fixed T, as


Journal Pre-proof

in Fig. 7b, the pure solid with composition AB (X1) melts over a temperature interval. As it

heats up it reaches the solidus at TS1 and begins to melt via the univariant reaction:

AB → A + Liquid, (11)

remaining at temperature TS1 until all AB is consumed. Temperature then rises and melt

forms by the congruent continuous reaction A → Liquid until the liquidus is reached. A

melting reaction such as reaction 11, by which a solid phase (or phases; here just AB) breaks

f
oo
down to form melt plus a different solid (or solids; here just A) is an example of incongruent

melting. The composition of the melt (here XP) formed by the incongruent reaction is
pr
governed by the location of the peritectic point, P, and the solid phase that is produced with
e-
the melt (here A) is called a peritectic mineral.
Pr

Examination of Fig. 7c shows that all intermediate bulk compositions to the left of XP

undergo incongruent melting. For those to the left of AB, for example X2, the subsolidus
al

assemblage A + AB first melts by reaction 10 at TS1, then by the reaction A → Liquid. For
n

those intermediate compositions between X1 and XP, for example X3, a subsolidus assemblage
ur

of AB + B melts first by the congruent discontinuous reaction AB + B → Liquid at TS2, then


Jo

by reaction 11, then by the reaction A → Liquid until the liquidus is reached. Bulk

compositions more enriched in B than XP only undergo congruent melting.

For the interested reader, phase equilibria in simple (binary, ternary and quaternary)

melt-bearing systems are described elegantly and in detail by Cox et al. (1979). Although

these are described mostly from the perspective of igneous petrology (i.e. heat out driving

crystallisation), those describing equilibrium crystallisation are the inverse of equilibrium

melting described above.


Journal Pre-proof

4.3. Types of phase diagram

4.3.1. Compatibility diagrams and petrogenetic grids

We have already discussed a simple P–T phase diagram in the unary system Al2SiO5

(Fig. 6). This diagram shows three univariant (discontinuous) reactions emanating from an

invariant point––otherwise called a reaction bundle––for the aluminosilicate minerals. More

complex diagrams in larger chemical systems, known as petrogenetic grids, involve multiple

invariant points and reaction bundles, the latter of which are sometimes termed

f
oo
Schreinemakers’ bundles after the Dutch physical chemist Franciscus Schreinemakers, who

developed the rules governing their geometry (Schreinemakers, 1915; Zen, 1966).
pr
We illustrate the logic behind deciphering reactions and the graphical construction of
e-
reaction bundles and petrogenetic grids in P–T space using a simple example in the four-
Pr

component KASH (K2O–Al2O3–SiO2–H2O) subsystem (Fig. 8). Figure 8a is a ternary

compatibility diagram showing the compositions of the phases muscovite, K-feldspar,


al

sillimanite, melt (L) and H2O projected from quartz (the SiO2 apex). Projecting from quartz
n

allows us to easily establish graphically (i.e. on a piece of paper or a screen) what the
ur

univariant reactions are, but only those involving quartz––in other words, quartz is
Jo

considered to be in excess. Note that in this simple example, the only solid solution phase is

melt, which can contain variable quantities of H2O (it lies on a straight line whose extension

intersects the H2O apex).

In this four-component system, and considering six phases, the phase rule tells us

there is only one invariant equilibrium where all phases coexist at a single P–T point. Each

univariant reaction line stemming from the invariant point involves five phases, meaning

there are six reactions in total, each one excluding one of the six phases. Considering quartz
Journal Pre-proof

to be in excess leaves five univariant reactions, which by convention are labelled with the

absent phase in square brackets.

We first consider the reaction that excludes melt, i.e. [L]. With quartz in excess and

excluding melt, the compatibility diagram for the remaining phases (inset to Fig. 8b) has

muscovite at its centre and K-feldspar, sillimanite and H2O at each apex. As muscovite is

located entirely within the area bounded by the other three phases, we deduce the reaction to

be: Ms → Sil + Kfs + H2O. Petrographic observations (and alternative graphical projections)

show that quartz is consumed along with muscovite, such that the full subsolidus dehydration

f
oo
reaction is:

pr
Ms (+ Qz) → Sil + Kfs + H2O [L]. (12)
e-
Pr

This is a terminal reaction in KASH, at temperatures above which muscovite is no longer

stable (in quartz-bearing rocks). The Clapeyron equation discussed earlier predicts that
al

reaction 12 has a positive slope with H2O on the high-T side. Figure 8b shows reaction 12
n

plotted (qualitatively) in P–T space extending from lower to higher T. However, common
ur

sense tells us that, as we increase temperature (and P) along this univariant line, at some
Jo

point the assemblage will begin to melt. With the addition of melt (L), the assemblage

becomes invariant (and [L] terminates at the invariant point IP in Fig. 8b). We may draw

reaction 12 continuing through IP to higher T but, as H2O is no longer stable with respect to

melt, it is termed the metastable extension of this reaction (and is dashed). With one reaction

plotted, we complete the reaction bundle simply by following Schreinemakers’ metastable

rule: “The metastable extension of the [phase-absent] reaction must fall in the sector in

which that phase is stable in all possible assemblages” (Spear, 1995).


Journal Pre-proof

Next consider the H2O-absent reaction [H2O]. With quartz in excess and excluding

H2O, muscovite is located within the triangle joining sillimanite, K-feldspar and melt, and the

full reaction is:

Ms (+ Qz) → Sil + Kfs + L [H2O]. (13)

The Clapeyron equation predicts that this fluid-absent incongruent melting reaction has a

positive slope with melt on the high-T side, and must also must emanate from IP. However,

f
oo
Schreinemakers’ rule tells us that the metastable extension of reaction 12 ([L]) must be on the

same side as melt in reaction 13 [H2O], and that the metastable extension of reaction 13 must
pr
be on the same side of the line as H2O in reaction 12, giving us the topology shown in Fig.
e-
8c.
Pr

Following the same logic allows us to deduce the muscovite-absent and sillimanite-

absent reactions to be:


n al

Sil + Kfs + H2O (+ Qz) → L [Ms] and (14)


ur

Ms + Kfs + H2O (+ Qz) → L [Sil], (15)


Jo

respectively. These fluid-present congruent melting reactions lie on the low-P (Fig. 8d) and

high-P (Fig. 8e) sides of the metastable extensions of reactions 12 and 13, respectively, and

together define the H2O-saturated (‘wet’) solidus. Again, with quartz in excess these are

terminal (albeit in the retrograde sense) reactions as, by definition, melt cannot be stable at

temperatures below the solidus. Melt, as the sole product, lies on the high-T (and high S) side

of reactions 14 and 15, but with H2O as a reactant, these equilibria are associated with a

negative volume change, and consequently have a steep negative P/T. Importantly, as the
Journal Pre-proof

solubility of H2O in melt is positively correlated with P (Hamilton et al., 1964; Holtz et al.,

2001), the slope of the H2O-saturated solidus becomes shallower with decreasing pressure.

The final reaction is that excluding K-feldspar (Fig. 8f). In the compatibility diagram,

rather than forming a triangle surrounding one phase, the four phases participating in the

reaction define a quadrilateral with muscovite and H2O defining one pair of opposing

corners, and sillimanite and melt the other. The crossed tie-lines from each pair of corners

define the reaction:

f
oo
Ms + H2O (+ Qz) → Sil + L [Kfs]. (16)

pr
This reaction is not a terminal reaction, as either muscovite or H2O can be consumed
e-
depending on their relative abundance, which is a function of the bulk composition. A
Pr

petrogenetic grid in the enlarged KFMASH system appropriate to metapelitic rocks is shown

and discussed in section 7.


al

It is important to note that petrogenetic grids show all of the possible univariant
n

reactions a system might encounter (or ‘see’), projected into the pressure–temperature plane
ur

of interest, hence their alternative name of P–T projections. However, specific bulk
Jo

compositions may see only a portion of any reaction, or none of it. In addition, petrogenetic

grids require that the system includes only a relatively small number of components such that

univariant equilibria can be calculated. Although this is possible for metapelites via the

simplified KFMASH system (Carrington and Harley, 1995; Grant, 1985; Thompson, 1982;

White et al., 2001, 2014a), based on the pioneering work of J. B. Thompson (Thompson Jr,

1957), equilibria governing the partial melting of metabasic rocks in plausible chemical

systems (e.g. NCFMASH) are of higher variance and this approach is unsuitable. Thus, while

petrogenetic grids are extremely useful, for example in illustrating qualitatively which phases
Journal Pre-proof

might appear or disappear during melting along a specified P–T trajectory, and are key to

understanding the underlying topology of phase diagrams, they generally do not provide

useful quantitative constraints on the behaviour of natural rocks during anatexis (White et al.,

2007).

4.3.2. P–T–X pseudosections

Within the context of petrology, the term pseudosection owes its origin to Roger

Powell (Powell, 1978) to describe equilibrium phase diagrams depicting stable multivariant

f
oo
assemblage fields for specified bulk rock compositions (Powell et al., 1998), although such

diagrams were in use in the early 1970s (Hensen, 1971). Pseudosections are two-dimensional
pr
(Cartesian) diagrams mostly constructed in P–T–X space, such that the three main types are
e-
P–T, T–X and P–X pseudosections. Although others prefer the terms phase-diagram sections
Pr

(Connolly and Petrini, 2002; Tinkham and Ghent, 2005a), equilibrium assemblage diagrams

(Capitani and Petrakakis, 2010) or, in the case of P–T pseudosections, isochemical phase
al

diagrams, the ‘pseudo’ prefix is intended to indicate that the compositions of the phases
n

predicted to be stable generally do not plot within the compositional plane of interest (i.e.
ur

‘pseudo’ as in ‘almost’ or ‘similar to’ rather than ‘false’ or ‘spurious’). Notwithstanding, over
Jo

the past few decades calculated pseudosections have become the tool of choice in providing

quantitative constraints on the (evolving) P–T (and other) conditions during metamorphism

and partial melting. Much of what follows in this review regarding the behaviour of crustal

rocks during anatexis is illustrated with reference to pseudosections.

The most commonly used diagrams are P–T pseudosections that are constructed for a

fixed bulk composition (X), where the bulk composition most commonly was measured via

XRF analysis of a whole-rock powder or estimated based on the abundance (through image

analysis) and composition (through electron probe microanalysis (EPMA)) of the constituent
Journal Pre-proof

minerals. P–T pseudosections are essentially maps showing the equilibrium assemblages that

are predicted to be stable for a specified rock composition in a P–T window of interest. An

example is shown in Fig. 9a, a P–T pseudosection drawn for a synthetic composition in the

six-component NKFMASH system (the compositions used to calculate this and all other

pseudosections in subsequent figures in terms of mol.% oxides normalised to 100% are

provided in Table 1), as used in the discussion of the experiments of Taylor et al. (2015b),

and calculated for the P–T range from 0.5 and 1.5 GPa, and from 750 to 1200 C.

f
The logic behind these diagrams and reading them is straightforward. The diagram is

oo
clearly subdivided by the depth of colour into different assemblage fields, which are labelled

pr
with the stable phases, and are adjacent to fields that are either paler or darker in shade. By

convention, the depth of shading of the assemblage fields reflects increasing variance (F).
e-
Crossing into a paler field requires the production of a new phase (decreasing variance by 1),
Pr

whereas moving into a darker field indicates that a phase is consumed (increasing variance by

1). For example, for the experimental composition used, at 750 C and 1.0 GPa (point A in
al

Fig. 9a) the equilibrium assemblage should be kyanite–garnet–biotite–alkali feldspar–quartz–


n

liquid, a trivariant (F = 3) assemblage in the modelled system. Heating at fixed P thereafter


ur

first predicts the consumption of kyanite at ~755 C, associated with moving into the darker
Jo

quadrivariant (F = 4) field containing garnet–biotite–alkali feldspar–quartz–liquid. Further

heating leads to the appearance of orthopyroxene then the consumption of biotite over a

narrow temperature interval (~3 C) at ~810 C across the paler narrow trivariant field

containing garnet–orthopyroxene–biotite–alkali feldspar–quartz–liquid. With further heating,

first orthopyroxene then quartz is consumed until at 1000 C, 1.0 GPa (point B in Fig. 9a) the

composition should contain just garnet and melt (F = 7). Isothermal decompression from

point B predicts the reappearance of orthopyroxene then spinel, and then exhaustion of garnet

at 0.65 GPa.
Journal Pre-proof

Inspection of Fig. 9a shows that many of the lines bounding the assemblage fields

meet at points, from which exactly four lines emanate. The reason for this is self-evident

when examining phase relations within the dashed black box in Fig. 9a (see inset). This

region essentially comprises the intersection between two lines across which a mineral

disappears––one with a steep positive P/T that, when crossing to higher T, leads to the

disappearance of alkali feldspar, and the other with a shallower negative P/T that, on

crossing to higher P, leads to the disappearance of orthopyroxene. Thus, there is a pale low-

f
variance field (F = 5; Grt–Opx–Fsp–L) that is bounded by two F = 6 fields, one lacking alkali

oo
feldspar and another lacking orthopyroxene, and which lies opposite an F = 7 field that lacks

pr
both minerals. With the exception of univariant lines, this logic holds for all pseudosections–

–i.e. at an intersection, the lowest-variance field with variance n is bounded on either side by
e-
fields of variance n + 1 and lies opposite a field of variance n + 2.
Pr

In addition to the assemblage field boundaries, which correspond to the conditions at

which the abundance (‘mode’) of a phase becomes zero, pseudosections can also be
al

contoured for the mode and composition of phases, termed mode isopleths (or isomodes) and
n

compositional isopleths, respectively. As an example, the mode isopleths for biotite at 6, 4,


ur

and 2 mol.% are shown (dotted red lines in Fig. 9a; note that THERMOCALC, which was used
Jo

to calculate Fig. 9, outputs the abundance of phases as mol.% based on one-oxide molar

proportions such that they approximate vol.%). These provide additional and useful

information, in particular in illustrating the rate of change (consumption or growth) of a

phase within the P–T window of interest. For example, at point A, a little more than 6 vol.%

biotite is predicted in the modelled composition. Heating thereafter causes biotite to be

consumed (i.e. biotite is a reactant) at a more-or-less steady rate until, as the path intersects

the narrow trivariant field in which orthopyroxene appears, there is a little more than 2 vol.%

remaining. Thus, over a temperature interval of ~60 C (750 to 810 C), ~4 vol.% biotite has
Journal Pre-proof

been consumed at a rate of 1 vol.% every ~15 C. However, the trivariant field containing

coexisting biotite and orthopyroxene is very narrow, and the remaining 2 vol.% biotite is all

consumed over ~3 C, representing a reaction rate (here taken to mean the rate of change of

the abundance (mode) of a reactant or product as a function of changing T and/or P) an order

of magnitude or so faster.

This illustrates an important point, that most of the (re)action in pseudosections

occurs within the lowest-variance fields, as shown by the closely-spaced mode isopleths of

f
biotite in the divariant (F = 2) field in the bottom left corner where the assemblage

oo
cordierite–garnet–orthopyroxene–biotite–alkali feldspar–quartz–liquid is stable. Across this

field >6 vol.% biotite is consumed (with quartz) over ~3 C by a fluid-absent incongruent
pr
hydrate-breakdown reaction producing mostly melt and peritectic orthopyroxene.
e-
In an ideal world, Fig. 9a should replicate the assemblages that the experimentalist
Pr

might expect at any P–T of interest. For example, to examine Fe–Mg exchange between

garnet, orthopyroxene and melt without the complication of the presence of any other solid
al

phases, the experimentalist should be able to do so by performing an experiment at 0.7 GPa


n

and 950 C (the star in Fig. 9a). However, in constructing the pseudosection, some of the
ur

end-member thermodynamic data or a–X models (or solution models) describing the
Jo

energetics of mixing in the solid solution minerals may be poorly constrained, perhaps due to

a lack of suitable experiments that underpin these models. In addition, it is almost inevitable

that experimental charges will undergo some oxidation (converting some FeO to Fe2O3), and

possible or even likely that some minerals will have difficulty nucleating or dissolving over

the timescale of an experiment, such that equilibrium in may not be attained in part or all of

the experimental charge (for more details see White et al., 2011).

Now let us assume that the experimental petrologist wishes to know whether it is

possible to grow cordierite but no garnet in an experiment at 0.9 GPa (the dashed white line
Journal Pre-proof

in Fig. 9a) simply by adjusting the XMg (molar MgO/(MgO + FeO)) of the original

composition (in which XMg = 4.54/10.05 = 0.45). To address this problem, the experimentalist

can construct a T–X pseudosection at a fixed pressure of 0.9 GPa between two end-member

compositions in which XMg ranges, for example, from 0.1 to 0.9 by adjusting the FeO and

MgO contents accordingly (to 9.04 mol.% FeO and 1.01 mol.% MgO and 1.01 mol.% FeO

and 9.04 mol.% MgO, respectively). The result is shown in Fig. 9b, an isobaric T–X

pseudosection that can be read exactly as described earlier, but in which T increases upwards

in the y-direction and bulk composition (X in terms of XMg) varies in the x-direction. The

f
oo
vertical dashed white line in Fig. 9b represents the original composition used (Fig. 9a), such

that the phase equilibria calculated from left to right along the white dashed line in Fig. 9a are
pr
identical to those calculated from bottom to top along the equivalent line in Fig. 9b.
e-
Inspection of Fig. 9b indicates that cordierite is stable only in compositions with XMg
Pr

greater than ~0.7 and at temperatures between ~800 and 950 C (in the region bounded by the

thick black line). Only in compositions with XMg greater than 0.83 (to the right of the dashed
al

thick black line) is cordierite but no garnet predicted. Thus, the experimentalist might run a
n

charge on a composition with, for example, XMg = 0.9 at 920 C and 0.9 GPa (the white star
ur

in Fig. 9b).
Jo

Hopefully, the logic behind constructing a P–X pseudosection (not shown) for

variable XMg at fixed T, for example, should be clear. For this example, with XMg = 0.45, the

resulting phase equilibria correspond to a vertical line at the appropriate T in Fig. 9a.

4.3.3. Other phase diagrams

Although most calculated reactions, grids and pseudosections in the petrological

literature are P–T or T–X and P–X diagrams, other diagrams can be used to investigate

multiple compositional parameters at a fixed P–T condition (e.g. X–X diagrams; (Riesco et
Journal Pre-proof

al., 2005; White et al., 2003). Furthermore, other thermodynamic variables might be chosen

as the axes of choice, for example volume, entropy or activity (Powell et al., 2005). Although

such diagrams are generally less intuitive, they can be used to illustrate critically important

processes, including adiabatic or isentropic (constant S) melting of upwelling mantle

(Asimow et al., 1997; Iwamori et al., 1995; McKenzie and Bickle, 1988), and thermal

buffering of the crust due to strongly-endothermic partial melting reactions (Bea, 2012;

Schorn et al., 2018; Stüwe, 1995). Diagrams using chemical potential () as one or both axes

f
are particularly useful in understanding the spatial organisation of minerals in reaction

oo
textures, such as partial pseudomorphs, coronae and/or symplectites, which are particularly

pr
common in formerly melt-bearing granulites, including UHT rocks (Doukkari et al., 2018;

Gallien et al., 2012; Mictchell et al., 2020; Powell et al., 2019; White and Powell, 2011;
e-
White et al., 2008). However, the assumptions and simplifications required in their
Pr

construction mean that chemical potential diagrams are of limited use in providing

quantitative constraints on the processes of partial melting.


al

Of particular use in modelling partial melting are plots of molar proportion versus
n

temperature or pressure (Yakymchuk and Brown, 2014b), or along specified thermobaric


ur

ratios (Johnson et al., 2017), now colloquially called ‘modebox’ diagrams (Holland et al.,
Jo

2013). They are important in visualising how the abundance (mode) of phases with a system

(or rock) are expected to evolve during melting along a specified path (Johnson et al., 2008;

Yakymchuk et al., 2017). An example is given in Fig. 9c, which shows the changing

abundances of phases predicted along the isobaric (1.0 GPa) path from 750 to 1000 C (point

A to point B) on Fig. 9a. Along this path, the proportion of melt increases from ~0.3 to ~0.8

(30 to 80 vol.%), largely by the breakdown of biotite, quartz and alkali feldspar. A small

amount of peritectic orthopyroxene is produced by biotite breakdown, then consumed at


Journal Pre-proof

higher T. The abundance of garnet changes little with increasing T, indicating that garnet is

neither a major reactant nor product of melting along the chosen heating path.

4.4. Calculating phase diagrams

Having outlined the thermodynamic basis of phase equilibria, and the types of phase

diagrams most commonly drawn to constrain the conditions and processes of partial melting,

in particular pseudosections, we now outline the requirements and practicalities behind their

f
oo
calculation. Clearly, solving the equilibrium relationship (equation 6) is key, which requires

end-member thermodynamic data, a–X models describing the energetics of mixing in solid-

solution phases and appropriate software.


pr
e-
Pr

4.4.1. End-member data

Values for the thermodynamic properties of end members (to satisfy the left-hand-
al

side of equation 5) that describe most of the common rock-forming minerals, and more
n

recently fluid, melt and aqueous species, have existed for more than four decades (Berman,
ur

1988; Ghiorso and Sack, 1995; Gottschalk, 1996; Helgeson et al., 1978; Holland and Powell,
Jo

1985, 1990, 1998a, 2011; Miron et al., 2016; Naumov et al., 1974; Saxena et al., 2012). The

data sets most widely used currently, at least in the metamorphic community, are those of

Holland and Powell (1998a) and Holland and Powell (2011). These data sets represent a

collation of a large body of thermodynamic data including experimentally-determined phase

equilibria in simple and more complex chemical systems, calorimetric measurements on pure

end members and, increasingly, natural data from coexisting phases in rocks inferred to have

attained equilibrium. Importantly, the absolute values and uncertainties on the various

thermodynamic properties, which include the standard enthalpy of formation (ΔfH), entropy
Journal Pre-proof

(S), heat capacity (CP), and volume (V) as a function of temperature and pressure (the

coefficients of thermal expansion, and compressibility, ) are optimised to provide the best

statistical fit to the disparate sources of data, such that the full data set provides an internally

consistent thermodynamic framework for calculations. The different approaches used in the

derivation of the internally consistent data are summarised by Gottschalk (1996).

4.4.2. Activity–composition (solution) models

f
The energies associated with mixing between end members in variably complex solid

oo
solution phases (to satisfy the right-hand-side of equation 5) are themselves complex.

pr
Whereas some data are relatively well-constrained experimentally, for example the

interaction energies (Ws) relating to Fe2+–Mg substitution in minerals such as olivine,


e-
clinopyroxene, garnet and biotite (Bhattacharya et al., 1992; Ellis and Green, 1979; O'Neill
Pr

and Wood, 1979; von Seckendorff and O'Neill, 1993), in many other cases there is little or no

data, and the associated energetics of mixing must be estimated. This is achieved largely
al

using a heuristic approach–– we fix those things for which data exist and/or we understand
n

well, then make educated guesses to fit those parameters for which there is little or no data
ur

and/or that we do not understand well (Powell et al., 2014).


Jo

For example, based on existing experimental data on specific ferromagnesian

minerals, we may be able to make some general assumptions regarding the energetics of

mixing between Fe2+ and Mg in all ferromagnesian minerals, and solve for other

compositional variables. In essence, adjusting the interaction energies (Ws) within limits that

we think are sensible is a means of ‘tweaking’ or ‘tuning’ the a–X models such that they

behave as we think they should. For instance, many decades of experience tells us the order

of appearance of new minerals we might expect in a sequence of pelitic rocks undergoing

progressive prograde regional metamorphism (Barrow, 1893; Read, 1952), and the relative
Journal Pre-proof

XMg (Mg/(Mg + Fe2+)) of coexisting ferromagnesian minerals (Droop and Harte, 1995; Harte

and Hudson, 1979). If the sequence of isograds or composition of coexisting phases predicted

by calculations using the relevant a–X models is inconsistent with these observations, one or

more may need to be adjusted (Cesare et al., 2003; Johnson et al., 2003a; Pattison et al.,

2002). The a–X models, also known as solution models, are many and varied in their

applicability and complexity (see for example:

http://www.perplex.ethz.ch/perplex/datafiles/solution_model.dat), and many different models

may exist for particular minerals, such as garnet.

f
oo
4.4.3. Software
pr
Several software programs exist for calculating phase equilibria to address
e-
petrological problems (Asimow and Ghiorso, 1998; Connolly, 1990, 2005; de Capitani and
Pr

Petrakakis, 2010; Ghiorso et al., 2002; Gualda et al., 2019; Powell et al., 1998; Powell and

Holland, 1988; Spear and Menard, 1989). Increasingly, applications are being developed that
al

automate or refine particular aspects of the process, such that calculating phase diagrams is
n

becoming ever more ‘user friendly’ (Duesterhoeft and de Capitani, 2013; Mayne et al.,
ur

2020b; Pearce et al., 2015; Xiang, 2020; Xiang and Connolly, 2021). Here we briefly review
Jo

the three computer programs (or rather families of programs) most commonly used to

calculate subsolidus and suprasolidus phase equilibria in metamorphic studies, viz.

THERMOCALC (Powell and Holland, 1988), Perple_X (Connolly, 1990, 2005) and THERIAK-

DOMINO (de Capitani and Petrakakis, 2010). By metamorphic studies we mean scenarios

mostly involving the addition of heat to solid-dominated systems, of which partial melting is

the most important. We do not discuss the MELTS family of software, which is mainly

designed to investigate crystallising and fractionating melts and magmas (liquid-dominated

systems), representing systems from which heat is being removed, and which might be
Journal Pre-proof

considered closer to the realm of igneous studies if such classifications are useful (Asimow

and Ghiorso, 1998; Ghiorso et al., 2002; Ghiorso and Sack, 1995; Gualda et al., 2019).

Each of the three main software packages discussed here has benefits and limitations

that may/will change as the user gains experience and/or based on the problem being

addressed. In constructing pseudosections, Perple_X and THERIAK-DOMINO adopt a

method of Gibbs energy minimisation and are (at least superficially) rather easy to use. In

essence, the user selects the end member thermodynamic dataset to use, provides the bulk

composition or bulk compositional range in terms of the components (usually major oxides)

f
oo
of interest, defines the P–T window of interest, and selects the solution models for all the

phases they think might be stable within that window. The software then calculates the
pr
Gibbs’ free energy for all possible combinations of the chosen phases at regular spacings
e-
within the P–T window of interest (i.e. on a grid), for which the combination with the lowest
Pr

Gibbs energy is taken to be the equilibrium assemblage. After some time, which may be

seconds or many days depending on the complexity of the calculation, a pseudosection is (or
al

can be) plotted automatically using the output data. Provided the user has correctly included
n

solution models for all of the appropriate phases, the output diagram should correspond to the
ur

stable (lowest G) configuration.


Jo

THERMOCALC uses a different approach to calculate phase diagrams involving solid

solutions, which involves solving sets of non‐linear equations. In the construction of

pseudosections, using a bulk composition or bulk compositional range specified by the user,

THERMOCALC reads the end-member thermodynamic data and solution models to solve

statements of equality of chemical potential (i.e. equilibrium) among the end-members of a

specific combination of phases (assemblage) that the user thinks might be stable. If it is able

to find solutions, it outputs the equilibrium compositions and abundance of those phases at

the user-specified (generally P–T–X) conditions. However, the chosen combination of phases
Journal Pre-proof

need not be the most stable (lowest G) assemblage. Consequently, the user should have some

confidence that they are calculating a stable rather than metastable equilibrium based on

experience or on examination of a phase diagram for a similar bulk composition under

similar P–T–X conditions. In THERMOCALC, pseudosections are built up manually, line by

line and point by point, following the logical rules set out earlier in this section.

Results calculated using any one of these programs are no better or worse that those

generated using the others––for a given database of end-member thermodynamic data,

solution models and user-defined P–T–X constraints, they should produce identical phase

f
oo
diagrams. However, there are pros and cons, and perils and pitfalls, pertaining to the usability

and functionality of each. As stated previously, Perple_X and THERIAK-DOMINO are very
pr
easy to run, and a user with almost no experience can generate a pseudosection with minimal
e-
effort. However, both programs have a large library of solution models available, meaning it
Pr

is easy for users with less experience to pick inappropriate or incompatible models for one or

more phases, such that the resulting diagrams, whilst appearing perfectly reasonable, might
al

have little bearing on the problem at hand (see Gervais and Trapy, 2021) for a comparison of
n

different end-member data and a–X models). However, once a diagram is calculated,
ur

Perple_X and THERIAK-DOMINO have far more flexibility, for example in easily
Jo

extracting other data related to the chosen system or its component phases, such as V, S, a,

density (), and P- and S-wave velocities.

A limitation of THERMOCALC is that it only works with those solution models that

were calibrated specifically against one or more of the Holland and Powell end-member

thermodynamic databases (Holland and Powell, 1990, 2011), now collectively termed ‘HPx-

eos’ (Green et al., 2016; Holland et al., 2018; White et al., 2014a, b). It is more difficult to get

started on phase diagram calculations using THERMOCALC, but doing so arguably leads to a

deeper understanding of phase equilibria than other programs as the diagrams have to be
Journal Pre-proof

constructed from scratch. In terms of functionality, it is more difficult to extract other useful

data from the THERMOCALC output, although a great deal of progress has been made recently

and is ongoing.

Ultimately, the fundamentally different way in which THERMOCALC solves the

equilibrium relationship coupled with advances in the complexity of solution models and

expansion of the compositional systems that can be modelled mean it will probably outlast its

Gibbs energy minimiser competitors. To quote (or paraphrase) Jamie Connolly (ETH Zürich)

in response to a user who wished to undertake calculations using the complex new solution

f
oo
models of Holland et al. (2018): “…Perple_X has been teetering on the edge of dimensional

oblivion for a decade… the root(s) of the problem are polymorphic solutions, such as the
pr
pyroxenes and amphiboles… if worse comes to worst and it turns out that it is too
e-
(computationally) costly to resolve the composition space of solutions such as the Green et
Pr

al. (2016) amphibole model, then Perple_X will not be robust (its only virtue) and in that

case one might as well rely on expertise-based, non-linear, optimization which is both faster
al

and more precise” (see: https://groups.io/g/PerpleX/message/2201).


n
ur

4.5. Getting started – perils and pitfalls


Jo

We end this section by discussing some of practicalities that should be considered

before embarking on the construction of a pseudosection for a natural sample about which

little is known, along with some of the assumptions and limitations involved. While many of

these should be self-evident, it is worthwhile to clarify them.

4.5.1. Sample selection and petrographic characterisation


Journal Pre-proof

A first-order consideration is the question of which sample(s) to model. The first

assumption is that the sample under consideration has attained equilibrium on some scale.

Having identified an equilibration volume, which may be an entire hand-specimen and/or thin

section, and assuming the purpose is to constrain peak P–T conditions or the P–T path, it is

almost always preferable to attempt to identify samples with the lowest variance assemblage

(i.e. those containing the most minerals), as these generally occupy smaller (more tightly

constrained) assemblage fields than higher variance assemblages (see Fig. 9a, b and later

sections). In many cases, low-variance assemblages occur in rocks containing one or more

f
oo
conspicuous porphyroblasts. At a particular location (outcrop) and for a particular generalised

protolith (e.g. metabasite or metapelite), changes in variance reflect variations in bulk


pr
composition, which can be subtle such that porphyroblasts may only occur rarely in some
e-
layers. Where possible, it is best to collect several samples covering a range in bulk
Pr

compositions from any locality, as P–T conditions will generally be more tightly constrained

using the intersection of several calculated assemblage fields.


al

Once suitable samples have been collected, the next logical step is to make thin
n

sections and examine these under the petrological microscope or via other imaging
ur

techniques, for example those that are commonly available on a scanning electron
Jo

microscope (SEM). Petrographic study of the samples serves multiple purposes, including: (i)

assessing the degree of alteration (it is more difficult to accurately estimate the original

(high–T) concentrations of volatile phases in moderately- to heavily-altered samples, and

minerals present at or near the metamorphic peak may have been replaced); (ii) determining

whether one or more minerals shows significant compositional variation (zoning), which

leads to complexities in estimating an appropriate bulk composition for phase equilibrium

modelling and in constraining mineral compositions at the metamorphic peak; (iii)

categorising any inclusions within high-T minerals, identify the peak metamorphic
Journal Pre-proof

assemblage and document any high-T reaction microstructures, which may constrain the

prograde, peak and high-T retrograde path, respectively; and, (iv) identifying dateable

accessory or rock-forming minerals and their microstructural context that, along with analysis

of minerals separates, might provide temporal constraints on the metamorphic evolution of

the sample.

4.5.2. Bulk composition, including ferric iron and H2O

Once those samples to be modelled have been selected, their bulk major oxide

f
oo
compositions need to be determined. If the intention is to quantitatively constrain P–T (and

other) conditions, the modelled bulk compositions should be in the largest system available
pr
(currently MnO–Na2O–CaO–K2O–FeO– MgO–Al2O3–SiO2–H2O–TiO2–O
e-
(MnNCFMASHTO) for pelite/greywacke/granite compositions, and including Cr2O3 for
Pr

(ultra)basic compositions) such that it most closely approximates natural rock compositions.

Bulk compositions are most commonly measured using X-ray fluorescence (XRF),
al

but may also be determined by dissolution inductively-coupled mass spectrometry (ICP–MS).


n

In such cases, the (normally sawn) portion of the sampled rock to be analysed should be free
ur

of weathered portions and in most cases as compositionally homogeneous as possible.


Jo

Increasingly, metamorphic P–T(–t) studies use the composition of the constituent minerals

(measured by EPMA or through energy-dispersive spectroscopy (EDS) on an SEM) and their

modal abundance (estimated using image analysis or SEM/EPMA mapping, including

automated mineral liberation analysis) within polished thin sections to calculate a

composition (Clarke et al., 2001; Godet et al., 2020a, b; Johnson et al., 2004; Lanari et al.,

2013; Lasalle and Indares, 2014; Mitchell et al., 2019; Tinkham and Ghent, 2005b).

However, where minerals are zoned, it may be necessary to ‘remove’ core compositions from

the bulk composition to be modelled to account for those portions of the rock considered not
Journal Pre-proof

to be in communication (equilibrium) with the matrix (Evans, 2004; Marmo et al., 2002;

Tinkham and Ghent, 2005a; Vance and Mahar, 1998; Zuluaga et al., 2005).

Regardless of the method used, uncertainties on the measured concentrations of most

major oxides are generally small (< or << 10% relative), and probably within the

uncertainties associated with phase equilibrium modelling (see below). However, the most

commonly used methods for estimating bulk composition are unable to distinguish the

valency of relevant elements, of which, for our purposes, Fe is the most important (Boger et

al., 2012; Diener and Powell, 2010). Iron exists in many phases in its relatively reduced

f
oo
divalent form (Fe2+) and/or its more oxidised trivalent form (Fe3+), which form ferrous oxide

(FeO) and ferric oxide (Fe2O3), respectively. For example, pure magnetite (Fe3O4) comprises
pr
one mole each of FeO and Fe2O3, whereas pure hematite, which is more oxidised, contains
e-
only Fe2O3. The proportions of ferrous and ferric iron can be measured directly through a
Pr

redox titration of a sample powder, but these measurements are rather imprecise and involve

working with aggressive chemicals. In any case, relative to the redox conditions under which
al

they formed, high-T rocks can only become more oxidised on their return to, and residence
n

on, the surface, and subsequently during sample preparation. As such, Xferric values (Fe3+/Fe)
ur

measured using titrations should be regarded as maxima. In most high-T rocks, values for
Jo

Xferric between 0.1 (commonly used for Archean rocks; Berry et al., 2008) and 0.2 (commonly

used for Phanerozoic rocks) give reasonable results, although T–Xferric (or P–Xferric) diagrams

can be used where P (or T) and H2O contents are constrained (Boger et al., 2012; Diener and

Powell, 2010; Korhonen et al., 2012).

Potentially a more significant problem lies in constraining the abundance of H2O in

modelled bulk rock compositions, as it exerts a first-order control on melt compositions and

fertility, as discussed in section 5. The H2O contents are most commonly estimated based of

the loss on ignition (LOI), measured as part of a standard XRF analysis. However, LOI
Journal Pre-proof

values do not discriminate among the numerous volatile species that might be present (e.g.

H2O, CO2, CH4, N2, F, Cl, S, etc.), and consequently provide only a maximum constraint for

the H2O content of the analyte. Any alteration, which can be very fine-grained, for example

within feldspars, will also increase the H2O content of the rock compared to its original high-

T state. Alternatively, H2O contents can be estimated based on the measured abundance of

hydrous minerals inferred to have been present at the metamorphic peak in combination with

their measured or estimated OH contents. Similar to Xferric, H2O contents can be estimated

using T–XH2O pseudosections if other variables are reasonably well constrained (Johnson et

f
oo
al., 2003b; Korhonen et al., 2012; White et al., 2001). From a more theoretical standpoint, if

the purpose is to model partial melting of protolith composition that is minimally saturated in
pr
H2O as it crossed the solidus at a particular pressure, the H2O content can be adjusted
e-
manually such that < or <<1% H2O-saturated melt is produced at the solidus (Johnson and
Pr

Brown, 2004; Mayne et al., 2020b; White et al., 2001), which is the case for many of the

pseudosections that follow. Alternatively, in situations where H2O is considered to be present


al

as a free phase, its abundance can effectively be set to be infinite.


n

Finally, some consideration should be given to minor components that cannot


ur

(currently) be modelled. This can be done either using the measured or estimated mode and
Jo

composition of the minerals in which these components occur, or using the abundance of the

minor element. For example, in many crustal rocks, phosphorous, which currently cannot be

modelled as a component, is mostly concentrated in apatite (Ca5(PO4)3(OH, F, Cl)). In such

cases, as each mole of P2O5 in (end member) apatite is associated with 3.33 moles of CaO,

this amount can be removed from the bulk composition before any phase equilibrium

calculations (Johnson et al., 2003a). However, in many high-grade metapelitic rocks, most

phosphorous resides in monazite, and P2O5 can simply be ignored. Although MnO can be

modelled in metapelitic systems (White et al., 2014b), it is commonly ignored when


Journal Pre-proof

modelling high-T phase equilibria as Mn only plays a significant role in expanding the

stability field of garnet in greenschist facies rocks. If analysed, the concentrations of sulphur

can be used to subtract an appropriate quantity of Fe associated with any pyrite/pyrrhotite.

4.5.3. Uncertainties and realities

Searching the titles, abstracts and keywords of peer-reviewed articles reveals almost

1000 papers in the petrological literature that include the word ‘pseudosection’, with an

average of around one per year in the 1970s and ‘80s, growing to more than 50 per year since

f
oo
2011. Many such papers propose P–T constraints with a degree of accuracy and precision that

is both unreasonable and unjustified. Like all modelling, phase equilibrium calculations are
pr
associated with uncertainties, some of which are quantifiable and some of which are not
e-
(Palin et al., 2016a). These uncertainties can be subdivided into ‘thermodynamic’,
Pr

‘analytical’ and ‘interpretive’, some of which are intertwined.

For example, it is possible to calculate uncertainties on the end-member


al

thermodynamic data that can be propagated through calculations (Powell and Holland, 2008),
n

but less easy (or impossible) to calculate or estimate uncertainties on the a–X models.
ur

Analytical errors include those associated with measuring or estimating the bulk rock
Jo

composition to be modelled, and the composition and abundance of minerals, including

estimating Xferric in relevant minerals (Droop, 1987; Schumacher, 1991). In relation to

compositions of rocks and minerals, there are inevitable differences between natural systems

and modelled systems related to the omission of elements that cannot (yet) be included in the

modelling, particularly, for example, in relation to crustal melting, the halogens in hydrous

minerals (Bartoli, 2021; Johnson et al., 2008). Interpretive uncertainties might involve

petrographic observations, including the incorrect identification of the peak assemblage or

misinterpretation of reactions between minerals.


Journal Pre-proof

A more thorough treatment of the uncertainties involved in the method are provided

elsewhere (Palin et al., 2016a; Powell and Holland, 2008). Regardless of what is output by

the various computer programs used in calculations, for example a stability field for the peak

assemblage that is a few degrees and/or a few hundredths of a GPa in extent, in reality the 2

uncertainties of ± 50 °C and ± 0.1 GPa estimated (heuristically) for conventional

thermobarometry (Powell and Holland, 2008) are probably sensible minima for most P–T

constraints derived from pseudosections.

f
oo
5. Open vs closed systems
pr
e-
5.1 Definitions
Pr

Definitions of isolated, closed and open systems are provided in section 3.1. Here, we
al

introduce two end-member variants on the terminology for open systems: conditionally open,
n

which indicates that the system becomes episodically open to some infiltration or exfiltration
ur

of fluid or melt (Handy et al., 2001b), but is otherwise closed; and, unconditionally open,
Jo

where the system is always open. Whether natural anatectic systems are open or closed is

partly a matter of scale. At a kilometre scale, suprasolidus crust is generally considered to be

unconditionally open, because it is a source for granites emplaced at higher levels (Brown et

al., 2016; Handy et al., 2001b; Sawyer et al., 2011), even as it accumulates melt from deeper

in the suprasolidus crust (Brown, 2004; Carvalho et al., 2016; Diener and Fagereng, 2014;

Hall and Kisters, 2012; Marchildon and Brown, 2003; Morfin et al., 2013; Nyman et al.,

1995; Schwindinger and Weinberg, 2017; Yakymchuk et al., 2013). However, at the scale of

a thin section or hand sample, a system may remain closed (Yakymchuk, 2021). Outcrops

may comprise layered sequences of different compositions, some of which remained closed
Journal Pre-proof

systems (e.g. refractory metabasite boudins) whereas others were open (e.g. melt-rich

metapelites). As a consequence, phase equilibrium modelling of open system behaviour is

similarly scale-dependent, with implications for the methodology and applicability of the

results.

In studies of crustal melting, the treatment of open-system behaviour is generally

restricted to the gain or loss of fluid (H2O and, less commonly, CO2) and melt. However,

natural systems are inevitably more complex. For example, as melt migrates, batches may

blend with each other and/or entrain residual or peritectic minerals from the source (Clemens

f
oo
et al., 2011; Stevens et al., 2007) or xenocrysts and xenoliths from the host rocks, and

migrating aqueous fluids contain dissolved cations and minor quantities of other volatile

species.
pr
e-
Pr

5.2. Internally vs externally buffered H2O


al

The activity of H2O in a metamorphic system may be internally or externally


n

buffered. Closed systems, where the proportion of any fluid is generally low, are internally
ur

buffered. For suprasolidus rocks, the activity of H2O decreases as temperature increases, and
Jo

melting reactions at temperatures well above the solidus do not include a free fluid phase (i.e.

the system is H2O-undersaturated, with aH2O < 1). Such reactions commonly generate

anhydrous peritectic minerals through hydrate-breakdown (section 2.3), and the volume of

melt produced is limited by the volume of the hydrous reactant mineral present (Sawyer,

2010; Slagstad et al., 2005).

The ingress of an externally-derived fluid into a system above its wet solidus will

cause fluid-present (or fluid-fluxed) melting. In the case of H2O, providing the quantity of

fluid supplied is sufficiently large, the melt may become H2O saturated (aH2O = 1) and remain
Journal Pre-proof

so as long as the supply of H2O is maintained. As fluid is being supplied to the system, none

is required from hydrate breakdown, and H2O-fluxed melting proceeds by consuming

feldspars (plagioclase and/or alkali feldspar) and quartz by the generalised congruent

reaction:

Qtz + Fsp + H2O = L. (17)

The volume of melt produced by H2O-fluxed melting is limited by the supply of H2O or,

f
oo
where fluid:rock ratios are high, the abundance of one or more of the solid reactants.

These principles are well illustrated by migmatites developed in the contact aureole of
pr
the Ballachulish igneous complex in the west-central Scottish Highlands (Pattison and Harte,
e-
1988, 1997). The metasedimentary rocks in the lower-temperature western sector of the
Pr

contact zone of the aureole (i.e. above the 'melt-in' isograd) produced the largest volume of

melt, which led locally to the formation of schollen diatexite migmatites in the ‘Chaotic
al

Zone’, where the greatest fluid-fluxing occurred. By contrast, in the higher-temperature


n

eastern sector of the contact zone, fluid-absent partial melting led to the formation of net-
ur

structured metatexite migmatites, reflecting the lower volume of melt produced in these
Jo

rocks. In the case of fluid-present melting, heat is mainly consumed by melting rather than

increasing the temperature, whereas in the absence of a free fluid, heat is used to increase the

temperature until a major fluid-absent melting reaction is encountered.

A thorough review of the evidence for and role of fluid-present melting has been

provided by Weinberg and Hasalová (2015). Although fluid-present melting reactions

generally are congruent (Johannes, 1985; Stevens and Clemens, 1993), some may be

incongruent, producing peritectic minerals including hornblende (Slagstad et al., 2005),

garnet (Patiño Douce and Harris, 1998) and cordierite (Yardley and Barber, 1991). In
Journal Pre-proof

general, a role for fluid-present melting may be indicated if in situ melt fractions inferred

from the volume of leucosome in a migmatite complex are higher than can reasonably be

accounted for through fluid-absent melting at peak metamorphic conditions (Brown, 2013;

Pourteau et al., 2020), although care is required since melt may have infiltrated from deeper

levels in the crust. The trace element composition of granites can also be used to infer the

behaviour of plagioclase and micas during partial melting to estimate if H2O was a reactant

and if fluid-present melting occurred (Gao et al., 2017a; Inger and Harris, 1993).

The relative importance of fluid-present and fluid-absent partial melting in the

f
oo
generation of granites is debated (Clemens et al., 2016, 2020; Weinberg and Hasalová, 2015).

Fluid-fluxed melting has been argued to be important for the growth of continental crust at
pr
subduction zones (Collins et al., 2020b), but in orogenic belts the limited permeability of the
e-
deep crust at steady-state conditions (Yardley and Valley, 1997) may prohibit pervasive fluid
Pr

movement and restrict widespread fluid-present partial melting. However, shear zones that

develop during deformation may provide permeable conduits to allow ingress of hydrous
al

fluids and induce significant melting (Carvalho et al., 2016; Johnson et al., 2001c; Mogk,
n

1992; Sawyer, 2010; Weinberg and Hasalová, 2015). Although the efficacy and extent of
ur

fluid-present melting in the continental crust is debated, there is commonly a genetic link
Jo

between migmatites that formed through fluid-present melting and syn-anatectic deformation.

One of the biggest challenges in the petrogenesis of migmatites generated through

fluid-present partial melting is determining the source of H2O. To do so usually requires a

detailed understanding of the tectonic, metamorphic and magmatic history of a terrane

coupled with geochronology and stable isotope analysis of both the migmatites and the

inferred fluid sources. Fluids can be locally sourced or be derived far from the site of

anatexis. For example, H2O may readily be derived from sedimentary rocks beneath large

intrusions, such as the Bushveld Complex, in which dehydration reactions at depth may flux
Journal Pre-proof

melting of rocks closer to the basal contact (Johnson et al., 2003b). In regional metamorphic

terranes, two commonly inferred sources of H2O are: (1) underthrust or underplated

sedimentary rocks that undergo dehydration reactions without melting (Berger et al., 2008;

White et al., 2005); and, (2) fluids exsolved from the crystallization of hydrous granite

(Finger and Clemens, 1995) or basalt magmas (Annen and Sparks, 2002; Collins et al., 2016,

2020b; Pourteau et al., 2020). Diffusion of H2O from a magmatic source into rocks at

temperatures close to the wet solidus is a viable mechanism to produce granite melts, but it is

unlikely to generate a significant volume of magma (Tafur and Diener, 2020). Although fluid

f
oo
sources may be cryptic, oxygen isotope analysis can be used to differentiate alternative

sources, whether using whole-rock samples (Holk and Taylor, 2000; Johnson et al., 2003b;
pr
Sawyer, 2020; Taylor Jr, 1978; Yakymchuk et al., 2019b) or in situ analysis of minerals
e-
(Gordon et al., 2009; Higashino et al., 2019; Smithies et al., 2021).
Pr

5.3. Phase equilibrium modelling approaches to open system melting


n al

A first-order consequence of melt drainage is an increase in the density and strength


ur

of the residual crust (Diener and Fagereng, 2014; Yakymchuk and Brown, 2014b). Although
Jo

the effect of continuous melt loss on crustal rheology has not been quantified, accumulation

of melt up to the MCT (or a lower threshold, as discussed in section 2.4) allows episodic melt

drainage, leading to transient periods of strengthening and weakening in the source (Brown,

2010a; Diener and Fagereng, 2014). Melt loss generally occurs during prograde heating, but

it may also occur during cooling and melt crystallization, in particular where there is

synchronous deformation (Koblinger and Pattison, 2017). Evidence of melt loss includes: (1)

the preservation of peak metamorphic assemblages (Brown, 2002; White and Powell, 2002b);

(2) the residual (melt-deficient) composition of many migmatites and granulites compared to
Journal Pre-proof

their inferred protoliths (Johnson et al., 2013; Korhonen et al., 2010a); and, (3) the presence

of structures consistent with a decrease in volume (Bons et al., 2008; Diener and Fagereng,

2014). Melt may migrate from centimetres to metres, to accumulate in local melt-induced

deformation-band networks, or from metres to kilometres, to accumulate in crustal-scale

plutons, as discussed in Section 2.4.

Melt loss changes the composition of an anatectic system, which leads to a number of

effects, including: (1) an increase in the temperature of the solidus due to a decrease in bulk

H2O contents and aH2O (Korhonen et al., 2010a; Schwindinger et al., 2019; Stuck and Diener,

f
oo
2018); (2) shifts in the topology of a phase diagram due to preferential loss of those elements

concentrated in the extracted melt (notably H2O, K, Na, and Si; (Mayne et al., 2020a); (3) an
pr
increase in the stability of anhydrous minerals due to the extraction of relatively H2O-rich
e-
melt (Johnson et al., 2011; Yakymchuk and Brown, 2014a), and the preservation of high-
Pr

temperature mineral assemblages (Hernández-Montenegro et al., 2019; White and Powell,

2002a; White and Powell, 2010); (4) reduced fertility of the residue with increasing melt
al

extraction (Yakymchuk and Brown, 2014b); (5) depletion of the source in the relatively
n

incompatible elements that preferentially enter the melt (Mayne et al., 2020b); (6) an increase
ur

or decrease in the stability of accessory minerals depending on the concentration of essential


Jo

structural elements in the extracted melt (Rapp et al., 1987; Yakymchuk and Brown, 2014b);

and, (7) modified trace element compositions of the source (Kendrick and Yakymchuk, 2020)

and of accessory minerals (Holder et al., 2020). Of these effects, the preservation of high-

temperature mineral assemblages and an increase in the solidus temperature are the most

relevant to studies of natural migmatites and granulites. In particular, the solidus temperature

is important for interpreting geochronological data, as it strongly influences when accessory

minerals crystallize from melt during cooling (Korhonen et al., 2013b). In this section, we
Journal Pre-proof

consider strategies to address the effects of melt loss on both source and melt compositions

using phase equilibrium modelling.

The effect of melt drainage on the composition, fertility and mineral assemblage of

residual rocks can be investigated by forward modelling of phase equilibria. First, forward

modelling may be used to predict how the mineral assemblages and fertility of a source rock

change with increasing temperature through loss of melt (White and Powell, 2010;

Yakymchuk and Brown, 2014b). Second, forward modelling may be used to solve the inverse

problem where melt is added back into a residual composition to create a fictive but plausible

f
oo
subsolidus protolith, an approach generally called melt reintegration (Bartoli, 2017;

Korhonen et al., 2013a; White et al., 2004).


pr
Forward modelling of melt drainage has been used to consider: (1) a single melt loss
e-
event close to peak T (Indares et al., 2008); (2) multiple melt loss events to reflect a system
Pr

that episodically reaches a specified MCT during heating (Diener and Fagereng, 2014;

Johnson et al., 2011; Schwindinger et al., 2019; Stuck and Diener, 2018; Yakymchuk and
al

Brown, 2014b); and, (3) quasi-continuous melt extraction along a prograde path (Mayne et
n

al., 2016; Yakymchuk and Brown, 2019). The composition of the extracted melt is usually
ur

dictated by the source composition and the P–T conditions of melt drainage, as determined
Jo

using phase equilibrium modelling. However, the amount of melt to remove at this P–T

condition will vary depending on whether the modelled system is expected to be static

(Handy et al., 2001a) or dynamic (Marchildon and Brown, 2001; Stuart et al., 2018). For a

static system in which drainage is episodic, we may assume that melt loss will occur when

the melt fraction reaches the MCT (Rosenberg and Handy, 2005). In this case, it may be

appropriate to retain a portion of melt to account for that trapped along grain boundaries, as

inferred from microstructural studies of anatectic rocks (Holness and Sawyer, 2008;

Marchildon and Brown, 2002; Sawyer, 2001). For example, Yakymchuk and Brown (2014b)
Journal Pre-proof

modelled melt drainage from migmatites by removing 6 mol.% melt from the system when a

7 mol.% extraction threshold was reached along a P–T path, leaving 1 mol.% melt in the

system to represent that expected to have been trapped along grain boundaries. Whether melt

extraction was episodic or continuous does not appear to significantly impact the predicted

mineral assemblages and modes in the residue or the composition of the melt (Mayne et al.,

2020b).

To illustrate the effects of melt drainage in migmatites, Yakymchuk and Brown

(2014b) compared the fertility of metapelites and greywackes. They showed that an average

f
oo
amphibolite-facies metapelite generates ~36 vol.% melt along an isobaric heating path to 890

ºC at 1.2 GPa in a closed system, but in a conditionally open system undergoing four melt
pr
drainage events generates only ~30 vol.% along the equivalent path due to the loss of H2O
e-
and other incompatible components in melt that drained from the system (Fig. 10a). In this
Pr

case the residue does not generate additional melt during high-T decompression due to its

H2O-depleted composition, and in some cases the trapped melt may begin to crystallize and
al

may cross the solidus. Thus, an important consequence of episodic melt extraction is that the
n

efficacy of decompression melting is significantly decreased (Yakymchuk and Brown,


ur

2014b), calling into question tectonic models that rely on decompression melting in the
Jo

genesis of gneiss domes (Teyssier and Whitney, 2002). By contrast, the fertility of

greywackes is less sensitive to melt loss (Fig. 10b). For a metagreywacke composition

following the same P–T path as the metapelite, an undrained system yields 21 vol.% melt,

whereas a drained system produces ~16 vol.% melt (Yakymchuk and Brown (2014b).

Similar to other rock types, metabasites will generate less melt in a drained system

when compared with an undrained system (Stuck and Diener, 2018). However, like

metapelites and metagreywackes (Johnson et al., 2011; Yakymchuk and Brown, 2014b)),

although the modelled residual mineral assemblage does not change significantly with
Journal Pre-proof

episodic melt drainage, both reactants and products may be stabilised to higher temperatures

when compared with an undrained system (Palin et al., 2016c). Major and trace element

compositions of melt are also expected to be slightly different between drained and undrained

systems (Huang et al., 2020; Kendrick and Yakymchuk, 2020).

Reintegration of melt into a residual composition requires assumptions about the

number of steps to use and the total amount of melt to reintroduce into the system, the P–T

conditions at which each step should occur, and the composition of the melt at each step. One

approach is to use the amount of melt required to shift an elevated solidus down temperature

f
oo
to a predetermined P–T condition using the composition of melt predicted by phase

equilibrium modelling (Diener et al., 2008; White et al., 2004). Alternative approaches use
pr
either the composition of a melt produced in an experiment at a known P–T condition or the
e-
composition of nanogranite/melt inclusions (Bartoli, 2017, 2019). Melt may be reintegrated
Pr

in a single step (Guilmette et al., 2011; Indares et al., 2008), in multiple steps (Alessio et al.,

2017; Korhonen et al., 2013a; Yakymchuk et al., 2015), or using T–Mmelt diagrams (Anderson
al

et al., 2013). These different approaches generally produce diagrams with similar topologies,
n

although the fertility of the reintegrated composition may vary (Bartoli, 2017).
ur

The largest uncertainty in modelling melt reintegration is the assumed prograde P–T
Jo

path, which affects the minerals present at the P–T condition of melt loss and the composition

of the melt. Although reintegration of melt into a residual migmatite or granulite offers

insight into the original composition of the protoliths in circumstances where subsolidus

equivalents are not exposed, considering the uncertainties involved, the results of this

procedure should be treated with caution.

6. Melting of metaigneous rocks


Journal Pre-proof

Igneous rocks and their deformed and metamorphosed equivalents (orthogneisses)

comprise by far the largest proportion of the crust, which is bimodal in terms of its chemical

composition (Bonnefoi et al., 1995; Collins et al., 2020b; Daly, 1925; Keller et al., 2015) and,

at least since the end of the Archean, its hypsometry (elevation) (Flament et al., 2008;

Schubert and Reymer, 1985; Wise, 1974). On the modern Earth, this bimodality is clearly

manifest as oceanic crust that is dominantly basaltic in composition and subaqueous, and

continental crust that is on average andesitic in composition and mostly subaerial. However,

continental crust also shows pronounced compositional bimodality, with peaks at ~50 wt%

f
oo
SiO2 and ~70–75 wt% SiO2, in particular for the intrusive rock record (Keller et al., 2015).

The most abundant magmatic rock type within the middle to upper continental crust is
pr
granite, with mafic rocks and/or residual granulites dominating the lower crust (Hans
e-
Wedepohl, 1995; Rudnick and Gao, 2003; Rudnick and Fountain, 1995).
Pr

Although a detailed treatment of crustal magmatism per se is beyond the scope of this

contribution, in this section, we discuss the most important (by volume) crustal magmatic
al

protoliths, placing a particular emphasis on the basic (basaltic/gabbroic) and felsic


n

(rhyolitic/granitic sensu lato) end members. In terms of phase equilibria, we focus on the
ur

anatexis of metabasic rocks, for which there are far more experimental constraints (see
Jo

section 4.1.2).

6.1. Magmatic protoliths

Mafic crust is formed through melting of the mantle driven by decompression (for

example at mid-ocean ridges and in plumes), advection of heat (e.g. in plumes) or the

introduction of H2O (for example, slab-derived fluids beneath island arcs). Regardless of the

mechanism, these primitive (juvenile) magmatic rocks represent the source from which
Journal Pre-proof

almost all continental crust was ultimately derived (Albarède, 1998; Hawkesworth and

Kemp, 2006; Taylor and McLennan, 1985). However, the processes by which continental

crust forms from mafic crust, whether through partial melting or fractional crystallisation,

and how these processes may have changed through time are topics of longstanding and

ongoing debate (Annen et al., 2006; Arndt, 2013; Collins et al., 2020b; Davidson et al., 2007;

Jacob et al., 2021; Jagoutz and Klein, 2018; Johnson et al., 2017; Moyen et al., 2021).

However, there is agreement that to partially melt mafic crust at temperatures that may

plausibly be attained during orogenesis (up to ~1000 °C), mafic rocks must be hydrated,

f
oo
requiring a source of H2O in the generation of continental crust (Campbell and Taylor, 1983).

Of particular importance to understanding the formation of the Archean continental


pr
crust is partial melting of hydrated mafic rocks to generate the tonalite–trondhjemite–
e-
granodiorite (TTG) suite, and on modern Earth compositionally-similar high-silica adakites
Pr

(Barker and Arth, 1976; Castillo, 2006; Condie, 2014; Martin et al., 2005; Moyen, 2011;

Moyen and Martin, 2012; Rollinson and Martin, 2005; Smithies, 2000). TTGs are sodic
al

granites with high Sr/Y and La/Yb ratios, and they are the volumetrically-dominant rock-type
n

within the deeper levels of exposed Archean continental crust (Condie, 1981; Martin, 1994;
ur

Windley, 1996). These TTGs represent the most primitive (least evolved, with K2O/Na2O
Jo

ratios < or << 0.6) continental crust (Smithies et al., 2021). Although some TTG suites may

be the result of fractional crystallisation of basaltic magmas (Kramers, 1988; Liou and Guo,

2019), there is general agreement that most were derived by partial melting of amphibolite

within the stability field of garnet, corresponding to source depths of 25–50 km (Johnson et

al., 2017; Smithies et al., 2019). Whereas some propose TTGs represent supra-subduction-

zone magmas (Foley, 2008; Foley et al., 2003; Kusky and Polat, 1999; Polat et al., 2014;

Windley et al., 2020) others argue for their formation near the base of thick mafic plateau-

like mafic crust in a stagnant-lid setting that may have characterised the early Earth (Bédard,
Journal Pre-proof

2006a; Bédard, 2018; Johnson et al., 2017; Smithies et al., 2009, 2019). In contrast to TTGs,

and although rare, high-silica adakites occur in the modern rock record. Most form at plate

margins and are interpreted to indicate high-pressure melting of down-going oceanic

lithosphere associated with ‘hot’ (young) subduction zones, in which the ascending slab-

melts variably interact with the mantle wedge (Martin et al., 2005; Smithies, 2000).

The most evolved volumetrically-abundant igneous rocks within the continental crust

are granites sensu stricto. These are high silica, high-K rocks that dominate active and

ancient continental arcs, and which span the entire geological rock record, first becoming

f
oo
abundant towards the end of the Archean (Chowdhury et al., 2020; Johnson et al., 2019;

Moyen, 2020a; Moyen et al., 2003; Nebel et al., 2018; Zhao et al., 2017). High-K granites
pr
can form by a range of processes, including partial melting of common metasedimentary
e-
rocks (see section 7), low-degree melting of a wide range of igneous protoliths and fractional
Pr

crystallisation of mafic and other magmas. Fractionation clearly has a role to play in the

petrogenesis of high-K granites, which ultimately drives the formation of pegmatites and
al

hydrothermal fluids that are the source of numerous rare and more common metals and other
n

ore deposits on which humankind depends.


ur

In terms of partial melting of different source lithologies, and based originally on data
Jo

from granites in the Lachlan Fold Belt, southeastern Australia (Chappell, 1974), a key

compositional subdivision of granite is into I-, S- and A-types (Chappell, 1999; Chappell and

White, 1992; Clemens et al., 1986; Clemens, 2003; Collins, 1996; Collins et al., 2020a; Eby,

1992; Foden et al., 2015; Frost et al., 2001). I-type granites, which form during collision and

in the lower crust of some arcs, are mostly calcic and sodic metaluminous (ASI < 1) to

weakly-peraluminous (ASI > 1) rocks derived from predominantly mafic igneous (hence the

‘I’) precursors, although elevated zircon δO18 suggests that many incorporate a supracrustal

component (Kemp et al., 2007). In addition, the relative role of partial melting versus
Journal Pre-proof

fractionation is debated (Collins et al., 2020b; Jagoutz and Schmidt, 2012; Keller et al.,

2015). By contrast, syn- to post-collisional S-type granites are high-K peraluminous rocks

mostly derived by anatexis of sedimentary (‘S’) protoliths (Collins and Richards, 2008;

White and Chappell, 1988), although Sr and Nd isotope data indicate a mantle-derived

component is also commonly involved (Healy et al., 2004). Post-collisional or anorogenic A-

type granites, which are commonly located in extension (rift) settings, are hot, dry magmas

that mostly reflect relatively low-P partial melting of residual (melt-depleted) source rocks of

varying composition that underwent an earlier metamorphic cycle (Clemens et al., 1986;

f
oo
Collins et al., 1982; Eby, 1992; Patiño Douce, 1997; Turner and Rushmer, 2009; Whalen et

al., 1987).
pr
e-
6.2. Field and petrographic characteristics of migmatised metaigneous rocks
Pr

6.2.1. Metabasic rocks


al

Relative to more fertile protoliths, such as pelites and greywackes (section 7), there
n

have been relatively few detailed field-based studies on partially-melted basic rocks (i.e.,
ur

metabasic migmatites). On a regional scale, areas where such metamorphic studies have been
Jo

undertaken include the Grenville front in Quebec (Sawyer, 1991) and the Kapuskasing Zone

in Ontario (Hartel and Pattison, 1996), Canada, the Lewisian Complex of NW Scotland

(Johnson et al., 2012), the Ivrea Zone in northern Italy (Kunz et al., 2014), and the

Sveconorwegian orogen of SW Sweden (Hansen et al., 2015). Many of the features described

below are nicely illustrated in these (and other) publications, in particular in the Atlas of

Migmatites (Sawyer (2008).

Hydrated metabasic rocks generally comprise assemblages dominated by hornblende

and plagioclase, commonly with quartz, as they begin to partially melt. Depending on bulk
Journal Pre-proof

composition and the metamorphic pressure, they may also contain variable quantities of

garnet, clinopyroxene, biotite, epidote, titanite, magnetite, ilmenite and rutile, and less

commonly gedrite, muscovite and/or K-feldspar. Although much less abundant than in

migmatites with pelitic or greywacke protoliths, metabasic migmatites may contain zircon or,

if silica-undersaturated, baddeleyite, allowing them to be dated (Fischer et al., 2021).

When the inferred degree of partial melting is small, leucosomes may form irregular

isolated pockets or discontinuous stromatic veins, the latter of which may mimic original

magmatic and/or deformation-induced layering (Fig. 11a, b). As temperature (and the

f
oo
inferred degree of melting) increases, leucosomes are seen to coalesce into larger-scale

structures, commonly sheets or more irregular accumulations facilitating drainage of melt


pr
from the system (Fig. 11c, d). The leucosome in metabasic migmatites consists mostly of
e-
plagioclase and quartz with minor K-feldspar and may contain variable quantities of
Pr

entrained material. Leucosome may also contain peritectic minerals, including garnet,

clinopyroxene and/or orthopyroxene, although these are commonly altered to some degree,
al

and in some cases hydrous peritectic minerals, including hornblende, biotite and epidote
n

(Yakymchuk et al., 2019b), consistent with melting via incongruent reactions. However,
ur

peritectic minerals are commonly absent, suggesting H2O-fluxed congruent melting may play
Jo

an important role in the anatexis of many metabasic rocks (Collins et al., 2020b; Nehring,

2012; Pourteau et al., 2020; Yakymchuk et al., 2019b). As peak temperature increases

further, the rocks generally become coarser-grained and the proportions of hornblende,

plagioclase and quartz decrease, whereas those of garnet, clinopyroxene and/or

orthopyroxene increase.

Collectively, the field, petrographic and experimental data (section 4.1) indicate that

the following two generalised reactions are of fundamental importance to the partial melting

of hydrous mafic rocks at crustal depths:


Journal Pre-proof

Hbl + Qtz ± Bt ± Pl → Cpx ± Grt ± Opx + L, and (18)

Pl + Qtz + H2O ± Kfs → L. (19)

(Hartel and Pattison, 1996; Sen and Dunn, 1994). In the case of fluid-absent melting, a first

order control on the amount of melt that can be generated by reaction 18 is the abundance of

hornblende in the protolith. In terms of fluid-present melting, given a supply of H2O, the

quantity of melt that can be produced is limited by the abundance of quartz and feldspar in

f
oo
the protolith, which in primitive mafic compositions may be low.

6.2.2. Metamorphosed felsic igneous rocks


pr
e-
Similar to metabasic rocks, there are comparatively few studies documenting the
Pr

features of partially-melted felsic igneous rocks. Exceptions include the Central Gneiss

Complex in Alaska, USA (Kenah and Hollister, 1983; Lappin and Hollister, 1980; McLellan,
al

1988), the Opatica Subprovince in the Canadian Shield (Sawyer, 2010), the Grenville
n

Province in Ontario, Canada (Slagstad et al., 2005), the Gallatin Range of Montana, USA
ur

(Mogk, 1992), the Central Alps (Berger et al., 2008), the Bohemian Massif in the Czech
Jo

Republic (Hasalová et al., 2008a, b; Schulmann et al., 2008; Štípská et al., 2019), the

Lewisian Gneiss Complex of NW Scotland (Johnson et al., 2013), the Campo Belo

metamorphic complex in the São Francisco Craton, Brazil (Carvalho et al., 2016) and the

Taishan region of the Western Shandong Province, China (Yakymchuk et al., 2019b).

Notably, unlike other common crustal rock types that have undergone anatexis, the colour

contrast between leucosome and residuum in felsic igneous rocks, which are both

quartzofeldspathic, may be subtle, and clear migmatitic features indicative of partial melting

become much harder to recognize (Johnson et al., 2013; Sawyer, 2010).


Journal Pre-proof

Partially-melted felsic orthogneisses exhibit a full range of migmatitic structures,

from stromatic, patch and net-structured metatexites to nebulites and schollen diatexites

(Sawyer, 2008, 2010; Slagstad et al., 2005) (Fig. 11e, f). However, a key observation is that

the volumes of melt produced by fluid-absent melting, as inferred from the presence of

anhydrous peritectic minerals within leucosome, are small, even where peak metamorphic

temperatures were extreme (950–1000 °C) (Johnson et al., 2013); Fig. 11e,f). For example,

reaction 18 also describes fluid-absent melting in calcic intermediate and felsic ‘I-type’

igneous rocks, such as tonalites and trondhjemites (Johnson et al., 2017), but is limited by the

f
oo
abundance of hornblende (with or without biotite), which is much lower than in metabasic

rocks. Fluid-absent melting of peraluminous ‘S-type’ granites proceeds mostly by reactions


pr
consuming biotite and/or muscovite, and mimics reactions that govern suprasolidus phase
e-
equilibria in metapelites and aluminous metagreywackes, as described in section 7. In
Pr

partially-melted felsic orthogneisses, larger melt volumes, including diatexites, only occur

where melting (for example, by reaction 19) was fluxed by ingress of an H2O-rich fluid
al

(Carvalho et al., 2016; Clos et al., 2019; Sawyer, 2010) due to the high modal abundance of
n

quartz and feldspars.


ur
Jo

6.3. Calculated phase equilibria

Fully-quantitative calculations investigating partial melting of hydrated mafic rocks

(amphibolites) and other Ca-rich, K-deficient intermediate to felsic protoliths (diorites,

tonalites, etc.) have only been possible since publication of the a–X models of Green et al.

(2016). Importantly, in addition to a new solution model for (trondhjemitic–tonalitic) melt,

this work included models for high-T clinopyroxene and hornblende. Of particular

importance, the hornblende model includes K, such that both biotite-present and biotite-
Journal Pre-proof

absent equilibria may be considered. Some key papers utilising these models include those by

Palin et al. (2016c), Palin et al. (2016b), Johnson et al. (2017), White et al. (2017), Feisel et

al. (2018), Kunz and White (2019), Hernández-Uribe and Palin (2019), Huang et al. (2020)

and Kendrick and Yakymchuk (2020). We illustrate the calculated phase equilibria with

reference to metabasic protoliths. Intermediate to felsic ‘I-type’ compositions show similar

topologies, although hornblende is consumed at lower T and quartz and feldspars at higher T

in these lithologies. The phase equilibria for K-rich peraluminous protoliths (granites) are

similar to those described in the following section.

f
oo
6.3.1. Fluid-absent melting
pr
Given the importance of the role of partial melting of metabasic rocks to the
e-
production of primitive continental crust in the early Earth, we first illustrate phase relations
Pr

with reference to a bulk composition that has been shown to be appropriate to the production

of common TTGs with respect to both major and trace element compositions (Johnson et al.,
al

2017; Smithies et al., 2009). The modelled composition, an average (n = 10) of the c. 3.5 Ga
n

Coucal basalts (hereafter CF2) from the East Pilbara Terrane in Western Australia, is
ur

compositionally evolved, exhibiting a low XMg (0.38) and enrichment in K, LILE, Th and
Jo

LREE relative to primary basalts such as MORB. A P–T pseudosection calculated between

0.35 and 1.35 GPa and from 630 to 1000 ºC for CF2, and containing sufficient H2O to

minimally saturate the solidus at 1.2 GPa, is shown in Fig. 12a, with a simplified version

highlighting the key petrological features shown in Fig. 12b. We describe phase relations

with reference to melting along three paths with thermobaric (T/P) ratios of 700, 900 and

1100 °C/GPa (Fig. 12), which correspond to relatively low, intermediate and high crustal

geotherms. The changing abundance of phases along these three prograde melting paths are

shown in Fig. 13.


Journal Pre-proof

Within the modelled P–T window, at high-T subsolidus conditions the phase

assemblage corresponds to a quartz-amphibolite containing >50 mol.% hornblende. Minor

biotite is present at all pressures, and coexists with minor titanite at higher P and ilmenite at

lower P. At P > 1.2 GPa small quantities of muscovite are predicted to be present and epidote

is predicted at similar pressures in other bulk compositions––these minerals may play a minor

role in fluid-absent melting of metaigneous rocks in the deep crust. The wet solidus occurs at

temperatures between ~650 °C at high P and ~700 °C at low P. Significant melting begins

through fluid-absent reactions that consume biotite, hornblende and quartz, producing ~15%

f
oo
melt as biotite is exhausted, and a total of 25–30% melt as hornblende (± quartz) is exhausted

at 800–870 °C (Figs 12 and 13). Above these temperatures, melting continues by


pr
consumption of anhydrous minerals, decreasing the H2O content of the melt.
e-
At relatively low P (yellow fields in Fig. 12b), neither garnet nor rutile is predicted to
Pr

be stable; at intermediate P (green fields) garnet is stable without rutile; at high P (blue

fields) garnet and rutile coexist. These fields correspond to the low-pressure, medium-
al

pressure and high-pressure TTG subdivisions of (Moyen, 2011). Plagioclase is predicted to


n

be stable throughout the P–T window of interest, consistent with experiments on quartz-rich
ur

amphibolite starting compositions. Importantly, garnet, which defines the lower pressure
Jo

limit for MP-TTGs, is predicted to be stable in CF2 at pressures significantly lower (by up to

0.8 GPa) than suggested by melting experiments using other amphibolite starting materials

(Moyen and Stevens, 2006).

The main mineralogical effects of the paths with different thermobaric ratios are

shown in Fig. 13a. With increasing T/P or along lower pressure isobaric heating paths, the

proportion of garnet decreases significantly and of quartz a little. Rutile is predicted at high-T

along the 700 °C/GPa path, but ilmenite is stable at higher thermobaric ratios. For the

composition of CF2, orthopyroxene is only predicted along the 1100 °C/GPa path. In all
Journal Pre-proof

cases melt fractions show a broadly linear increase with T, although fertility increases

slightly with increasing thermobaric ratio (Fig. 13).

To illustrate the compositional evolution of the melt, we use as an example partial

melting along the 900 °C/GPa path at melt fractions between 5 and 40 mol.% (Fig. 13d). At

low melt fractions (5–10 mol.%), the calculated melts are H2O-rich and granitic with

K2O/Na2O > 0.8. At 20% melting, the calculated K2O/Na2O ratio (0.54) of the melt is similar

to the upper range observed in TTGs, and at 30% melting this ratio (0.38) is very close to the

average measured value in TTGs. In addition, using published partition coefficients and the

f
oo
predicted melt residua (Fig. 13), and the methodology detailed in section 9, trace element

modelling of the calculated melt compositions along the 900 °C/GPa at melt fractions of
pr
around 25 vol.% show good correspondence with the composition of the average composition
e-
of the TTGs from the East Pilbara Terrane (Johnson et al., 2017).
Pr

To illustrate the effects of changing major oxide bulk composition on suprasolidus

phase equilibria in metabasites, in Fig. 14 we show for comparison P–T pseudosections for
al

an average N-MORB composition (Sun and McDonough, 1989) with plagioclase in excess,
n

and an average Eoarchean basalt from Isua, SW Greenland (Hoffmann et al., 2011) with
ur

hornblende in excess, both of which contain sufficient H2O to minimally hydrate the solidus
Jo

at 1.2 GPa. Relative to CF2, the average N-MORB contains less silica, much more CaO, has

a lower K/Na ratio and a significantly higher XMg of 0.64 (Table 1). Although these

compositional differences cannot be treated in isolation, the most notable cumulative effects

on the average N-MORB during heating are that quartz is consumed at lower T, hornblende is

stable to much higher T (>1000 °C) and there is only a very narrow field in which

suprasolidus biotite is stable (compare Fig. 14a with Fig. 12a). In terms of pressure, the

maximum stability of rutile and ilmenite occurs at much lower P and, of particular

significance, the lowest P stability of garnet is at ~0.95 GPa compared to ~0.75 GPa in CF2,
Journal Pre-proof

due mainly to the much higher XMg of N-MORB (Johnson et al., 2017). In addition,

plagioclase is not predicted to be stable in the high P, low T part of the modelled P–T

window.

Relative to CF2, the average Isua basalt is much more primitive and contains less

SiO2, lower TiO2 and Al2O3, less Na2O and K2O (although the K/Na ratio is similar) and

almost three times the MgO, with a much higher XMg (0.67) that is similar to the average N-

MORB (Table 1). Compared to CF2, the principal effects on the average Isua basalt are that

the maximum thermal stability of quartz is reduced even more than for the average N-

f
oo
MORB, hornblende is stable to T > 1000 °C and the lower stabilities of rutile and garnet and

the upper stability of ilmenite occur at higher pressures than for the average N-MORB
pr
(compare Fig. 14b with Figs 12a and 14a). The other notable difference is that the stability of
e-
plagioclase is greatly reduced, and plagioclase occurs only in the low P–T part of the
Pr

diagram, at P < 0.85 GPa and T < 830 °C, such that the H2O-saturated solidus at P > 0.85

GPa occurs at temperatures up to 50 °C higher than in the other two compositions. In terms
al

of fertility, at a given suprasolidus T, CF2 produces more melt than the average N-MORB,
n

whereas the average Isua basalt is even less fertile (Figs 12a, 14a, b). However, at least in
ur

terms of major oxides, the compositions of melts produced from these different protoliths
Jo

varies little, evolving from granite at low melt fractions to trondhjemite–tonalite at higher

melt fractions.

As stated previously, intermediate to felsic metaluminous (‘I-type) granites will also

melt mostly by reactions consuming hornblende (reaction 18). The main difference in the

phase equilibria for these more evolved (silicic) rocks will be a reduction in the stability of

hornblende, an increase in the stability of quartz, plagioclase and biotite, and, in the most

evolved compositions, muscovite and/or K-feldspar. For a more comprehensive treatment of


Journal Pre-proof

the effects of varying major oxide compositions, the reader is referred to the articles by Palin

et al. (2016c) and White et al. (2017).

6.3.2. Fluid-present (H2O-fluxed) melting

Partial melting driven by an influx of H2O-rich fluid is argued to be important in the

generation of some magmatic rocks, as discussed in section 5. In particular, H2O-fluxed

melting of mafic underplate in arcs has recently been argued to be a primary driver in the

production of silicic arc magmas (Collins et al., 2016, 2020b), although the relative

f
oo
importance of fluid-present versus fluid-absent melting and the role of fractional

crystallisation is debated (Clemens et al., 2021; Jagoutz et al., 2011; Jagoutz and Schmidt,

2012; Sisson et al., 2005).


pr
e-
To illustrate the process of modelling H2O-fluxed melting of metabasites, we take as
Pr

an example the composition of a ‘fine-grained, quench-textured hornblende gabbro’ (sample

87S35A) used in the experiments of Sisson et al. (2005) to investigate the potential origin of
al

high-K granites in continental arcs. We model the effects of introducing a pure H2O fluid
n

phase at 0.8 GPa, chosen to approximate the base of thickened continental arc crust (see
ur

(Collins et al., 2020b) for more details). For reference, a P–T pseudosection for the
Jo

composition of sample 87S35A containing sufficient H2O to minimally saturate the solidus at

0.8 GPa is shown in Fig. 15, in which plagioclase and hornblende are in excess. Compared to

the samples modelled so far in this section, this composition is depleted in SiO2, moderately

enriched in K2O, such that suprasolidus K-feldspar is stable at P > 0.7 GPa, and has an

intermediate XMg of 0.56. Garnet is only stable in the high P–T corner of Fig. 15a.

An isobaric (P = 0.8 GPa) T–X pseudosection is shown in Fig. 15b, for which T varies

from 650 to 800 °C (thick dashed white line in Fig. 15a) and X (= H2O content in wt%) varies

from the value used in Fig. 15a (~1 wt.%) to 6 wt%, corresponding to the addition of an extra
Journal Pre-proof

5 wt% H2O. Melt fractions (F) varying from 5 to 50 mol.% are shown as the red dashed lines.

Two arrowed paths are shown on Fig. 15b, the yellow arrow representing introduction of

H2O at 700 °C, ~50 °C above the solidus, and the pink arrow representing influx at 750 °C.

In both cases, for minimally-saturated conditions (yellow and pink stars on Fig. 15a, left hand

edge of the yellow and pink arrows on Fig. 15b) in which all the H2O is locked within

amphibole and biotite, or just amphibole, respectively, trivial amounts of melt (<5 mol.%) are

produced, quantities that may be insufficient for segregation and migration to occur

(Rosenberg and Handy, 2005). However, with the addition of just 1 wt.% H2O (bringing the

f
oo
total to ~2 wt%; yellow and pink diamonds), melt fractions increase to ~12 mol.% and ~15

mol.%, respectively. With the addition of 4 wt% H2O (bringing the total to ~5 wt%; yellow
pr
and pink discs) melt fractions are predicted to be >35 mol.% and >40 mol.%, respectively,
e-
volumes that may be sufficient for bulk flow to occur (Arzi, 1978; Renner et al., 2000;
Pr

Rosenberg and Handy, 2005; Vigneresse et al., 1996). Increasing H2O content still further at

750 °C (pink arrow) produces more melt. However, with 4 wt% added H2O (i.e. 5 wt% H2O
al

in total), at 700 °C the melt is saturated (i.e. the path has crossed the H2O-out curve), such
n

that the excess H2O forms a separate fluid phase that coexists with melt. Whether melt drains
ur

from the system likely depends on whether the local structural setting allows melt
Jo

accumulation or not (see section 5).

In Fig. 15b, following either path towards higher H2O contents (i.e. to the right in the

figure) leads to the exhaustion of first K-feldspar then, following the consumption of

clinopyroxene and growth of titanite at 750 °C, quartz and eventually ilmenite. Plagioclase is

also consumed, although it is not exhausted at the maximum H2O contents shown, consistent

with fluid-fluxed melting by reaction 19. While generally underappreciated, the importance

of H2O-fluxed melting to the generation of both modern and ancient continental crust is

gaining increasing traction (Collins et al., 2020b; Nehring, 2012; Pourteau et al., 2020;
Journal Pre-proof

Weinberg and Hasalová, 2015; Yakymchuk et al., 2019b) and is an exciting avenue of future

research.

7. Melting of clastic metasedimentary rocks

Clastic sedimentary rocks accumulate in many different tectonic settings, leading to

wide variation in their bulk compositions depending on provenance and maturity. Within

f
oo
sequences of deformed and metamorphosed sedimentary rocks (paragneisses), metapelites

and metagreywackes are particularly important as they represent the primary source rocks of
pr
peraluminous (‘S-type’) granites (Chappell, 1974; Chappell and White, 2001), although other
e-
source rocks may also produce minor quantities of peraluminous granite (Clarke, 2019;
Pr

Clemens, 2006; Gao et al., 2016a). This section focusses on these two key protoliths.

Metapelites have clay-rich protoliths (pelites) that accumulate in many subaqueous


al

environments, from abyssal plains to lakes. Metapelites are widely studied, in part because
n

during prograde metamorphism they develop a distinctive suite of index minerals (Barrow,
ur

1893; Barrow, 1912) that can be readily used to qualitatively constrain P–T conditions in
Jo

both subsolidus and suprasolidus rocks (Chinner, 1966; Johnson et al., 2001c). However, the

pelite component of most paragneiss sequences is subordinate to coarser-grained clastic

sedimentary rocks such as greywacke. Greywacke, a type of sandstone (or psammite), is a

term that covers a compositionally diverse suite of sediments with varying proportions of

different clastic components. As a consequence, the mineral assemblages and fertility of

metagreywackes are also widely variable, since both are sensitive to bulk composition

(Johnson et al., 2008).


Journal Pre-proof

If temperatures are sufficiently high, metapelites and metagreywackes are fertile

sources, although metapelites will produce a larger volume of melt at any given P–T

condition than metagreywackes (Johnson et al., 2008). Importantly, anatectic (and

subsolidus) metapelites and metagreywackes commonly contain a wide variety of mineral

chronometers, including garnet, zircon, monazite, titanite and apatite. Strategies linking the

ages derived from these minerals to stages of the evolution of melt-bearing rocks (i.e.

petrochronology) are explored in sections 9 and 10.

The presence of a large volume of peraluminous granite in orogenic systems—such as

f
oo
in the Himalaya (Harris et al., 1995)––requires the accumulation then burial of sediment to

pressures and temperatures sufficient for partial melting, although the timescale may still be
pr
rather short (Harris et al., 2000). Their presence commonly reflects late orogenic melting
e-
during continent–continent collision (Barbarin, 1996; Sylvester, 1998), although
Pr

peraluminous granites may also form during back-arc extension in accretionary orogens

(Collins and Richards, 2008). Peraluminous granite is scarce in the early to middle Archean
al

compared with TTGs (section 6), which may provide support for an origin unrelated to plate
n

tectonic processes. These Archean granites may represent low-volume partial melting of
ur

metabasic rocks, partial melting of sodic granites themselves or anatexis of metasedimentary


Jo

rocks (Johnson et al., 2017; Li et al., 2020; Watkins et al., 2007).

7.1. Field and petrographic observations

Metapelite and metagreywacke are commonly interlayered in metasedimentary

successions (Fig. 2e), which can lead to a complex distribution of leucosome (Brown and

Solar, 1998; Marchildon and Brown, 2003; Weinberg et al., 2013; Yakymchuk et al., 2013).

A common observation is that metapelite layers contain more leucosome than


Journal Pre-proof

metagreywacke layers, but this need not relate to the amount of melt generated in situ (Diener

and Fagereng, 2014). In the field, fluid-absent melting of metapelite and metagreywacke may

be identified based on the spatial association of leucosome and anhydrous peritectic minerals,

whereas a large proportion of leucosome retained in an outcrop may provide support for

fluid-present melting, which generally does not produce peritectic minerals. The

concentration and alignment of phyllosilicate minerals (biotite, muscovite) into layers

between leucosomes in some metasedimentary migmatites may create a potential barrier to

cross-layer melt migration (Laporte and Watson, 1995). However, some structures that

f
oo
develop in different orientations to the dominant fabric (e.g. axial planar foliation and shear

bands) can allow melt to locally cross these barriers (e.g. (Weinberg et al., 2013).
pr
In metasedimentary migmatites, the leucosome and melanosome generally record
e-
different information about the metamorphic history. Since the melanosome is dominated by
Pr

residual minerals, it is more likely to record important petrographic information about the

prograde to peak metamorphic P–T evolution. Furthermore, except at its margins, the
al

melanosome is less susceptible to the retrogression that accompanies local melt


n

crystallization (White and Powell, 2010). Notably, inclusions shielded in peritectic minerals
ur

(e.g. garnet) provide important information about the prograde history and also may host
Jo

accessory mineral chronometers that were not influenced by retrograde processes.

By contrast, the petrography of the leucosome usually provides information about the

cooling and melt crystallization history, providing insight into the processes that influenced

the evolution of the melt during fractional crystallization and deformation (Brown et al.,

2016; Koblinger and Pattison, 2017). Biotite-rich selvages adjacent to leucosomes in

metasedimentary migmatites are common features (Fig. 2e). Although they may have

multiple origins (Brown, 2002), in many cases these selvages are related to recrystallization

driven by diffusion of H2O from the leucosome into the melanosome as melt crystallizes
Journal Pre-proof

(White et al., 2008). However, the common preservation of close-to-peak mineral

assemblages in many migmatites and granulites demonstrates that extensive reaction with

melt is uncommon, consistent with melt drainage (Brown, 2002).

7.2. Compositions and typical minerals

The bulk composition of pelites varies depending on the relative proportions of clay

minerals, quartz, feldspar and other minerals. Clay minerals, such as kaolinite, illite and

f
oo
montmorillonite, are aluminous and hydrous, features that allow metapelites to develop

distinctive aluminosilicate mineral assemblages during metamorphism across a range of T/P


pr
conditions and to be fertile protoliths with respect to melt production. Oxidation state also
e-
plays a role in controlling the suprasolidus metamorphic mineral assemblage of metapelites
Pr

(Boger et al., 2012).

Metapelites can be divided into aluminous and subaluminous varieties based on their
al

position in the A(K)FM ternary diagram relative to a garnet–cordierite tie line (Fig. 16). The
n

average compositions of many shales (Condie, 1993) and their amphibolite-facies equivalents
ur

(Ague, 1991; Shaw, 1956) are aluminous, and most contain a higher molar proportion of Fe2+
Jo

than Mg (Fig. 16). Prograde subsolidus metamorphism is not an isochemical process, and the

bulk composition of metapelites immediately prior to partial melting has been modified from

that of the original pelite by the progressive loss of H2O and water-soluble components (e.g.

SiO2) to become more Al-rich (Ague, 1991). Aluminous metapelites contain peraluminous

minerals such as garnet and kyanite at high pressure, sillimanite at intermediate pressure, and

cordierite and orthopyroxene at low pressure. Suprasolidus andalusite is not expected to be

stable except at very low pressures and temperatures just above the solidus (Cesare et al.,

2003; Greenfield et al., 1996).


Journal Pre-proof

Although rare, magnesium-rich metapelites are important in studies of ultrahigh-

temperature (UHT) metamorphism because their compositions allow the development of

sapphirine-quartz and osumilite-bearing assemblages at UHT conditions (Harley, 2008;

Kelsey and Hand, 2015). The assemblage sapphirine + quartz is restricted to rare aluminous

bulk rock compositions with XMg > 0.5–0.6 (Kelsey and Hand, 2015). The Mg- and Al-rich

compositions are most likely inherited from a sedimentary protolith, but some may have been

produced from syn-metamorphic metasomatism (Kelsey and Hand, 2015). Overall, these

rocks represent a small proportion of common (meta)sedimentary sequences, but since they

f
oo
generally generate less melt than common metapelites, they play only a limited role in crustal

differentiation.
pr
Greywackes have highly variable compositions according to their tectonic setting,
e-
reflecting a combination of source characteristics (mafic island arcs versus more felsic
Pr

continental arcs) and sediment maturity. For example, more mature greywackes from passive

margin turbidite sequences may have >75 wt.% SiO2, indicating of a high proportion of
al

quartz fraction, whereas less mature greywackes from arc settings may contain much less
n

SiO2. Notwithstanding, greywackes generally contain more SiO2 than pelites, and average
ur

compositions are aluminous on an A(K)FM diagram (Fig. 16) and commonly develop similar
Jo

aluminous mineral assemblages to aluminous metapelites (Johnson et al., 2008; Yakymchuk

and Brown, 2014b). However, sub-aluminous compositions may lack sillimanite and contain

more biotite. Metaluminous variant develop hornblende- and clinopyroxene-bearing

assemblages above the solidus (Patiño Douce and Beard, 1996; Vielzeuf and Montel, 1994;

Vielzeuf and Schmidt, 2001), and relatively Na-rich compositions may stabilize

orthoamphibole (e.g. gedrite) at suprasolidus conditions (Johnson et al., 2008).

7.3. Reactions
Journal Pre-proof

We first consider the reactions crossed during suprasolidus heating in the simplified

KFMASH system. Note that the absence of Na and Ca from this modelled system precludes

consideration of plagioclase, which is a key reactant and product mineral in anatectic

reactions (Spear et al., 1999). Nevertheless, the KFMASH provides a useful system to

investigate the behaviour of the key ferro-magnesian minerals that participate in suprasolidus

reactions. A petrogenetic grid from White et al. (2014a) along with simplified A(K)FM

diagrams from (Spear et al., 1999) is shown in Fig. 17 (see also (Grant, 1985; Thompson,

f
oo
1982). Although this simplified system shows the key reaction relationships for metapelites

and metagreywackes, in natural systems these are multivariant (continuous) rather than

univariant equilibria.
pr
e-
Consider an isobaric heating path at 0.6 GPa in the KFMASH system (Fig. 17)––
Pr

quartz is considered to be in excess, so is not included in the reactions, although it is almost

always a reactant. The first melting reaction encountered with heating equates to crossing of
al

the wet solidus by the reaction:


n
ur

Bt + Ms + Kfs + H2O + Grt → Liq. (20)


Jo

In the absence of any external supply of fluid, the limiting reactant is usually H2O as the

amount of H2O present within pore spaces will be very small (< or <<1 vol.%; (Yardley and

Valley, 1997). As a result, the quantity of H2O-saturated melt produced is also small. Further

heating under fluid-absent conditions (aH2O < 1) results in the terminal breakdown of

muscovite via the incongruent melting reaction:

Ms + Grt → Sil + Kfs + Bt + Crd + Liq, (21)


Journal Pre-proof

which is the equivalent of reaction 12 in the simpler KASH system. The distinctive products

of this reaction are K-feldspar and, depending on pressure, sillimanite or kyanite. At higher

temperatures, biotite and sillimanite breakdown occurs via the incongruent melting reaction:

Bt + Sil → Grt + Crd + Kfs + Liq. (22)

which marks the transition to granulite facies conditions (Waters, 1988). In quartz-bearing

f
oo
aluminous pelites in KFMASH, biotite will be completely consumed by reaction 22, resulting

in the formation of a nominally-anhydrous mineral assemblage (garnet–sillimanite–-


pr
cordierite) coexisting with melt. However, in subaluminous pelites in KFMASH, sillimanite
e-
will be completely consumed before biotite by reaction 22, so that further heating leads to the
Pr

reaction:
al

Bt + Grt → Opx + Crd + Kfs + Liq. (23)


n
ur

Reaction 23 produces orthopyroxene–cordierite granulites, with or without biotite or garnet


Jo

(Johnson et al., 2001c). In aluminous pelites, at temperatures above those of reaction 22, the

reaction:

Grt + Sil → Spl + Crd + Kfs + Liq. (24)

leads to the formation of aluminous (hercynitic) spinel.

To summarize, melting initiates at the wet solidus, which consumes any free H2O.

Further heating causes muscovite-breakdown melting to produce K-feldspar and


Journal Pre-proof

aluminosilicate over a very narrow T interval such that it is effectively univariant, even in

natural rocks. At higher temperatures, melt is generated by the incongruent breakdown of

biotite and/or sillimanite via reactions that generate anhydrous (or low aH2O cordierite)

peritectic minerals. At temperatures above biotite and sillimanite exhaustion, melting

continues through the consumption of nominally anhydrous minerals. Although a

simplification, the KFMASH phase equilibria provide an extremely useful qualitative

framework in which to understand partial melting equilibria in natural pelites and greywackes

(Droop et al., 2003; Harley and Thompson, 2004; Johnson et al., 2001a; Kelsey et al., 2005;

f
oo
Pattison and Harte, 1984; Spear et al., 1999; Thompson Jr, 1957; White et al., 2001, 2007,

2014a).
pr
e-
7.4. Average metapelite
Pr

Now consider a P–T pseudosection calculated based on the composition of an average


al

amphibolite-facies metapelite (Ague, 1991) in the enlarged and more robust


n

MnNCKFMASHTO chemical system (Fig. 18a). The terminal muscovite-out reaction occurs
ur

at relatively low temperatures above the H2O-saturated solidus (at ~680 °C at P > 0.5 GPa
Jo

and ~780 °C at 1.2 GPa). Notably, orthopyroxene is not stable in the pseudosection for the

average metapelite due to its aluminous bulk composition (Figs 16, 18a). Garnet is stable

across most of the pseudosection; at mid-crustal pressures, biotite is absent above ~850 °C,

whereas at lower pressures and temperatures above ~730 °C, cordierite-bearing assemblages

are stable (Fig. 18b). In contrast to the univariant (discontinuous) reactions in the KFMASH

petrogenetic grid (Fig. 17), reactions in the MnNCKFMASHTO chemical system are almost

always continuous (F > or >> 1) and a closer approximation to how natural rocks behave.

The continuous nature of suprasolidus reactions can be visualized through mol.% isopleths
Journal Pre-proof

for a particular phase, such as mol.% melt, as shown in Fig. 18c, or using an isobaric T–molar

proportion diagram (Fig. 19a).

Consider isobaric heating at 0.6 GPa for a minimally-hydrated average amphibolite-

facies metapelite (arrow in Fig. 19a). The wet solidus is crossed at ~675 ºC consuming any

free H2O present, which is generally assumed to be minimal, by a reaction similar to reaction

20 (White et al., 2007). With further heating, the proportion of muscovite decreases slightly

from the wet solidus to ~700 ºC, where it is completely consumed within ~2 ºC by an

expanded version of reaction 21 in the simplified KFMASH system. During further heating,

f
oo
biotite, quartz, sillimanite and plagioclase are consumed to produce garnet, K-feldspar and

melt. At ~810 ºC cordierite is predicted to become stable and further heating consumes
pr
biotite, quartz and plagioclase to produce cordierite, K-feldspar, garnet and melt by an
e-
expanded version of reaction 22 in the simplified KFMASH system (Fig. 19a), which
Pr

consumes all biotite over a small temperature interval of ~3 ºC. Considering the strongly

endothermic nature of biotite-breakdown melting, a limiting factor may be the rate of heat
al

flow (Schorn et al., 2018). Notwithstanding, if heating continues after exhaustion of biotite,
n

consumption of quartz and feldspar continues to generate melt until quartz is exhausted at
ur

~910 ºC; spinel first forms at ~920 ºC (Fig. 19a).


Jo

Melt is generated continuously, albeit at an uneven rate, during suprasolidus heating

in metapelites (Fig. 18c). In addition, for a closed system, melt is also generated during

decompression due to the positive slopes of melt isopleths in P–T space (Fig. 18c; see section

5.5 for open system scenarios). For the average metapelite, ~67 mol.% melt is generated

during isobaric heating between the wet solidus at ~675 ºC and ~950 ºC, an average of ~0.2

mol.% melt/ºC along the modelled path. However, there are two abrupt increases in the

proportion of melt at ~700 ºC and ~810 ºC. The first is related to the muscovite-out reaction,

which occurs over a narrow temperature range (similar to reaction 21), generating ~3 mol.%
Journal Pre-proof

melt between ~700 and ~702 ºC. The second is during the growth of cordierite at the expense

of biotite (via a reaction similar to 22), which generates ~5 mol.% melt over ~3ºC (Fig. 19a).

Heating at higher pressures will generate less melt at a given temperature than at lower

pressure (Fig. 18c). However, the amount of melt and the reaction sequence experienced by

natural metapelites will vary depending on the particular bulk rock composition.

7.5. Greywacke

f
oo
Greywackes can be more compositionally diverse than pelites and experience different

reaction sequences, have different mineral assemblages, and be variably fertile (Johnson et
pr
al., 2008). Consider an average passive margin greywacke that is representative of turbidite
e-
sequences commonly involved in orogenesis at plate margins. An MnNCKFMASHTO P–T
Pr

pseudosection for this composition is shown in Figure 20a. Muscovite is present at low

temperatures and P > 0.75 GPa, biotite is absent at high temperatures, cordierite is stable at
al

low pressures, and garnet is stable over the entire P–T range shown (Fig. 20b). Notable
n

differences include the absence of a field representing muscovite breakdown to K-feldspar


ur

and melt (reaction 21) and the presence of orthopyroxene at low pressure due to the less
Jo

aluminous nature of this composition (Figs 16, 20b).

Along an isobaric heating path at 0.6 GPa for the greywacke (Fig. 19b), muscovite is

not stable having been consumed under subsolidus conditions. Melting begins at the wet

solidus at ~680 ºC by a reaction similar to reaction 20, consuming the small amount of free

H2O and producing ~2 mol.% melt. With increasing temperature, biotite is partially

consumed to produce garnet and melt. Cordierite first appears at ~790 ºC by the reaction of

biotite and sillimanite to cordierite, K-feldspar, minor garnet and melt (via a reaction similar

to 22) over a narrow temperature range (~5 ºC). Further heating results in the complete
Journal Pre-proof

consumption of biotite at ~840 ºC and K-feldspar at ~905 ºC (Fig. 19b). At higher

temperatures melting continues through the breakdown of quartz, plagioclase and cordierite.

Orthopyroxene is not stable along this heating path but it is predicted to be stable at lower

pressures, where crossing of the narrow field containing coexisting orthopyroxene and biotite

is similar to the KFMASH reaction 23 (Fig. 20b).

The proportion of melt in the greywacke increases continuously along the modelled

isobaric heating path, although slightly lower melt productivity is predicted with increasing

pressure (Fig. 20c). From the wet solidus at ~680 to ~950 ºC, ~33 mol.% melt is predicted, an

f
oo
average rate of ~0.1 mol.%/ºC, roughly half that in the metapelite. One abrupt increase in

melt proportion occurs at ~795 ºC by a reaction similar to 22 that produces ~2 mol.% melt

over ~5 ºC (Fig. 20c).


pr
e-
Pr

7.6. Melt compositions


al

Melt compositions predicted from the phase equilibrium modelling of metapelite and
n

greywacke vary with pressure and temperature, reflecting differences in the starting
ur

compositions and the evolving residual mineral assemblages. We emphasize that the wide
Jo

array of possible metapelite and greywacke starting compositions means that the range of

melt compositions is wider than those discussed here.

For the modelled metapelite and metagreywacke, the concentration of SiO2 in the

melt increases with heating (Fig. 21a, d). This partly reflects the dilution of H2O at higher

melt fractions with increasing temperature. At temperatures above quartz exhaustion (Fig.

21a), the concentration of SiO2 is expected to decrease with further heating because quartz is

no longer present to buffer the composition of the melt, which has implications for the use of

trace element equilibria that rely on the activity of silica, such as the Ti-in-zircon (Ferry and
Journal Pre-proof

Watson, 2007; Kirkland et al., 2021) and Zr-in-titanite thermometers (Hayden et al., 2008;

Kirkland et al., 2020).

Over the modelled P–T range, all melt compositions derived from the metapelite and

greywacke are peraluminous (Figs 21b, e; Fig. 22a), i.e. molar Al2O3/[CaO + Na2O + K2O] >

1 (Clarke, 1992; Shand, 1927). However, this is unlikely to be the general case, especially for

sub-aluminous greywacke compositions. For the modelled aluminous compositions, the

A/CNK values increase with temperature resulting in progressively more peraluminous melt

compositions (Fig. 21b, e), which eventually reach the ‘strongly’ peraluminous field at >800

f
oo
ºC (c.f. (Clarke, 2019). The presence or absence of sillimanite plays a role in controlling the

A/CNK value of the melt, as indicated by the inflection in A/CNK contours in Fig. 21b, e.
pr
The molar proportion of Fe, or Fe# (molar Fe2+/(Fe2+ + Mg)), is used to differentiate
e-
ferroan from magnesian granites (Frost et al., 2001). All of the modelled melt compositions
Pr

are magnesian, which includes granites with Fe# < 0.75 at the modelled concentrations of

SiO2 (Frost et al., 2001). The Fe# of melt is expected to decrease up temperature due to the
al

preferential partitioning of Fe into garnet for the modelled metapelite and greywacke. After
n

exhaustion of biotite, the Fe# initially decreases then increases with continued heating (Fig.
ur

21c, f).
Jo

The variable proportions of plagioclase to clay minerals in pelite versus psammite

source rocks has been used to develop a geochemical proxy for granites generated from these

two rock types (Chappell and White, 1992; Sylvester, 1998). Sylvester (1998) proposed that

strongly-peraluminous pelite-derived melts generally have low CaO/Na2O ratios (<0.3),

whereas psammite-sourced melt has higher ratios (>0.3). These different ratios are a

consequence of the behaviour of plagioclase and the Ca-rich ferromagnesian minerals during

partial melting. However, the compositional diversity of both pelites and psammites is greater
Journal Pre-proof

than the compositions considered by Sylvester (1998), such that the CaO/Na2O ratio may be

less useful than envisaged.

For the two compositions modelled, the metapelite is predicted to generate melt

compositions with CaO/Na2O ratios from <0.1 to >0.3 (Fig. 22b), whereas the greywacke is

predicted to yield melt compositions with ratios <0.3 (Fig. 22b), inconsistent with the 0.3

threshold of Sylvester (1998). The reason for this difference is that the modelled

compositions have similar and sufficient quantities of both plagioclase and garnet (Fig. 19),

whereas the pelites considered by Sylvester (1998) had substantially less plagioclase, leading

f
oo
to exhaustion of plagioclase at lower temperatures. Loss of plagioclase means Na2O

partitions into the melt whereas CaO entered peritectic garnet, resulting in lower CaO/Na2O
pr
ratios in the melt. In the average metapelite modelled here, plagioclase is stable to ~950 ºC
e-
(Fig. 18a), such that the CaO/Na2O ratio in melt does not decrease. Thus, although the
Pr

amount of plagioclase in the source may be an important control on melt composition and in

distinguishing pelitic from psammitic sources, the situation is more complex, and other trace
al

element proxies (e.g. Rb/Ba and Rb/Sr; Sylvester, 1998) may be required to distinguish the
n

putative source.
ur

Lastly, as a check on how close to natural melt compositions those derived using
Jo

phase equilibrium modelling are, we may compare the compositions obtained from

experiments on crustal rocks or synthetic compositions with those derived by modelling of

the same starting materials (Johnson et al., 2008; White et al., 2011; Bartoli and Carvalho,

2021; Bartoli, 2021). Furthermore, the degree of sophistication now achieved in phase

equilibrium modelling means that this approach also has the potential to highlight not only

differences, but also problems with the experiments (Johnson et al., 2008; White et al., 2011;

Bartoli, 2021; Bartoli and Carvalho, 2021). Indeed, the recent study by Bartoli (2021) has

confirmed that muscovite was a metastable phase in the experiments of Patiño Douce and
Journal Pre-proof

Harris (1998) and that, contrary to the original authors’ assertions, that metastability seriously

affected melt compositions.

8. Melting during exhumation from ultrahigh-pressure metamorphic conditions

In orogenic belts, crustal rocks that have been returned to shallow depths following

deep subduction to ultrahigh-pressure (UHP) conditions—defined by the stability of coesite

f
oo
rather than quartz—commonly show features interpreted to record partial melting. These

features include multiphase solid inclusions (including nanogranites) and/or glassy inclusions
pr
in rock-forming and accessory minerals, microstructures in which quartz or feldspar infills
e-
grain boundaries, commonly with low dihedral angles at triple grain-boundary junctions, and
Pr

macroscopic leucosomes (Feng et al., 2021; Ferrero et al., 2019; Gao et al., 2017b; Wang et

al., 2016, 2017, 2020). However, where and when during the P–T evolution these features
al

formed, and whether they are related to the generation and evolution of supercritical fluid, or
n

fluid-present or fluid-absent melting remain open questions.


ur

With changing P–T conditions, silicate-buffered fluids at high-pressure (HP) and


Jo

UHP metamorphic conditions vary dramatically in composition and properties, including

solute content, density, polymerization and viscosity (Hack and Thompson, 2011; Mysen,

2014). As pressure increases, at temperatures below the wet solidus minerals become more

soluble in aqueous fluid, whereas at temperatures above the wet solidus silicate melts can

accommodate more H2O in their structure (Fig. 23). Enhanced mutual solubility implies that

aqueous fluids and silicate melts become increasingly similar until pressure exceeds the

second critical endpoint for the particular silicate–H2O system (Fig. 23, SCE 1, 2 or 3,

according to composition), where the solvus between fluid and melt closes, and a discrete
Journal Pre-proof

solidus can no longer be distinguished (Hack et al., 2007; Hermann et al., 2006). Along the

critical line, which extends from each critical endpoint to higher temperatures, the solute

content increases (Fig. 23). At pressures above the second critical endpoints and critical lines,

miscibility yields a supercritical fluid that varies continuously from aqueous fluid at lower

temperatures and solute content <30% to supercritical fluid with increasing solute content

and properties similar to hydrous melt at higher temperatures (Fig. 23).

Informally, we distinguish sub-critical from supercritical P–T conditions. At sub-

critical P–T conditions (i.e. at pressures below the second critical endpoints or critical lines,

f
oo
but at temperatures above the fluid-saturated solidus for common crustal source rock

compositions), a saturated hydrous melt coexists with an aqueous fluid (Fig. 23), whereas at
pr
supercritical P–T conditions (i.e. at P–T conditions above the second critical endpoints or
e-
critical lines), hydrous melt and aqueous fluid are completely miscible and are replaced by a
Pr

single supercritical fluid (Fig. 23). In particular, even at sub-critical pressures above ~1 GPa,

silicate-rich fluids/H2O-rich melts become largely depolymerized, decreasing viscosity and


al

increasing mobility (Audétat and Keppler, 2004).


n

Partial melting at HP or UHP metamorphic conditions is sometimes argued to have


ur

been fluid-present, but more commonly is inferred to have been related to fluid-absent
Jo

hydrate breakdown reactions, such as Ph + Cpx + Qz → Bt + Pl + Grt + L (Auzanneau et al.,

2006), even where microstructural evidence of hydrate breakdown is lacking. However, in

some of these cases the P–T conditions were higher than those of the second critical endpoint

for the protolith composition (Fig. 23), suggesting that leucosomes may have crystallized

from supercritical fluids or from hydrous melts that evolved from supercritical fluids. Better

characterisation of where these leucosomes form along the P–T path and by what mechanism

is of great interest, because: (1) ascending fluid/melt transports mass and heat, which

facilitates hydration and retrogression of UHP metamorphic rocks at shallower HP eclogite


Journal Pre-proof

and amphibolite facies conditions (Chen et al., 2007; Wang et al., 2017; Zheng et al., 2009);

and, (2) the presence of melt lowers the viscosity of the crust and promotes exhumation

(Brown and Rushmer, 2006b; Ferrero et al., 2015; Gerya and Meilick, 2011; Labrousse et al.,

2011, 2015; Sizova et al., 2012). Thus, the evolution of supercritical fluid to hydrous melt

during decompression from UHP to HP conditions (i.e. crossing from the coesite to the

quartz stability field; Fig. 23) may be an underappreciated mechanism assisting the

exhumation of deeply-subducted crust (Wang et al., 2020). In addition, drainage of fluid

and/or melt from deeply-subducted lithosphere may lead to substantial metasomatic

f
oo
modification of the overlying crust and/or mantle (Borghini et al., 2020; Hermann and

Rubatto, 2014; Hermann et al., 2006, 2013; Keppler, 2017; Malaspina et al., 2006, 2009,
pr
2010; Scambelluri et al., 2008; Scambelluri and Philippot, 2001; Zheng, 2012).
e-
During subduction and prograde metamorphism of the lithosphere, any fluid and most
Pr

melt generated via hydrate-breakdown reactions is generally lost, which may lead to fluid-

deficient or fluid-absent conditions approaching peak pressures (Yardley and Valley, 1997).
al

Notwithstanding, some hydrous minerals are stable at HP and UHP conditions, including
n

phengitic mica, sodic amphibole, epidote–zoisite and lawsonite, which collectively may
ur

retain significant amounts of H2O at upper mantle depths. Furthermore, at peak UHP
Jo

conditions eclogites may store up to several hundred ppm primary 'water' (as OH;

colloquially water) structurally bound within nominally anhydrous minerals (NAMs),

including pyroxene, garnet and rutile in eclogites and, possibly, metastable feldspar in

gneisses (Wang et al., 2018). In addition, molecular H2O may be preserved as fluid

inclusions. In these circumstances, since the solubility of water in NAMs generally decreases

with decreasing pressure, during exhumation from pressures above the second critical end

point exsolution of water from NAMs (strictly dehydroxylation) will generate a supercritical

pore fluid (Hermann and Rubatto, 2014; Hermann et al., 2013; Wang et al., 2017, 2020;
Journal Pre-proof

Zheng et al., 2011; Zheng and Hermann, 2014). With ongoing exhumation, this fluid

dissolves more solute, becoming denser, more viscous and more polymerized until it stalls

and partially or completely crystallizes. Exactly how the fluid evolves during exhumation

depends not only on the P–T path but also on: (1) the variance of the phase assemblage—the

host silicate mineral assemblage in equilibrium with the supercritical fluid—because in high

variance assemblages (components >> phases) in which minerals exhibit extensive solid

solution, the fluid composition is not buffered by the mineral assemblage alone but also

depends on the rock/fluid ratio; and, (2) whether the fluid has sufficient volume to migrate

f
oo
from one host rock into another, which may lead to solute loss or gain by precipitation or

dissolution in the new chemical environment (Hack et al., 2007).


pr
In contrast to Hermann et al. (2006), we do not arbitrarily separate supercritical fluid
e-
from hydrous melt at 65 wt% solute (Frezzotti and Ferrando, 2015). With decreasing pressure
Pr

and increasing temperature along the critical line (Fig. 23) solute content increases from ~65

wt% at the second critical endpoint to ~95 wt% at 900 °C, and the supercritical fluid
al

increasingly has properties closer to those of hydrous melt with, for example, a fivefold
n

increase in viscosity (Audétat and Keppler, 2004; Hack and Thompson, 2011). Although we
ur

acknowledge that at high temperatures the difference between a high-solute supercritical fluid
Jo

and a hydrous melt could be considered semantic, the advantage of thinking in terms of a

supercritical fluid that evolves with decreasing P or increasing T to hydrous melt is that the

process involved is solute dissolution, and distinguishing between fluid-present and fluid-

absent (hydrate breakdown) melting becomes moot.

The distinction between supercritical fluid and hydrous melt has implications for the

generation of leucosomes during decompression from UHP to HP metamorphic conditions,

which may represent crystallized supercritical fluid or cumulate material precipitated from

supercritical fluid rather than melt or cumulate material precipitated from melt. Such an
Journal Pre-proof

origin has been proposed for leucosome sheets and composite granite–quartz veins that occur

in UHP eclogite in the central Sulu belt of eastern China (Wang et al., 2017, 2020). In this

case, the leucosomes crystallized at pressures from 3.5 to 2.2 GPa, well above the pressure of

the second critical endpoint for felsic compositions, supporting their origin from supercritical

fluid.

More than three dozen UHP terrains have been identified during the last four decades,

many of which contain leucosome interpreted to have formed by either fluid-present or fluid-

absent partial melting during exhumation. We anticipate that leucosomes in many of these

f
oo
UHP metamorphic belts will be found to have formed by crystallization from a solute-rich

supercritical fluid or hydrous melt itself derived from an evolving solute-rich supercritical
pr
fluid during exhumation. For example, leucosomes in migmatitic paragneisses on an island in
e-
Jøkelbugt at the eastern margin of the North-East Greenland eclogite province (Gilotti, 1993)
Pr

were generated during decompression from P–T conditions of 3.6 GPa and 970 °C (Lang and

Gilotti, 2007; Lang and Gilotti, 2015). Melting was inferred to have occurred by fluid-absent
al

phengite-breakdown at P between 2.7 and 2.2 GPa, but it may be more likely that leucosomes
n

crystallised from a supercritical fluid formed by dehydroxylation of nominally anhydrous


ur

minerals in the inferred peak metamorphic mineral assemblage. Other examples of


Jo

metasedimentary rocks decompressing from extreme P–T conditions include diamondiferous

quartzo-feldspathic rocks from the Saxonian Erzgebirge, Germany (~7 GPa, >1200 °C;

Massonne, 2003; Massonne and Fockenberg, 2012) and paragneisses from the Kokchetav

Massif, Kazakhstan (>4.5 GPa, 1000 °C; Stepanov et al., 2014, 2016). Under these extreme

conditions, whether the processes involved are magmatic or metamorphic, and whether the

liquid is a solute-rich supercritical fluid or a hydrous melt are difficult to distinguish and

largely semantic.
Journal Pre-proof

Migmatitic continental crust, including eclogite, in the Western Gneiss region of

Norway has been interpreted to record fluid-present melting coeval with peak P–T conditions

(>2.5 GPa, ~800 °C; Ganzhorn et al., 2014; Labrousse et al., 2011, 2015). Although

continuous release and loss of fluid during progressive metamorphism likely led to fluid

absent conditions at high pressures (Yardley and Valley, 1997), infiltration of fluid associated

with deformation during subduction and/or exhumation could have occurred. Additional

sources of fluid during exhumation include dehydroxylation of nominally anhydrous minerals

and breakdown of hydrous minerals. Thus, any fluid attending peak metamorphism and

f
oo
exhumation was likely to have been solute-rich supercritical fluid that could have increased

in volume during exhumation and evolved to a hydrous melt at P < 1.5 GPa. This expectation
pr
is consistent with zircon U–Pb geochronology and trace element concentrations that show
e-
earlier leucosomes crystallized in equilibrium with garnet in the absence of plagioclase,
Pr

whereas younger leucosomes crystallized in the absence of garnet but in the presence of

plagioclase, suggesting an evolution during crystallization from higher to lower pressure


al

(Gordon et al., 2013). Similarly, a recent study of scapolite pegmatites from the northern
n

segment of the Western Gneiss region (Bryden and Jamieson, 2020) considers a possible
ur

supercritical fluid source in the petrogenesis of the precursor melts, even though
Jo

crystallization occurred at much shallow crustal depths under amphibolite facies conditions.

In summary, leucosomes and pegmatites in UHP eclogites provide insight into the

generation and evolution of supercritical fluid and/or hydrous melt during exhumation of

deeply subducted continental crust. In many cases, leucosomes and pegmatites may have

crystallised from hydrous melt that evolved from a supercritical fluid. As supercritical fluid

generated by exsolution of water from nominally anhydrous minerals during decompression

increases in volume, it may become interconnected along grain boundaries and drain through

the exhuming crust. As it migrates, the fluid will evolve to a denser, more viscous and more
Journal Pre-proof

polymerized hydrous melt by dissolution of the silicate matrix. During progressive

entrapment of the resulting hydrous melt across a crystallization interval beginning at the

critical line for the appropriate chemical system (Fig. 23), the chemical potential of H2O will

be higher in the hydrous melt than in the host rocks, promoting diffusion of H2O from the

melt into the host rocks driving crystallization of the melt and retrogression of the hosts

(White and Powell, 2010). After solidification, some leucosomes may record evidence of

limited partial melting due to phengite breakdown (e.g. Fig. 3b) during ongoing exhumation

to amphibolite facies conditions. However, at these P–T conditions, this reaction only

f
oo
produces a small volume of melt (<< the MCT) that likely will remain trapped in the host

(Wang et al., 2017, 2020).


pr
e-
Pr

9. Behaviour of trace elements in anatectic systems


al

9.1. Importance of trace elements as petrogenetic indicators


n
ur

Characterizing the behaviour of trace elements during crustal melting provides a tool
Jo

to understand the petrogenesis of the continental crust and secular change in its composition.

Empirical, experimental and modelling approaches are used to determine which elements

partition into anatectic melt and which remain in the residue. Here, we describe modelling

trace element behaviour during anatexis, explore the assumptions and limitations of this

approach and discuss some trace element modelling applications to migmatites. The

behaviour of some trace elements is also controlled by the dissolution and growth of

accessory minerals, which is discussed in more detail in section 10.


Journal Pre-proof

9.2. Modelling methodology

Trace element concentrations in melt and its complementary residue are modelled

following the methodology established for aqueous and igneous processes using a mass

balance approach coupled with Nernst partition coefficients (Kd):

Kd =Cmineral
i /Cmelt
i . (25)

f
oo
where Cmineral
i is the concentration of trace element i in a mineral and Cmelt
i is the

pr
concentration of this element in the melt. Concentrations are in weight fractions (e.g. mg/kg)

but the formula weights cancel each other such that Kd is dimensionless.
e-
We can model a variety of modal and non-modal melting scenarios, such as batch
Pr

melting, fractional melting, and continuous melting (Hanson, 1978; Shaw, 1970). Mass

balance between solid (residue) and liquid in a system dictates that:


n al

C0 = F CL + (1 – F) Cs. (26)
ur
Jo

where C0 is the concentration of element C in the system, CL is the concentration in the liquid

and Cs is the concentration in the solid, and F is the weight fraction of melt in the system.

This equation may be combined with partition coefficients (Kd) as follows, to determine the

distribution of the trace element between the solid assemblage and the equilibrated residue:

D = X.Kd = CS/CL. (27)


Journal Pre-proof

where D is the bulk solid (residue)–melt partition coefficient and X is the weight fraction of

each mineral in the solid residue. The equation for batch melting may be derived by

combining equations 26 and 27 into:

CL/C0 = 1 / (D (1 – F) + F). (28)

The concentration of a trace element in melt is a function of the concentration of the trace

element in the system, the bulk solid–melt partition coefficient, and the melt fraction. The

f
oo
concentration of the trace element in the solid residue may be calculated from the mass

balance equation above. Because the proportion of minerals in the residue changes with
pr
progressive melting (see sections 6 and 7), this requires changes to X and F (and Kd; see
e-
section 9.3) in the equations above at each modelled P–T condition.
Pr

Modelling trace element behaviour during melting requires: data for mineral/melt

partition coefficients (Bédard, 2006b; Hanson, 1978) or expressions that account for changes
al

in partition coefficients with P, T and composition (Bédard, 2007); the trace element
n

composition of the system (e.g. a measured whole-rock composition); and, the weight
ur

fractions and composition of minerals and melt (and possibly fluid) at specified P–T
Jo

conditions. With the exception of the partition coefficients, these variables can be retrieved

from phase equilibrium modelling or estimated based on petrographic observations (Sawyer,

1987). The calculations can be done manually, but automated software packages are being

developed that integrate phase equilibrium modelling with trace element partitioning (Mayne

et al., 2020b). Most, although not all, partition coefficients are determined through

experiments.
Journal Pre-proof

Open system behaviour can be modelled to account for the extraction of a single

batch or multiple batches of melt through mass balance. For example, the extraction of 6 wt%

melt from a rock with 7 wt% melt will result in a new system concentration as follows:

C0 = (93/94) CS + (1/94) CL. (29)

The new calculated concentrations can then be applied to the next step in the melting process

and used to evaluate the changing composition of the residue and different batches of melt

f
oo
generated through open system melting.

9.3. Assumptions and limitations


pr
e-
Trace elements that are concentrated in rock-forming minerals (e.g. the large-ion
Pr

lithophile elements: K, Rb, Sr, Cs, Ba) can be modelled based on mass balance and partition

coefficients. However, the behaviour of trace elements concentrated primarily in accessory


al

minerals—such as most high-field-strength elements (e.g. Zr, Hf, Th)—is more difficult to
n

model given uncertainties in their behaviour during melting (Yakymchuk et al., 2017). In
ur

addition, there are several key assumptions that go into trace element modelling in anatectic
Jo

systems, including: (1) that trace elements behave according to Henry’s law; (2) that the

partition coefficient (or expression for how it changes) is applicable to the P–T–X conditions

of the modelled system; and, (3) that all minerals are in equilibrium with melt and each other.

An underlying assumption in using partition coefficients is that concentrations of the

trace element of interest are low enough that they follow Henry’s law (Mysen, 1978; Wood

and Fraser, 1976), whereby the concentration in the mineral is proportional to its

concentration in the melt. The threshold for Henry’s law behaviour varies for different

elements and minerals, but it is generally <<1 wt% of the element concerned. Therefore,
Journal Pre-proof

modelling the behaviour of trace elements in rock-forming minerals, for example Yb in

garnet (Johnson et al., 2017; Semprich et al., 2015; Yakymchuk et al., 2020), is more robust

using this approach than modelling trace elements that can reach wt% levels in minerals

where they are essential structural constituents, such as Th in monazite (Skora and Blundy,

2010; Yakymchuk et al., 2018).

Partition coefficients will vary as a function of temperature and pressure, and the

compositions of melt and minerals. For some elements, such as Y in garnet, these variations

are small or negligible (Pyle et al., 2001). For others, a careful parameterization of partition

f
oo
coefficients is needed. Consider trace element partitioning between plagioclase and melt.

Partitioning of Sr between plagioclase and melt is influenced by the anorthite content of


pr
plagioclase and pressure, whereas partitioning of REE between plagioclase and melt is
e-
sensitive to the H2O content of the melt (Bédard, 2006b). There is still a paucity of partition
Pr

coefficients for key high-temperature metamorphic minerals. For example, the available

partition coefficients for trace elements in cordierite are based on an empirical approach
al

comparing the concentrations of trace elements in minerals and those in leucosome (Bea et
n

al., 1994), but for a variety of reasons the leucosome may not have represented the
ur

composition of equilibrated melt (Sawyer, 1987; Yakymchuk, 2019). In addition, variable


Jo

speciation of trace elements sensitive to oxidation state (e.g. Eu3+ and Eu2+) requires multiple

partition coefficients (Holder et al., 2020).

A crucial assumption in the modelling of trace elements in anatectic systems (e.g.

migmatites and most granulites) is that minerals and melt achieved perfect equilibrium. In

such circumstances, phases are compositionally homogenous, can adjust their compositions

and modes instantly to changes in P–T–X conditions with no kinetic hindrances, and are

available to the reacting volume (i.e. there is no shielding of some minerals as inclusions in

others). In natural migmatites, minerals may record zoning in trace element concentrations
Journal Pre-proof

(Jung and Hellebrand, 2006; Pyle and Spear, 1999, 2003; Spear and Kohn, 1996) and

accessory minerals commonly occur as inclusions in rock-forming minerals (Watson, 1996).

Some of these features can be accounted for, such as REE zoning in garnet (Kendrick and

Yakymchuk, 2020) and inclusion of accessory minerals in rock-forming minerals

(Yakymchuk and Brown, 2014a), but others are more difficult to model, such as the kinetic

hindrance to the dissolution of accessory minerals in melt (Harrison and Watson, 1983, 1984;

Rapp and Watson, 1986). In general, high-temperature metamorphic rocks, such as

migmatites and granulites, are less susceptible to kinetic barriers (Carlson et al., 2015).

f
oo
However, careful consideration of trace element diffusivities in the modelled minerals at the

temperatures and timescales of interest is required. Finally, the different diffusivities of major
pr
and trace elements in minerals and melt makes the approach of using a single ‘effective bulk
e-
composition’ (Lanari and Engi, 2017) problematic, especially when rocks preserve
Pr

microstructural evidence of chemical potential gradients, such as symplectites and coronae.


al

9.4. Applications
n
ur

Studies of the concentrations and distributions of trace elements in migmatites and


Jo

granulites provide key insights into: (1) establishing the protoliths of these rocks; (2)

elucidating open- and closed-system processes that may have occurred during anatexis

(Sawyer, 1987; Slagstad et al., 2005); (3) linking the growth of mineral chronometers to P–T

paths (i.e. petrochronology; (Kylander-Clark et al., 2013); (4) establishing the degree of

melting and melt loss (Guernina and Sawyer, 2003); (5) understanding how melt exfiltration

leads to compositional stratification of the continental crust (Morfin et al., 2013; Sawyer,

2020; Yakymchuk and Brown, 2019); (6) assessing chemical equilibrium during melt–

residue separation (Nabelek, 1999; Sawyer, 1991); and, (7) linking granites to their source
Journal Pre-proof

rocks (Iles et al., 2018; Johnson et al., 2018). Many of these applications require an

assessment of melt compositions, particularly during the early stage of anatexis, and of melt

that subsequently drained from the system.

The first step in any study of an anatectic system is to develop a sampling strategy, as

discussed by Sawyer (2008). It is also important to estimate an initial composition for the

melt prior to melt extraction, which can be accomplished using trace element concentrations

(e.g. REE, LILE) of whole rocks, minerals and nanogranites (Bartoli et al., 2016) or glass

compositions from experimental studies on similar rocks (Beard and Lofgren, 1991; Gardien

f
oo
et al., 1995). Most leucosome compositions do not represent unmodified melt; more

commonly they represent a product of fractional crystallization, and/or melt–residue


pr
interaction and/or open-system infiltration of melt (Brown et al., 2016; Morfin et al., 2014;
e-
Sawyer, 1987, 2020). Trace element concentrations are generally complementary to the
Pr

major oxides in deciding whether leucosome compositions represent unfractionated anatectic

melt, cumulate-rich magma or complementary fractionated and evolved compositions


al

(Brown et al., 2016; Morfin et al., 2014; Solar and Brown, 2001b). These processes are
n

particularly well expressed in some diatexite terrains, where compositions vary from residue-
ur

rich diatexites that have lost melt to strongly-fractionated leucogranites (Milord et al., 2001;
Jo

Sawyer, 1998, 2020).

A powerful application of trace element modelling is inferring the P–T conditions of

partial melting of source rocks that produced granite, particularly whether garnet was present

or not in the source during melting. This approach has proven particularly useful in

estimating pressures related to the generation of tonalite–trondhjemite–granodiorite (TTG)

melts from hydrated metabasites in Archean cratons (Johnson et al., 2017, 2019; Moyen and

Stevens, 2006, 2013; Yakymchuk et al., 2020). Furthermore, some of the complexities in
Journal Pre-proof

trace element behaviour related to garnet fractionation and drainage of melt in the source can

be integrated into the modelling (Kendrick and Yakymchuk, 2020).

Petrochronology involves integration of the trace element compositions of dated

accessory minerals and particular rock-forming minerals that sequester (during growth) or

release (during breakdown) these elements during metamorphism (Taylor et al., 2016). For

example, heavy rare earth element (HREE) concentrations in zircon can be linked to the

growth and breakdown of garnet (Rubatto, 2002; Rubatto et al., 2013; Rubatto and Hermann,

2007; Taylor et al., 2015a) and equilibrium between these minerals may be assessed using

f
oo
REE partitioning arrays (Taylor et al., 2017). Other examples include Eu anomalies in

monazite as a monitor for the behaviour of feldspar during melting and melt crystallization
pr
(Dumond et al., 2015; Rubatto, 2002), linking Y concentrations in monazite to garnet growth
e-
and breakdown (e.g. Hacker et al., 2019), linking Th/U ratios in zircon to the behaviour of
Pr

monazite (Harley and Kelly, 2007; Harley et al., 2007; Rubatto, 2017; Yakymchuk et al.,

2018), and tracking oxidation state using Eu anomalies in major and accessory minerals
al

during partial melting (Holder et al., 2020; Yakymchuk et al., 2019a).


n

Three important considerations in applying petrochronology to anatectic systems are:


ur

(1) that whole-rock compositions will influence the trace element concentrations in accessory
Jo

minerals (Kaczmarek et al., 2008; Schaltegger et al., 1999; Yakymchuk et al., 2019a), which

is why absolute values for concentrations are less reliable than the ratios between the

accessory and rock-forming minerals of interest (Taylor et al., 2017); (2) that enhanced

mobility of Pb in zircon at high temperature (>900ºC) combined with the limited diffusivity

of many trace elements in accessory minerals may result in ages that are decoupled from

trace element compositions (Flowers et al., 2010; Kunz et al., 2018); and, (3) that equilibrium

between accessory and rock-forming minerals may have been hindered by kinetic issues

associated with mineral dissolution in melt (Watt et al., 1996), as well as sluggish
Journal Pre-proof

intracrystalline diffusion of trace elements in accessory minerals (Bea, 1991) and possibly

melt (Acosta-Vigil et al., 2017), and armouring of accessory minerals in residual or growing

minerals (Watson, 1996). Notwithstanding these limitations, petrochronology applied to

anatectic systems provides key constraints on the absolute timing of rock-forming mineral

growth that is difficult to achieve with other methods. An example of forward modelling

accessory mineral compositions is presented in section 10.

Trace element geochronology of rock-forming minerals (e.g. garnet) is now routine,

avoiding the assumptions and limitations of linking accessory mineral ages and

f
oo
concentrations of trace elements with the growth and breakdown of the P–T sensitive rock-

forming minerals. Garnet geochronology can be measured with two radioactive isotope
pr
systems: 147Sm–143Nd and 176Lu–176Hf (Baxter et al., 2017; Cheng et al., 2018). In general,
e-
the Lu–Hf system has a higher closure temperature (Bloch et al., 2015; Scherer et al., 2000;
Pr

Smit et al., 2013) and it may be more robust for estimating high-temperature metamorphic

processes, such as the timing of garnet growth and partial melting. Although the ages
al

retrieved from these techniques may be quite precise, how they relate to the P–T path is not
n

always straightforward, especially if whole garnet aliquots were analysed. In these


ur

situations, the spatial zoning of the parent isotope in garnet provides additional context––if
Jo

the parent isotope (Lu or Sm) is concentrated in the garnet cores, then the age will be biased

towards the early stage of garnet growth, whereas high concentrations in garnet rims may

bias ages towards the later stages of garnet growth (commonly the metamorphic peak).

Alternatively, rims may be enriched due to garnet resorption and preferential retention of the

parent isotope, which complicates a simple age interpretation. Avoiding these complexities

may require micro-sampling of garnet core and rim domains (Dragovic et al., 2015; Ducea et

al., 2003), but this approach may be limited by the size of the garnets and the concentration

of parent and daughter isotopes in the micro-sampled aliquots. Nonetheless, garnet


Journal Pre-proof

geochronology is a powerful tool especially when linked with the results of accessory

mineral geochronology.

10. Accessory Minerals

10.1 Importance of accessory minerals to studies of anatectic systems

f
oo
The importance of accessory minerals far outweighs their abundance in common rock

types. They are the main hosts for multiple trace elements (Ayres and Harris, 1997; Bea et
pr
al., 1994; Spear and Pyle, 2002; Ward and Werner, 1964; Zeng et al., 2005) and volatile
e-
elements such as S (Evans et al., 2010). Accessory minerals are used routinely as
Pr

chronometers (Harley and Nandakumar, 2014; Harris et al., 2000; Rubatto et al., 2009, 2013;

Sawyer, 1991; Taylor et al., 2016), particularly in migmatites and granulites, and they may
al

control the isotope composition of the melt that drains to shallower crust to form granites
n

(Ayres and Harris, 1997; Zeng et al., 2005). We divide this section into those accessory
ur

minerals we can model well, those we would like to model better, and those we cannot
Jo

(currently) consider. We also include some applications and insights regarding the behaviour

of accessory minerals in migmatites and granulites.

10.2. Minerals we can model reasonably well

Many of the accessory minerals that contain Fe and Ti as essential structural

constituents can be modelled reasonably well in anatectic systems (Diener and Powell, 2010;

White et al., 2000, 2007). In general, this includes the common oxide accessory minerals
Journal Pre-proof

(rutile, ilmenite, magnetite, spinel, hematite) and some silicates (titanite). Some accessory

minerals may be treated as pure phases lacking solid solution, such as rutile and titanite. In

these cases, our confidence in modelling their behaviour is related to our understanding of the

underlying thermodynamic properties of these minerals (Holland and Powell, 1998b, 2011;

Powell et al., 2014) and our ability to model other phases that incorporate significant amounts

of the essential structural constituents they contain, such as Ti and Fe3+ (Boger et al., 2012;

Diener and Powell, 2012; Wheller and Powell, 2014) For accessory minerals that exhibit

solid solution (e.g. spinel group and ilmenite), our confidence in terms of modelling is based

f
oo
on the underlying activity–composition models (Diener and Powell, 2010; Powell et al.,

2014), which are limited by the restricted number of experiments examining the distribution
pr
of ferric and ferrous iron between coexisting rock-forming and accessory minerals.
e-
Pr

10.3. Minerals we can model, but would like to do better


al

The accessory minerals zircon, monazite and apatite have been the focus of both
n

experimental melting studies (Boehnke et al., 2013; Harrison and Watson, 1983; Klimm et
ur

al., 2008; Stepanov et al., 2012; Watson and Harrison, 1983, 1984) and empirical studies
Jo

(Bingen et al., 1996; Cottle et al., 2018; Glasson et al., 2019; Kunz et al., 2018; Rubatto et al.,

2006; Williams, 2001), as they are important for both geochronology (Prent et al., 2020;

Rubatto, 2017; Taylor et al., 2016; Weinberg et al., 2020) and thermometry (Ferry and

Watson, 2007). Consequently, we have a reasonable understanding of the behaviour of these

minerals in simple experimental systems, although various complications arise in natural

anatectic systems that limit our ability to model quantitatively their growth and breakdown.

Assuming equilibrium among melt and minerals, the behaviour of zircon, monazite

and apatite is mostly controlled by the solubility limits of their essential structural
Journal Pre-proof

constituents in melt. Phase equilibrium modelling coupled with experimentally-derived

solubility models predicts that zircon, monazite and apatite should dissolve during heating

and melting to saturate the increasing fraction of melt in their essential structural constituents.

An example of a solubility equation for zircon (Gervasoni et al., 2016) is:

lnZrsat = (4.29 ± 0.34) – (1.35 ± 0.10) x lnG + (0.0056 ± 0.0002) x T(°C), (30)

where G is a compositional parameter for the melt (= 3 x (Al2O3 + SiO2)/(Na2O + K2O + CaO

f
oo
+ MgO + FeO)) in molar proportions and with 1σ uncertainties, and Zrsat is the concentration

of Zr in zircon-saturated melt in µg/g. Although the formulations of solubility expressions


pr
vary, most have temperature and melt composition as variables. The amount of zircon
e-
dissolution (Zrdis%) in a rock required to saturate the melt can be modelled through mass
Pr

balance considering the weight fraction of melt (F) and the total concentration of Zr (Zrsystem)

within the system. In a simple model, Zr is assumed to reside only in zircon and melt (Kelsey
al

et al., 2008), for which Zrdis% = F(Zrsat/ Zrsystem). A similar approach can be applied to
n

monazite for the LREE, which are essential structural constituents in that mineral, assuming
ur

that phosphorous is also available.


Jo

Examples of the modelled behaviour of zircon and monazite during melting of an

average amphibolite-facies metapelite (Ague, 1991) and an average passive margin

greywacke (Yakymchuk and Brown, 2014a) are shown in Fig. 24 for a closed system

scenario, considering bulk rock compositions with 150 ppm Zr and 150 ppm LREE, and that

the only repository for these elements are zircon (Zr) and monazite (LREE). For both rock

types, the amount of zircon (Fig. 24a, c) and monazite (Fig. 24b, d) decreases with increasing

temperature in a non-linear manner until each is completely dissolved between 800 and

900ºC. At no point does the proportion of either mineral in the system increase with heating
Journal Pre-proof

(Fig. 24), such that prograde zircon and monazite growth at suprasolidus conditions is not

expected in a closed system. However, the corollary is that the predicted proportion of both

minerals increases as T decreases, such that growth is expected to occur during cooling and

melt crystallization. Zircon and monazite are modelled to be stable to higher temperatures for

the greywacke than the metapelite for the same bulk rock concentrations of Zr and LREE.

This difference reflects the lower fertility of the greywacke compared with the metapelite.

For higher concentrations of Zr and LREE in the modelled system, exhaustion of zircon and

monazite is expected to occur at even higher temperatures. For example, zircon is expected to

f
oo
persist to >940 ºC for bulk rock Zr concentrations of 500 ppm in the metapelite compared

with 840–880 ºC with 300 ppm Zr (Fig. 24a).


pr
The examples above demonstrate that new growth of accessory minerals in anatectic
e-
systems is only expected during cooling and melt crystallization when the melt becomes
Pr

supersaturated with respect to those elements that are essential structural constituents of the

accessory minerals considered (Kelsey et al., 2008; Kelsey and Powell, 2011; Wang et al.,
al

2014; Yakymchuk, 2017b; Yakymchuk and Brown, 2014a; Yakymchuk et al., 2017).
n

However, this sequence of dissolution and growth contrasts with some interpretations that
ur

suggest prograde suprasolidus growth of accessory minerals can occur by Ostwald ripening
Jo

(Kawakami et al., 2013; Nemchin et al., 2001; Vavra et al., 1999). Another mechanism to

recrystallize zircon without necessarily increasing its mode is melt-induced dissolution–

reprecipitation (Hay et al., 2010; Rubatto and Hermann, 2003). Notwithstanding, mass

balance calculations support the inference that most suprasolidus zircon, monazite and apatite

growth should occur during cooling and melt crystallization (Kohn et al., 2015).

Consider a mass balance approach for zircon and its essential structural constituent,

Zr, in an anatectic system that maintains equilibrium. Zircon solubility in melt is related to

the amount of Zr that can be accommodated by the melt, which is a function of T, P and melt
Journal Pre-proof

composition (Boehnke et al., 2013; Watson and Harrison, 1983). However, other factors will

influence the behaviour of zircon, such the amount of Zr in the system (Kelsey et al., 2008;

Yakymchuk and Brown, 2014a). Another factor that can affect zircon solubility is the

presence and behaviour of other rock-forming or accessory minerals that can accommodate

Zr. For example, garnet and hornblende are both relatively Zr-rich rock-forming minerals

(Fraser et al., 1997; Kohn et al., 2015) so that the growth of peritectic garnet or hornblende

during heating and partial melting will accommodate some Zr, limiting the amount of Zr

available to the melt and enhancing dissolution of zircon. For suprasolidus prograde growth,

f
oo
breakdown of a Zr-rich mineral such as ilmenite could liberate enough Zr to create a local

volume of melt that is saturated with respect to zircon (Bingen et al., 2001; Glasson et al.,
pr
2019). However, with further heating and an increase in melt volume this zircon may be
e-
consumed. New studies directed at identifying potential sources of Zr that may be responsible
Pr

for prograde suprasolidus zircon growth would be useful in this context. A final consideration

is that zircon (and other accessory minerals) may be shielded from dissolution in stable or
al

growing peritectic minerals (e.g. garnet) during melting. These grains are not expected to
n

participate in equilibration with melt and other minerals, and are more likely to record
ur

evidence of the prograde subsolidus history.


Jo

Monazite solubility in melt is also a function of T, P and melt composition (Stepanov

et al., 2012), but its behaviour is complicated by competition with apatite for phosphorous in

anatectic systems (Duc-Tin and Keppler, 2015; Johnson et al., 2015; Schwindinger et al.,

2020; Yakymchuk, 2017b). Similar to zircon, the mode of monazite in an anatectic system is

expected to decrease during melting as the melt saturates in the LREE (Kelsey et al., 2008).

One exception may be dissolution of LREE-rich apatite during heating, in which diffusivity

of LREE away from the apatite–melt interface may be much slower than for P, resulting in a

local monazite growth (García-Arias et al., 2012; Wolf and London, 1995). However, as with
Journal Pre-proof

zircon, further heating may lead to dissolution of this monazite as diffusion rates and melt

volumes increase. The main limiting factor in our ability to model monazite is the behaviour

of Th in anatectic systems. Thorium substitution into monazite (as the huttonite or cheralite

end member) is expected to increase the stability of monazite allowing it to persist to UHT

conditions (Weller et al., 2020; Williams et al., 2018). With respect to calculated phase

equilibria, we cannot currently integrate Th in monazite or melt, which is a crucial next step

in modelling the behaviour of this important accessory mineral chronometer in formerly

suprasolidus rocks.

f
oo
Trace element concentrations in zircon and monazite are linked to the behaviour of

rock-forming minerals (petrochronology; section 9.4). An example of forward modelling


pr
applied to petrochronology for an average amphibolite-facies metapelite (Ague, 1991) is
e-
shown in Fig. 25. Accessory mineral compositional proxies for garnet behaviour are the
Pr

concentration of HREE in zircon and monazite, and of Y in monazite (Taylor et al., 2016).

The chondrite-normalized Gd/Yb (GdN/YbN) ratio of zircon is commonly used to infer


al

whether zircon grew or equilibrated in the presence of garnet, because Yb is preferentially


n

incorporated into garnet (Rubatto, 2002). In general, low GdN/YbN ratios characterise zircon
ur

in equilibrium with relatively low P and low T mineral assemblages, where garnet
Jo

proportions are low or where garnet is absent (Fig. 25a). The amount of garnet in the residue

has a strong influence on the slope of chondrite-normalized HREE concentrations in zircon.

A high proportion of garnet is consistent with relatively flat HREE patterns whereas low

proportions are indicated by positive slopes (Fig. 25d). Concentrations of Y in monazite are

also commonly linked to growth in the presence or absence of garnet, because garnet is a key

repository of Y in metamorphic rocks (Williams et al., 2007) and concentrations of Y in

monazite (calculated as wt% Y2O3 in Fig. 25b) are inversely proportional to the mode of

garnet (Fig. 25b, c). Similar to zircon, monazite in equilibrium with a garnet-rich residue has
Journal Pre-proof

low concentrations of HREE, resulting in chondrite-normalized REE patterns with steep

negative slopes (Fig. 25e).

Relating the concentrations of trace elements in accessory minerals to a metamorphic

P–T path requires knowledge of when that accessory mineral grew. Although the behaviour

of accessory minerals during anatexis is complex, during heating along an isobaric

suprasolidus path at 0.6 GPa for an average amphibolite-facies metapelite and an undrained

system, as shown in Fig. 25, the proportion of monazite and zircon are expected to decrease

(Fig. 25c). Along this path, the proportion of garnet in the system is low at the solidus, but

f
oo
increases with heating to ~900ºC when its abundance remains constant (Fig. 25c). Most

zircon and monazite growth is expected to occur during cooling and melt crystallization from
pr
around 880–850 ºC until the solidus is crossed, during which the amount of garnet in the
e-
system decreases to ~1 wt% (Fig. 25c). In Fig. 25, the first zircon and monazite to crystallize
Pr

from melt at ~880–850 ºC are expected to have the lowest concentrations of the HREE (Fig.

25d, e). Monazite is also expected to have the lowest concentration of Y at ~850 ºC, which
al

reflects the high modal abundance of garnet in the system. Monazite with high concentrations
n

of Y, and zircon and monazite with relatively high concentrations of the HREE (and low
ur

GdN/YbN ratios), record crystallization near the solidus. Alternatively, high-Y monazite and
Jo

low-HREE monazite and zircon could record prograde or retrograde subsolidus monazite or

zircon growth.

There are three key assumptions in phase equilibrium modelling when applied to

accessory minerals and petrochronology. First, we assume that the trace elements have

equilibrated between the rock-forming and accessory minerals, which is difficult to reconcile

with the limited diffusivity of the HREE in zircon, monazite (Cherniak, 2010; Cherniak et al.,

1997) and silicate liquid (Holycross et al., 2018; Mungall et al., 1999). Second, we assume

there is no temperature dependence of partitioning of REE (and Y) in garnet, zircon,


Journal Pre-proof

monazite and melt, which is a simplification (Hacker et al., 2019; Taylor et al., 2017).

Finally, we assume that zircon and monazite grow and dissolve instantaneous in melt based

on mass balance consideration, which is unlikely to be true in nature. Notwithstanding,

linking the results of phase equilibrium modelling with partitioning expressions between

rock-forming and accessory minerals provides a framework in which to interpret the

geological significance of petrochronological data from zircon and monazite.

The behaviour of apatite is anatectic systems is expected to be similar to zircon and

monazite whereby, on heating, the fraction of melt increases and the amount of apatite

f
oo
decreases until the melt becomes saturated in phosphorous (Watson and Harrison, 1984).

There are several experimentally-determined solubility expressions for apatite (Bea et al.,
pr
1992; Harrison and Watson, 1984; Pichavant et al., 1992; Tollari et al., 2006; Wolf and
e-
London, 1994), but these do not reproduce well the measured concentrations of P in natural
Pr

glasses and nanogranites (Yakymchuk and Acosta-Vigil, 2019). Nonetheless, apatite

solubility in melt increases with aluminium saturation index and temperature. A complication
al

in modelling apatite behaviour is that calcium is an essential structural constituent of many


n

rock-forming minerals, limiting the availability of Ca. New insights into apatite behaviour
ur

awaits advances in the in situ analysis of Nd isotope ratios (Fisher et al., 2020; Hammerli et
Jo

al., 2019).

Xenotime is a Y-bearing phosphate mineral that is normally restricted to subsolidus

metamorphic conditions, although it does occur in some migmatites (Engi, 2017; Pyle et al.,

2001; Spear and Pyle, 2002). It has been used as a thermometer when combined with the

compositions of monazite and garnet (Pyle et al., 2001) and can be used as a U–Pb

chronometer (Cottle et al., 2018; Parrish, 1990; Shrestha et al., 2019). Xenotime is expected

to be consumed during heating and, above the solidus, should saturate the melt in in P (Wolf

and London, 1995), although the solubility of P in xenotime-saturated melt is expected to be


Journal Pre-proof

quite low. In addition, peritectic garnet grown during incongruent melting will incorporate Y

and HREE possibly at the expense of xenotime. This is supported by the presence of

xenotime as inclusions in garnet but its absence from the matrix of high-grade rocks (Pyle

and Spear, 1999). Spear and Pyle (2010) used phase equilibrium modelling to show that

xenotime is expected to be completely consumed just above the solidus, although the specific

P–T conditions of xenotime exhaustion are sensitive to the bulk rock concentration of Y.

Xenotime should grow during melt crystallization (Crowley et al., 2009; Pyle and Spear,

1999). In general, we have a reasonable understanding of xenotime behaviour in migmatites

f
oo
and a rudimentary ability to model its behaviour. Further studies should consider the role of

xenotime in partitioning HREE during relatively low-temperature anatexis.


pr
e-
10.4. Minerals we cannot (yet) model
Pr

Allanite is a LREE-rich epidote-group mineral and a key repository for the LREE and

Th in high-pressure intermediate to basic metamorphic rocks (Engi, 2017; Hermann, 2002).


al

Although it is more common in low- to medium-temperature metamorphic rocks, allanite has


n

been documented in upper-amphibolite facies Ca-rich rocks (Bingen et al., 1996; Pan and
ur

Fleet, 1996) including migmatites (Gregory et al., 2012). Experimental studies have
Jo

demonstrated that allanite solubility in melt is a function of temperature, melt composition,

pressure and LREE concentrations (Klimm et al., 2008). Similar to other accessory minerals,

allanite is expected to break down during partial melting. Berger et al. (2008) suggested that

epidote in leucosomes formed by partial resorption of earlier-formed epidote then growth

during subsequent melt crystallization, consistent with the inference that allanite should break

down during melting and grow during melt crystallization. One complication in

understanding the behaviour of allanite is the role of fluorine and complex solid solution

between trace elements and major elements in epidote-group minerals (Gieré and Sorensen,
Journal Pre-proof

2004). Nonetheless, furthering our understanding of allanite in high-pressure rocks is crucial

for studies into the role of melt and mobilization of Th and LREE in subduction systems

(Hermann et al., 2006). Finally, allanite has potential as a U–Pb chronometer (Gregory et al.,

2012; Smye et al., 2014), although we need a better understanding of its stability and

behaviour in migmatites to be able to interpret these data.

10.5 Implications for isotope geochemistry

Isotope ratios of trace elements are commonly used to link granites to their sources at

f
oo
deeper crustal levels. In particular, neodymium and hafnium isotope ratios are used to infer if

the source rocks of granites were radiogenic (i.e. juvenile) or non-radiogenic (i.e. evolved),
pr
and to model the duration of crustal residence (Gardiner et al., 2019a, b, 2020; Vervoort and
e-
Kemp, 2016). Accessory minerals are the dominant hosts of REE and Hf in high-grade
Pr

metamorphic rocks (Bea and Montero, 1999; Hermann, 2002) and can exert a first order

influence on the isotopic composition of drained melt (Ayres and Harris, 1997; Zeng et al.,
al

2005).
n

Zircon is the dominant host of Hf in many anatectic systems and Lu is concentrated in


ur

phosphate minerals (e.g. apatite and monazite) as well as garnet. Therefore, the behaviour of
Jo

these minerals during partial melting will affect the Lu/Hf ratio of the melt and the Hf isotope

ratio of zircon that crystallizes from it. In a closed system undergoing melting, Hf liberated

from dissolved zircon may be incorporated into newly grown zircon during crystallization,

and this neoblastic zircon will have an Hf isotope composition similar to that of the older

zircon (Flowerdew et al., 2006; Li et al., 2018; Wu et al., 2007). In metasedimentary

migmatites, this implies that the Hf isotope composition of the melt (and zircon that

crystallizes from it) may reflect an isotopically-diverse population of detrital zircon. The

breakdown of Lu-rich minerals (e.g. garnet and apatite) may also release relatively radiogenic
Journal Pre-proof

Hf into the melt that could be incorporated into new zircon that will have a more radiogenic

(more juvenile) isotope ratio (Chen et al., 2015). Similar to Hafnium, Nd isotope ratios in the

melt are controlled by the behaviour of monazite and apatite (Ayres and Harris, 1997;

Hammerli et al., 2018; Hammerli et al., 2014; Korhonen et al., 2010b; Macera et al., 2011;

Villaros et al., 2009; Watt and Harley, 1993; Zeng et al., 2005). Equilibration between these

minerals and melt is an important consideration in linking the Nd isotope ratio of granites to

their sources due to slow dissolution of these two minerals in melt (Rapp and Watson, 1986).

f
oo
11. Summary and future research
pr
e-
In this review, we hope to have provided information to help both novices and
Pr

professionals in their studies of the fundamental process of irreversible differentiation of the

continental crust by partial melting. This process requires the production, segregation and
al

migration of silica-rich (felsic) melts from deeper source rocks to shallower sinks where they
n

erupt or, more commonly, crystallise as granite sensu lato. Evidence of partial melting and
ur

drainage of melt from crustal rocks is preserved in outcrops and thin sections of migmatite
Jo

and residual granulite. At the crustal scale, these processes are implicit in the complementary

compositions of residual lower crust and enriched upper crust.

Investigation of partial melting and open versus closed system processes utilises

calculated phase equilibria, which may now be applied to a wide variety of crustal source

rocks, including basalts and gabbros, clastic sedimentary rocks such as greywackes, siltstones

and mudstones (pelites), and granites themselves. Leucosomes and pegmatites in ultrahigh-

pressure metamorphic rocks have become a focus of recent studies, but rather than being

formed by aqueous fluid-present or fluid-absent hydrate-breakdown melting, they may have


Journal Pre-proof

crystallised from hydrous melt that evolved from a supercritical fluid. We also introduced the

reader to recent studies that have advanced our understanding of the behaviour of trace

elements and the role of accessory minerals during partial melting, including the implications

for isotope geochemistry.

To conclude, we offer some directions in which future studies might be focussed.

 In relation to natural examples, it would be helpful to have a better understanding of the

mechanisms by which melt moves from source to sink in various tectonic settings. As a first

f
oo
step, more quantitative studies at thin section, outcrop and crustal scale are required. In

pr
particular, we need quantitative assessments of the proportion of melt that reaches the surface

versus the proportion that crystallises at depth. This information will permit a more complete
e-
understanding of crustal behaviour through 2D and 3D numerical modelling of geodynamics
Pr

through Earth history.


al

 Along with the mechanisms, future studies should focus on the timescales of melt
n

generation, segregation, migration, emplacement and crystallization. We should combine


ur

high-precision geochronology (e.g. CA–ID–TIMS) with field observations and samples from
Jo

various crustal depths. This approach will yield a better understanding of the geological

significance of ages determined using less precise geochronometers from both source rocks

(migmatites and residual granulites) and granites. In addition, recent developments in

reaction cell technology mean that we will soon be able to undertake in situ Lu–Hf dating of

garnet (and other minerals) by LA–ICP–MS spot analysis. The method will provide a means

to constrain the metamorphic (and anatectic) history of rocks that may lack traditional

chronometers, including garnet-bearing mafic and ultramafic rocks, in a manner that is far
Journal Pre-proof

more time and cost efficient than is currently available. A better determination of timescales

will improve our understanding of anatectic systems as a whole.

 With respect to calculated phase equilibria, there is a need for more targeted experiments

in simple (synthetic) compositional systems to enable the energetic contributions of

individual components in solid solutions to be better understood, which will allow a–X

(solution) models to be refined. This is particularly important for the ferric iron component in

ferromagnesian minerals, where our understanding of a–X relationships is hindered by the

f
oo
paucity of experimental data. In addition, we need continued improvements to the a–X

pr
thermodynamic models used in phase equilibrium calculations, including extending the range

of compositions to which they may be applied and the P–T conditions for which they may be
e-
used. To supplement new data from targeted experimental studies, we should increase the use
Pr

of comprehensive compositional data (i.e. major and trace elements) for coexisting minerals

in equilibrated natural samples.


n al

 To improve our understanding of open versus closed system melting, we should critically
ur

assess the role of H2O-fluxed melting in producing and reworking continental crust. In
Jo

particular, we should better characterise the sources of H2O, the solute content of the

attending fluid, and the thermochemical consequences of pervasive fluid infiltration during

high-grade metamorphism, as well as developing a better understanding of the mechanics of

fluid infiltration. This is of particular importance for arc magmatism on the modern Earth and

potentially for TTG production in the early Earth.

 The role of supercritical fluid during exhumation of deeply-subducted continental crust is

underappreciated in the UHP metamorphic community. As a first step, an assessment of the


Journal Pre-proof

potential role of dehydroxylation and the formation of a supercritical fluid in the generation

of hydrous melt during exhumation of all major UHP terranes would be useful.

 The importance of trace element partitioning and the role of accessory minerals in the

petrogenesis of granites have been appreciated for several decades. However, it would be

useful to have a better understanding of the behaviour of accessory minerals in anatectic

systems. This requires additional empirical and experimental studies on the partitioning of

elements between the accessory minerals, rock-forming minerals and melt, in particular for

f
oo
those elements that form essential structural constituents of the accessory minerals. This is

pr
crucial if we are to better assess the geological significance of the precise ages now being

retrieved from accessory minerals. A wider range of partitioning data will allow better links
e-
to be made between the growth or recrystallisation of accessory minerals and estimated P–T
Pr

conditions along a metamorphic P–T path. Furthermore, we recommend increasing the use of

rapid mapping of trace elements by LA–ICP–MS as a complement to major element mapping


al

by EPMA to better link accessory mineral behaviour with that of the rock-forming minerals.
n

This would also enable a better understanding of trace element partitioning between the rock-
ur

forming and accessory minerals with melt in anatectic systems, which is important for
Jo

furthering our understanding of long-term crustal differentiation using elemental and isotopic

data compiled in large datasets. Two additional goals are the identification of potential

sources of Zr that may support prograde suprasolidus growth of zircon, and the integration of

Th into monazite solubility models to refine our understanding of monazite ages retrieved

from formerly suprasolidus crustal rocks. Lastly, for both rocks and accessory minerals, we

should determine the relationship between changes in elemental or isotopic composition and

tectonic setting in both modern and ancient environments.


Journal Pre-proof

To close, we identify two topics that should be addressed in future studies. First, we

should develop a better understanding of the role of bolide impacts on large-scale crustal

melting and differentiation, particularly to assess the contribution of these processes to the

evolution of the mantle, crust, hydrosphere, atmosphere and biosphere on Earth and other

terrestrial bodies. Second, we should assess the physicochemical drivers behind the formation

of rare metal deposits from granite magmas and pegmatites. Our enhanced understanding of

the processes involved in the concentration of these elements in evolved melts is crucial for

identifying new deposits of these raw materials, since they are becoming increasingly

f
oo
important to current and future technological advances, particularly in green-energy

technologies.
pr
e-
Author contributions
Pr

Tim Johnson: Conceptualization; Visualization; Writing – sections of original draft; Writing

– review & editing. Chris Yakymchuk: Visualization; Writing – sections of original draft;
al

Writing – review & editing. Michael Brown: Conceptualization; Writing – sections of


n

original draft; Writing – review & editing.


ur
Jo

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

Over several decades we have collectively benefited from countless discussions with

colleagues too numerous to mention individually—we are grateful to you all. Tim Horscroft

is thanked for the invitation to write this review. We are particularly indebted to Antoine
Journal Pre-proof

Godet and Roberto Weinberg for their perceptive, detailed and helpful reviews, and for their

resilience in getting through the entire manuscript. Douwe van Hinsbergen is acknowledged

for his editorial handling and generosity in extending deadlines. Any remaining errors or

misconceptions are entirely ours.

Appendix

Glossary of terms relating to partial melting of crustal rocks

Many of these terms (and others) are also explained in the Atlas of Migmatites (Sawyer,

f
oo
2008).

 Anatexis—see partial melting,


pr
 Congruent melting: partial melting in which one or more reactants (minerals with or
e-
without volatile species) break down to form a melt of the same composition as the only
Pr

product.

 Diatexite—see migmatite.
al

 Dihedral angle: the angle subtended at the junctions of two solid grains in textural
n

equilibrium with a fluid or melt phase is the fluid–solid–solid dihedral angle.


ur

 Exhumation: the exposure of a rock unit on the land surface by removal of the
Jo

overburden at the metamorphic peak (not to be confused with uplift, which is the upward

movement of a rock unit relative to adjacent units).

 Fertility: the property of some protolith compositions to produce more melt than others at

the same P–T conditions.

 Fluid (also Volatile): a low-viscosity liquid or gas phase (commonly H2O) that may form

or be present in both subsolidus and suprasolidus rocks.

 Fusion: see partial melting.

 H2O-saturated solidus: see solidus.


Journal Pre-proof

 H2O-undersaturated solidus: see solidus.

 Heat: thermal energy within, or transferred to or from, a thermodynamic system.

 Heat producing element: an unstable atom that produces heat by radioactive decay, most

importantly isotopes of K, Th and U.

 Hydrous mineral (also Hydrate): a mineral containing OH– or H2O in its crystal

structure (e.g. mica, amphibole, epidote, cordierite, etc.); hence hydrate-breakdown

melting is incongruent melting involving the breakdown of one (or more) hydrous

f
minerals.

oo
 Incongruent melting: partial melting in which both liquid and one or more peritectic

pr
minerals are the products of the melting reaction.

 In-situ leucosome—see Leucosome


e-
 In-source leucosome—see Leucosome
Pr

 Internal energy: the total energy a substance possesses.

 Latent heat of fusion: at fixed pressure, the amount of heat required to convert one unit
al

of a substance from the solid to the liquid phase without changing the temperature of the
n
ur

system.

 Leucosome: the lighter-coloured part of the neosome that represents material crystallized
Jo

mostly from the segregated melt; it is commonly dominated by quartz and feldspar, and it

may or may not correspond to a melt composition. In-situ leucosome is the product of

crystallization of anatectic melt, or part of the anatectic melt, that has segregated from its

residuum, but has remained where it formed, whereas in-source leucosome is the product

of crystallisation of an anatectic melt, or part of an anatectic melt, that has migrated away

from the place where it formed, but is still located within its source layers.

 Liquidus: a line in P–T space above which a substance is completely liquid.


Journal Pre-proof

 Melanosome: the darker coloured part of the neosome that represents residual material

that did not melt and from which melt has drained.

 Melt: A viscous polymerised hydrous silicate liquid that begins to form when a rock

exceeds the temperatures of its solidus.

 Melt connectivity transition (MCT): the melt fraction (around 7 vol.% melt) above

which the melt forms an interconnected network and the strength of a partially-molten

rock drops by an order of magnitude or more.

 Metamorphic peak: the maximum temperature (or less commonly pressure) recorded by

f
oo
a metamorphic rock.

 Metamorphic facies: commonly-occurring sets of mineral assemblages developed in


pr
rocks of different compositions that recur in space and time.
e-
 Metatexite—see Migmatite
Pr

 Microstructure: the relative size, shape and spatial arrangement of grains that comprise a

rock (NB. texture is reserved for the internal features of the individual grains).
al

 Migmatite: a metamorphic rock formed by anatexis that is generally heterogeneous and


n

preserves evidence of partial melting at the microscopic to macroscopic scale


ur

(Yakymchuk, 2020). The first-order classification of migmatites is based on the inferred


Jo

melt fraction in the rock (Sawyer, 2008). Metatexite migmatites are characterized by a

low melt fraction, which allows pre-anatectic structures generally to be preserved, whereas

diatexite migmatites are characterized by high melt fraction, which generally means that

pre-anatectic structures have been disrupted by flow. The second-order classification of

migmatites relates to the amount of syn-anatectic strain. Patch, net-structured and

stromatic metatexite migmatites represent increasing syn-anatectic strain at low melt

fraction, whereas raft or schollen and schlieren diatexite migmatites represent


Journal Pre-proof

increasing syn-anatectic strain at high melt fraction. The interested reader is referred to

Sawyer (2008) for further details.

 Mineral assemblage: a collection of locally coexisting minerals interpreted to have

chemically interacted; the term equilibrium mineral assemblage may be used when a

state of chemical equilibrium is implied.

 Neosome: the part of a migmatite that was newly formed or reconstituted by partial

melting.

 Paleosome: the part of a migmatite that was unaffected by partial melting due to protolith

f
oo
composition or isolation from a hydrous fluid that fluxed melting.

 Partial melting (also Anatexis and Fusion): Partial melting is the process or action of
pr
transforming a substance into a liquid plus residual solids usually, but not exclusively, by
e-
the addition of heat.
Pr

 Peritectic mineral(s): the solid product(s) of an incongruent melting reaction.

 Prograde metamorphism: the changes in mineral assemblage and microstructure that


al

occur during heating.


n

 Protolith: the original crustal rock before modification by partial melting and loss of melt.
ur

 Reaction: the process by which one or more phases, the reactants, are converted to one or
Jo

more different phases, the products.

 Reaction rate: the rate of change of the abundance (mode) of reactants or products as a

function of changing T and/or P.

 Residue or Residuum (adj. residual): material from which partial melt has drained

leaving behind melt-depleted solids.

 Retrograde metamorphism: the changes in mineral assemblage and microstructure that

occur during cooling.


Journal Pre-proof

 Solidus: a line in P–T space at temperatures below which melt is not stable; this line

separates subsolidus from suprasolidus P–T conditions. The H2O-saturated (wet)

solidus is the locus of P–T conditions at which, with increasing T, H2O-saturated melt first

forms in equilibrium with a discrete H2O fluid phase. The H2O-undersaturated solidus is

the locus of P–T conditions at which, with increasing T, an H2O-undersaturated

composition will begin to melt.

 Ultrahigh pressure metamorphism: metamorphism that results in the presence of a high-

f
pressure polymorph of silica (coesite) or carbon (diamond) in a rock; the presence of

oo
either mineral, whatever its grain size or abundance, indicates a minimum pressure on the

order of 3 or 4 GPa, respectively.


pr
 Ultrahigh temperature metamorphism: metamorphism in which peak temperature
e-
exceeded 900°C at a pressure below the sillimanite–kyanite transition.
Pr

Declaration of interests
The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.
n al

References
ur

Acosta-Vigil, A., London, D., Morgan, G.B.V., Cesare, B., Buick, I., Hermann, J. and
Jo

Bartoli, O., 2017. Primary crustal melt compositions: Insights into the controls,

mechanisms and timing of generation from kinetics experiments and melt inclusions.

Lithos, 286-287: 454-479.

Ague, J.J., 1991. Evidence for major mass transfer and volume strain during regional

metamorphism of pelites. Geology, 19(8): 855-858.

Albarède, F., 1998. The growth of continental crust. Tectonophysics, 296(1-2): 1-14.

Alessio, K.L., Hand, M., Morrissey, L.J., Kelsey, D.E. and Payne, J.L., 2017. Melt

reintegration modelling: Testing against a subsolidus reference assemblage.

Geosciences (Switzerland), 7(3).


Journal Pre-proof

Anderson, J., Kelsey, D., Hand, M. and Collins, W., 2013. Conductively driven, high‐thermal

gradient metamorphism in the Anmatjira Range, Arunta region, central Australia.

Journal of Metamorphic Geology, 31(9): 1003-1026.

Annen, C., Blundy, J. and Sparks, R., 2006. The genesis of intermediate and silicic magmas

in deep crustal hot zones. Journal of Petrology, 47(3): 505-539.

Annen, C. and Sparks, R.S.J., 2002. Effects of repetitive emplacement of basaltic intrusions

on thermal evolution and melt generation in the crust. Earth and Planetary Science

Letters, 203(3-4): 937-955.

f
oo
Arndt, N.T., 2013. Formation and evolution of the continental crust. Geochemical

Perspectives, 2(3): I-VI+405-533.


pr
Arzi, A.A., 1978. Critical phenomena in the rheology of partially melted rocks.
e-
Tectonophysics, 44(1-4): 173-184.
Pr

Ashworth, J.R., 1985. Migmatites. Migmatites.

Ashworth, J.R. and Brown, M., 1990. High-temperature metamorphism and crustal anatexis,
al

2. Kluer Academic Publishers.


n

Asimow, P.D. and Ghiorso, M.S., 1998. Algorithmic modifications extending MELTS to
ur

calculate subsolidus phase relations. American Mineralogist, 83(9-10): 1127-1132.


Jo

Asimow, P.D., Hirschmann, M. and Stolper, E., 1997. An analysis of variations in isentropic

melt productivity. Philosophical Transactions of the Royal Society of London. Series

A: Mathematical, Physical and Engineering Sciences, 355(1723): 255-281.

Audétat, A. and Keppler, H., 2004. Viscosity of fluids in subduction zones. Science,

303(5657): 513-516.

Auzanneau, E., Vielzeuf, D. and Schmidt, M.W., 2006. Experimental evidence of

decompression melting during exhumation of subducted continental crust.

Contributions to Mineralogy and Petrology, 152(2): 125-148.


Journal Pre-proof

Ayres, M. and Harris, N., 1997. REE fractionation and Nd-isotope disequilibrium during

crustal anatexis: Constraints from Himalayan leucogranites. Chemical Geology,

139(1-4): 249-269.

Barbarin, B., 1996. Genesis of the two main types of peraluminous granitoids. Geology,

24(4): 295-298.

Barker, F. and Arth, J.G., 1976. Generation of trondhjemitic-tonalitic liquids and Archean

bimodal trondhjemite-basalt suites. Geology, 4(10): 596-600.

Barrow, G., 1893. On an Intrusion of Muscovite-biotite Gneiss in the South-eastern

f
oo
Highlands of Scotland, and its accompanying Metamorphism. Quarterly Journal of

the Geological Society, 49(1-4): 330-358.


pr
Barrow, G., 1912. On the geology of Lower Dee-side and the southern Highland Border.
e-
Proceedings of the Geologists' Association, 23(5): 274-IN1.
Pr

Bartoli, O., 2017. Phase equilibria modelling of residual migmatites and granulites: An

evaluation of the melt-reintegration approach. Journal of Metamorphic Geology,


al

35(8): 919-942.
n

Bartoli, O., 2019. Reintegrating nanogranitoid inclusion composition to reconstruct the


ur

prograde history of melt-depleted rocks. Geoscience Frontiers, 10(2): 517-525.


Jo

Bartoli, O., 2021. Granite geochemistry is not diagnostic of the role of water in the source.

Earth and Planetary Science Letters, 564:116927.

Bartoli, O., Acosta-Vigil, A., Ferrero, S. and Cesare, B., 2016. Granitoid magmas preserved

as melt inclusions in high-grade metamorphic rocks. American Mineralogist, 101(7):

1543-1559.

Baxter, E., Caddick, M. and Dragovic, B., 2017. Garnet: A rock-forming mineral

petrochronometer. Reviews in Mineralogy and Geochemistry, 83(1): 469-533.


Journal Pre-proof

Bartoli, O. and Carvalho, B.B., 2021. Anatectic granites in their source region: A comparison

between experiments, thermodynamic modelling and nanogranitoids. Lithos, 106046.

Bea, F., 1991. Geochemical modeling of low melt-fraction anatexis in a peraluminous

system: The Pena Negra Complex (central Spain). Geochimica et Cosmochimica

Acta, 55(7): 1859-1874.

Bea, F., 2012. The sources of energy for crustal melting and the geochemistry of heat-

producing elements. Lithos, 153: 278-291.

Bea, F., Fershtater, G. and Corretgé, L., 1992. The geochemistry of phosphorus in granite

f
oo
rocks and the effect of aluminium. Lithos, 29(1-2): 43-56.

Bea, F. and Montero, P., 1999. Behavior of accessory phases and redistribution of Zr, REE,
pr
Y, Th, and U during metamorphism and partial melting of metapelites in the lower
e-
crust: an example from the Kinzigite Formation of Ivrea-Verbano, NW Italy.
Pr

Geochimica et Cosmochimica Acta, 63(7-8): 1133-1153.

Bea, F., Pereira, M.D. and Stroh, A., 1994. Mineral/leucosome trace-element partitioning in a
al

peraluminous migmatite (a laser ablation-ICP-MS study). Chemical Geology, 117(1-


n

4): 291-312.
ur

Beard, J.S. and Lofgren, G.E., 1989. Effect of water on the composition of partial melts of
Jo

greenstone and amphibolite. Science, 244(4901): 195-197.

Beard, J.S. and Lofgren, G.E., 1991. Dehydration melting and water-saturated melting of

basaltic and andesitic greenstones and amphibolites at 1, 3, and 6.9 kb. Journal of

Petrology, 32(2): 365-401.

Bédard, J., 2007. Trace element partitioning coefficients between silicate melts and

orthopyroxene: parameterizations of D variations. Chemical Geology, 244(1-2): 263-

303.
Journal Pre-proof

Bédard, J.H., 2006a. A catalytic delamination-driven model for coupled genesis of Archaean

crust and sub-continental lithospheric mantle. Geochimica et Cosmochimica Acta,

70(5): 1188-1214.

Bédard, J.H., 2006b. Trace element partitioning in plagioclase feldspar. Geochimica et

Cosmochimica Acta, 70(14): 3717-3742.

Bédard, J.H., 2018. Stagnant lids and mantle overturns: Implications for Archaean tectonics,

magmagenesis, crustal growth, mantle evolution, and the start of plate tectonics.

Geoscience Frontiers, 9(1): 19-49.

f
oo
Berger, A., Burri, T., Alt-Epping, P. and Engi, M., 2008. Tectonically controlled fluid flow

and water-assisted melting in the middle crust: An example from the Central Alps.

Lithos, 102(3-4): 598-615.


pr
e-
Berman, R.G., 1988. Internally-consistent thermodynamic data for minerals in the system
Pr

Na2O-K2O-CaO-MgO-FeO-Fe2O3-Al2O3-SiO2-TiO2-H2O-CO2. Journal of Petrology,

29(2): 445-522.
al

Berry, A.J., Danyushevsky, L.V., St C. O’Neill, H., Newville, M. and Sutton, S.R., 2008.
n

Oxidation state of iron in komatiitic melt inclusions indicates hot Archaean mantle.
ur

Nature, 455(7215): 960-963.


Jo

Bhattacharya, A., Mohanty, L., Maji, A., Sen, S. and Raith, M., 1992. Non-ideal mixing in

the phlogopite-annite binary: constraints from experimental data on Mg−Fe

partitioning and a reformulation of the biotite-garnet geothermometer. Contributions

to Mineralogy and Petrology, 111(1): 87-93.

Bingen, B., Austrheim, H. and Whitehouse, M., 2001. Ilmenite as a source for zirconium

during high-grade metamorphism? Textural evidence from the Caledonides of

western Norway and implications for zircon geochronology. Journal of Petrology,

42(2): 355-375.
Journal Pre-proof

Bingen, B., Demaiffe, D. and Hertogen, J., 1996. Redistribution of rare earth elements,

thorium, and uranium over accessory minerals in the course of amphibolite to

granulite facies metamorphism: the role of apatite and monazite in orthogneisses from

southwestern Norway. Geochimica et Cosmochimica Acta, 60(8): 1341-1354.

Bloch, E., Ganguly, J., Hervig, R. and Cheng, W., 2015. 176 Lu–176 Hf geochronology of

garnet I: experimental determination of the diffusion kinetics of Lu 3+ and Hf 4+ in

garnet, closure temperatures and geochronological implications. Contributions to

Mineralogy and Petrology, 169(2): 12.

f
oo
Boehnke, P., Watson, E.B., Trail, D., Harrison, T.M. and Schmitt, A.K., 2013. Zircon

saturation re-revisited. Chemical Geology, 351: 324-334.


pr
Boettcher, A. and Wyllie, P., 1968. Melting of granite with excess water to 30 kilobars
e-
pressure. The Journal of Geology, 76(2): 235-244.
Pr

Boger, S.D., White, R.W. and Schulte, B., 2012. The importance of iron speciation

(Fe+2/Fe+3) in determining mineral assemblages: An example from the high-grade


al

aluminous metapelites of southeastern Madagascar. Journal of Metamorphic Geology,


n

30(9): 997-1018.
ur

Bonamici, C.E. and Duebendorfer, E.M., 2010. Scale-invariance and self-organized criticality
Jo

in migmatites of the southern Hualapai Mountains, Arizona. Journal of Structural

Geology, 32(8): 1114-1124.

Bonnefoi, C.C., Provost, A. and Albarede, F., 1995. The 'Daly gap' as a magmatic

catastrophe. Nature, 378(6554): 270-272.

Bons, P.D., Arnold, J., Elburg, M.A., Kalda, J., Soesoo, A. and van Milligen, B.P., 2004.

Melt extraction and accumulation from partially molten rocks. Lithos, 78(1-2): 25-42.
Journal Pre-proof

Bons, P.D., Druguet, E., Castaño, L.-M. and Elburg, M.A., 2008. Finding what is now not

there anymore: Recognizing missing fluid and magma volumes. Geology, 36(11):

851-854.

Borghini, A., Ferrero, S., O’Brien, P.J., Laurent, O., Günter, C. and Ziemann, M.A., 2020.

Cryptic metasomatic agent measured in situ in Variscan mantle rocks: Melt inclusions

in garnet of eclogite, Granulitgebirge, Germany. Journal of Metamorphic Geology,

38(3): 207-234.

Brown, C.R., Yakymchuk, C., Brown, M., Fanning, C.M., Korhonen, F.J., Piccoli, P.M. and

f
oo
Siddoway, C.S., 2016. From source to sink: Petrogenesis of cretaceous anatectic

granites from the Fosdick migmatite-granite complex, West Antarctica. Journal of

Petrology, 57(7): 1241-1278.


pr
e-
Brown, M., 2002. Retrograde processes in migmatites and granulites revisited. Journal of
Pr

Metamorphic Geology, 20(1): 25-40.

Brown, M., 2003. Hot orogens, tectonic switching, and creation of continental crust:
al

Comment. Geology, 31(1).


n

Brown, M., 2004. The mechanism of melt extraction from lower continental crust of orogens,
ur

Special Paper of the Geological Society of America, pp. 35-48.


Jo

Brown, M., 2005. Synergistic effects of melting and deformation: An example from the

Variscan belt, western France, Geological Society Special Publication, pp. 205-226.

Brown, M., 2007. Metamorphic conditions in orogenic belts: A record of secular change.

International Geology Review, 49(3): 193-234.

Brown, M., 2010a. Melting of the continental crust during orogenesis: the thermal,

rheological, and compositional consequences of melt transport from lower to upper

continental crust. Canadian Journal of Earth Sciences, 47(5): 655-694.

Brown, M., 2010b. Paired metamorphic belts revisited. Gondwana Research, 18(1): 46-59.
Journal Pre-proof

Brown, M., 2010c. The spatial and temporal patterning of the deep crust and implications for

the process of melt extraction. Philosophical Transactions of the Royal Society A:

Mathematical, Physical and Engineering Sciences, 368(1910): 11-51.

Brown, M., 2013. Granite: From genesis to emplacement. GSA bulletin, 125(7-8): 1079-

1113.

Brown, M., Averkin, Y.A., McLellan, E.L. and Sawyer, E.W., 1995. Melt segregation in

migmatites. Journal of Geophysical Research, 100(B8): 15,655-15,679.

Brown, M. and Johnson, T., 2018. Secular change in metamorphism and the onset of global

f
oo
plate tectonics. American Mineralogist, 103(2): 181-196.

Brown, M. and Johnson, T., 2019a. Global age, temperature and pressure data for secular
pr
change in metamorphism. EarthChem Library, DOI, 10.
e-
Brown, M. and Johnson, T., 2019b. Metamorphism and the evolution of subduction on Earth.
Pr

American Mineralogist, 104(8): 1065-1082.

Brown, M. and Johnson, T., 2019c. Time's arrow, time's cycle: Granulite metamorphism and
al

geodynamics. Mineralogical Magazine, 83(3): 323-338.


n

Brown, M. and Rushmer, T., 2006a. Evolution and differentiation of the continental crust.
ur

Cambridge University Press.


Jo

Brown, M. and Rushmer, T., 2006b. Introduction to Evolution and Differentiation of the

Continental Crust, Evolution and Differentiation of the Continental Crust. Cambridge

University Press, pp. 1-23.

Brown, M. and Solar, G.S., 1998. Granite ascent and emplacement during contractional

deformation in convergent orogens. Journal of Structural Geology, 20(9-10): 1365-

1393.
Journal Pre-proof

Brown, M. and Solar, G.S., 1999. The mechanism of ascent and emplacement of granite

magma during transpression: A syntectonic granite paradigm. Tectonophysics,

312(1): 1-33.

Brown, M.A., Brown, M., Carlson, W.D. and Denison, C., 1999. Topology of syntectonic

melt-flow networks in the deep crust: Inferences from three-dimensional images of

leucosome geometry in migmatites. American Mineralogist, 84(11-12): 1793-1818.

Bryden, C.D. and Jamieson, R.A., 2020. Scapolite pegmatite from the Nordøyane ultra-high

pressure domain, Western Gneiss Region, Norway: Partial melting driven by

f
oo
infiltration of mantle-derived fluid. Lithos, 364-365.

Burnham, C.W., 1979. Magmas and hydrothermal fluids. Geochemistry of Hydrothermal Ore

Deposits.: 71-136.
pr
e-
Campbell, I.H. and Taylor, S.R., 1983. No water, no granites‐No oceans, no continents.
Pr

Geophysical Research Letters, 10(11): 1061-1064.

Capitani, C.d. and Petrakakis, K., 2010. The computation of equilibrium assemblage
al

diagrams with Theriak/Domino software. American Mineralogist, 95(7): 1006-1016.


n

Carlson, W.D., Pattison, D.R. and Caddick, M.J., 2015. Beyond the equilibrium paradigm:
ur

How consideration of kinetics enhances metamorphic interpretation. American


Jo

Mineralogist, 100(8-9): 1659-1667.

Carmichael, D.M., 1969. On the mechanism of prograde metamorphic reactions in quartz-

bearing pelitic rocks. Contributions to Mineralogy and Petrology, 20(3): 244-267.

Carnot, S., 1824. Reflections on the motive power of fire, and on machines fitted to develop

that power. Paris: Bachelier.

Carrington, D.P. and Harley, S.L., 1995. Partial melting and phase relations in high-grade

metapelites: an experimental petrogenetic grid in the KFMASH system. Contributions

to Mineralogy and Petrology, 120(3-4): 270-291.


Journal Pre-proof

Carvalho, B., Sawyer, E. and Janasi, V., 2016. Crustal reworking in a shear zone:

transformation of metagranite to migmatite. Journal of Metamorphic Geology, 34(3):

237-264.

Castillo, P.R., 2006. An overview of adakite petrogenesis. Chinese Science Bulletin, 51(3):

257-268.

Castro, A., 2013. Tonalite–granodiorite suites as cotectic systems: a review of experimental

studies with applications to granitoid petrogenesis. Earth-Science Reviews, 124: 68-

95.

f
oo
Cesare, B., Acosta-Vigil, A., Bartoli, O. and Ferrero, S., 2015. What can we learn from melt

inclusions in migmatites and granulites? Lithos, 239: 186-216.


pr
Cesare, B., Acosta-Vigil, A., Ferrero, S. and Bartoli, O., 2011. Melt inclusions in migmatites
e-
and granulites. Journal of the Virtual Explorer, 38.
Pr

Cesare, B., Marchesi, C., Hermann, J. and Gómez-Pugnaire, M.T., 2003. Primary melt

inclusions in andalusite from anatectic graphitic metapelites: Implications for the


al

position of the Al2SiO5 triple point. Geology, 31(7): 573-576.


n

Chappell, B., 1974. Two contrasting granite types. Pacif. Geol., 8: 173-174.
ur

Chappell, B., 1999. Aluminium saturation in I-and S-type granites and the characterization of
Jo

fractionated haplogranites. Lithos, 46(3): 535-551.

Chappell, B. and White, A., 1992. I-and S-type granites in the Lachlan Fold Belt.

Transactions of the Royal Society of Edinburgh: Earth Sciences, 83(1-2): 1-26.

Chappell, B.W. and White, A.J., 2001. Two contrasting granite types: 25 years later.

Australian journal of earth sciences, 48(4): 489-499.

Chen, R.-X., Zheng, Y.-F., Gong, B., Zhao, Z.-F., Gao, T.-S., Chen, B. and Wu, Y.-B., 2007.

Origin of retrograde fluid in ultrahigh-pressure metamorphic rocks: constraints from


Journal Pre-proof

mineral hydrogen isotope and water content changes in eclogite–gneiss transitions in

the Sulu orogen. Geochimica et Cosmochimica Acta, 71(9): 2299-2325.

Chen, Y.X., Gao, P. and Zheng, Y.F., 2015. The anatectic effect on the zircon Hf isotope

composition of migmatites and associated granites. Lithos, 238: 174-184.

Cheng, H., Vervoort, J.D., Dragovic, B., Wilford, D. and Zhang, L., 2018. Coupled Lu–Hf

and Sm–Nd geochronology on a single eclogitic garnet from the Huwan shear zone,

China. Chemical Geology, 476: 208-222.

Cherniak, D., 2010. Diffusion in accessory minerals: zircon, titanite, apatite, monazite and

f
oo
xenotime. Reviews in mineralogy and geochemistry, 72(1): 827-869.

Cherniak, D., Hanchar, J. and Watson, E., 1997. Rare-earth diffusion in zircon. Chemical

Geology, 134(4): 289-301.


pr
e-
Chinner, G.A., 1966. The distribution of pressure and temperature during Dalradian
Pr

metamorphism. Quarterly Journal of the Geological Society of London, 122(1-4):

159-186.
al

Chowdhury, P., Chakraborty, S., Gerya, T.V., Cawood, P.A. and Capitanio, F.A., 2020. Peel-
n

back controlled lithospheric convergence explains the secular transitions in Archean


ur

metamorphism and magmatism. Earth and Planetary Science Letters, 538.


Jo

Cipar, J.H., Garber, J.M., Kylander-Clark, A.R. and Smye, A.J., 2020. Active crustal

differentiation beneath the Rio Grande Rift. Nature Geoscience: 1-6.

Clark, C., Fitzsimons, I.C., Healy, D. and Harley, S.L., 2011. How does the continental crust

get really hot? Elements, 7(4): 235-240.

Clarke, D.B., 1992. Granitoid rocks, 7. Springer Science & Business Media.

Clarke, D.B., 2019. The origins of strongly peraluminous granitoid rocks. Canadian

Mineralogist, 57(4): 529-550.


Journal Pre-proof

Clarke, G., Daczko, N. and Nockolds, C., 2001. A method for applying matrix corrections to

X‐ray intensity maps using the Bence–Albee algorithm and Matlab. Journal of

metamorphic Geology, 19(6): 635-644.

Clemens, J., Buick, I. and Stevens, G., 2016. Fluids, melting, granulites and granites: A

controversy–reply to the commentary of. Precambrian Research, 278: 400-404.

Clemens, J., Holloway, J.R. and White, A., 1986. Origin of an A-type granite; experimental

constraints. American Mineralogist, 71(3-4): 317-324.

Clemens, J., Stevens, G. and Mayne, M., 2021. Do arc silicic magmas form by fluid-fluxed

f
oo
melting of older arc crust or fractionation of basaltic magmas? Contributions to

Mineralogy and Petrology, 176(6): 1-28.


pr
Clemens, J.D., 2003. S-type granitic magmas - petrogenetic issues, models and evidence.
e-
Earth-Science Reviews, 61(1-2): 1-18.
Pr

Clemens, J.D., 2006. Melting of the continental crust: fluid regimes, melting reactions and

source-rock fertility. In: M. Brown and T. Rushmer (Editors), Evolution and


al

Differentiation of the Continental Crust. Cambridge University Press, Cambridge, pp.


n

296–331.
ur

Clemens, J.D. and Stevens, G., 2016. Melt segregation and magma interactions during crustal
Jo

melting: breaking out of the matrix. Earth-Science Reviews, 160: 333-349.

Clemens, J.D., Stevens, G. and Bryan, S.E., 2020. Conditions during the formation of granitic

magmas by crustal melting – Hot or cold; drenched, damp or dry? Earth-Science

Reviews, 200.

Clemens, J.D., Stevens, G. and Farina, F., 2011. The enigmatic sources of I-type granites:

The peritectic connexion. Lithos, 126(3-4): 174-181.

Clemens, J.D. and Wall, V.J., 1981. Origin and crystallization of some peraluminous (S-type)

granitic magmas. The Canadian Mineralogist, 19(1): 111-131.


Journal Pre-proof

Clos, F., Weinberg, R.F., Zibra, I. and Schwindinger, M., 2019. Magmatic and anatectic

history of a large Archean diapir: Insights from the migmatitic core of the Yalgoo

Dome, Yilgarn Craton. Lithos, 338: 18-33.

Coleman, D.S., Gray, W. and Glazner, A.F., 2004. Rethinking the emplacement and

evolution of zoned plutons: Geochronologic evidence for incremental assembly of the

Tuolumne Intrusive Suite, California. Geology, 32(5): 433-436.

Collins, W., Beams, S., White, A. and Chappell, B., 1982. Nature and origin of A-type

granites with particular reference to southeastern Australia. Contributions to

f
oo
mineralogy and petrology, 80(2): 189-200.

Collins, W. and Richards, S., 2008. Geodynamic significance of S-type granites in circum-

Pacific orogens. Geology, 36(7): 559-562.


pr
e-
Collins, W.J., 1996. Lachlan Fold Belt granitoids: Products of three-component mixing,
Pr

Special Paper of the Geological Society of America, pp. 171-181.

Collins, W.J., Huang, H.-Q., Bowden, P. and Kemp, A., 2020a. Repeated S–I–A-type granite
al

trilogy in the Lachlan Orogen and geochemical contrasts with A-type granites in
n

Nigeria: implications for petrogenesis and tectonic discrimination. Geological society,


ur

london, special publications, 491(1): 53-76.


Jo

Collins, W.J., Huang, H.-Q. and Jiang, X., 2016. Water-fluxed crustal melting produces

Cordilleran batholiths. Geology, 44(2): 143-146.

Collins, W.J., Murphy, J.B., Johnson, T.E. and Huang, H.-Q., 2020b. Critical role of water in

the formation of continental crust. Nature Geoscience: 1-8.

Condie, K.C., 1981. Archean greenstone belts. Elsevier.

Condie, K.C., 1993. Chemical composition and evolution of the upper continental crust:

Contrasting results from surface samples and shales. Chemical Geology, 104(1-4): 1-

37.
Journal Pre-proof

Condie, K.C., 2014. How to make a continent: Thirty-five years of TTG research, Modern

Approaches in Solid Earth Sciences. Springer International Publishing, pp. 179-193.

Connolly, J., 1990. Multivariable phase diagrams; an algorithm based on generalized

thermodynamics. American Journal of Science, 290(6): 666-718.

Connolly, J. and Petrini, K., 2002. An automated strategy for calculation of phase diagram

sections and retrieval of rock properties as a function of physical conditions. Journal

of Metamorphic Geology, 20(7): 697-708.

Connolly, J.A.D., 2005. Computation of phase equilibria by linear programming: A tool for

f
oo
geodynamic modeling and its application to subduction zone decarbonation. Earth and

Planetary Science Letters, 236(1-2): 524-541.


pr
Conrad, W.K., Nicholls, I.A. and Wall, V.J., 1988. Water-saturated and -undersaturated
e-
melting of metaluminous and peraluminous crustal compositions at 10 kb: Evidence
Pr

for the origin of silicic magmas in the taupo volcanic zone, new zealand, and other

occurrences. Journal of Petrology, 29(4): 765-803.


al

Cottle, J.M., Larson, K.P. and Yakymchuk, C., 2018. Contrasting accessory mineral behavior
n

in minimum-temperature melts: empirical constraints from the Himalayan


ur

metamorphic core. Lithos, 312: 57-71.


Jo

Cox, K., Bell, J. and Pankhurst, R., 1979. Pankhurst RJ (1979) The interpretation of igneous

rocks. George Allen, 86.

Crowley, J.L., Waters, D., Searle, M. and Bowring, S., 2009. Pleistocene melting and rapid

exhumation of the Nanga Parbat massif, Pakistan: Age and P–T conditions of

accessory mineral growth in migmatite and leucogranite. Earth and Planetary Science

Letters, 288(3-4): 408-420.


Journal Pre-proof

Cruden, A., 2006. Emplacement and growth of plutons: implications for rates of melting and

mass transfer in continental crust, Evolution and differentiation of the continental

crust. Cambridge University Press, pp. 455-519.

Cruden, A. and McCaffrey, K., 2001. Growth of plutons by floor subsidence: implications for

rates of emplacement, intrusion spacing and melt-extraction mechanisms. Physics and

Chemistry of the Earth, Part A: Solid Earth and Geodesy, 26(4-5): 303-315.

Daly, R.A., 1925. The geology of Ascension island, Proceedings of the American Academy

of Arts and Sciences. JSTOR, pp. 3-80.

f
oo
Davidson, J., Turner, S., Handley, H., Macpherson, C. and Dosseto, A., 2007. Amphibole

“sponge” in arc crust? Geology, 35(9): 787-790.


pr
de Capitani, C. and Petrakakis, K., 2010. The computation of equilibrium assemblage
e-
diagrams with Theriak/Domino software. American Mineralogist, 95(7): 1006-1016.
Pr

Defant, M.J. and Drummond, M.S., 1990. Derivation of some modern arc magmas by

melting of young subducted lithosphere. Nature, 347(6294): 662-665.


al

Deniel, C., Vidal, P., Fernandez, A., Le Fort, P. and Peucat, J.-J., 1987. Isotopic study of the
n

Manaslu granite (Himalaya, Nepal): inferences on the age and source of Himalayan
ur

leucogranites. Contributions to Mineralogy and Petrology, 96(1): 78-92.


Jo

Diener, J., White, R. and Powell, R., 2008. Granulite facies metamorphism and subsolidus

fluid‐absent reworking, Strangways Range, Arunta Block, central Australia. Journal

of Metamorphic Geology, 26(6): 603-622.

Diener, J.F.A. and Fagereng, A., 2014. The influence ofmelting and melt drainage on crustal

rheology during orogenesis. Journal of Geophysical Research: Solid Earth, 119(8):

6193-6210.

Diener, J.F.A. and Powell, R., 2010. Influence of ferric iron on the stability of mineral

assemblages. Journal of Metamorphic Geology, 28(6): 599-613.


Journal Pre-proof

Diener, J.F.A. and Powell, R., 2012. Revised activity-composition models for clinopyroxene

and amphibole. Journal of Metamorphic Geology, 30(2): 131-142.

Doukkari, S.A., Diener, J.F., Ouzegane, K. and Kienast, J.R., 2018. Mineral equilibrium

modelling and calculated chemical potential relations of reaction textures in the

ultrahigh‐temperature In Ouzzal terrane (In Hihaou area, Western Hoggar, Algeria).

Journal of Metamorphic Geology, 36(9): 1175-1198.

Dragovic, B., Baxter, E.F. and Caddick, M.J., 2015. Pulsed dehydration and garnet growth

during subduction revealed by zoned garnet geochronology and thermodynamic

f
oo
modeling, Sifnos, Greece. Earth and Planetary Science Letters, 413: 111-122.

Droop, G., 1987. A general equation for estimating Fe 3+ concentrations in ferromagnesian


pr
silicates and oxides from microprobe analyses, using stoichiometric criteria.
e-
Mineralogical magazine, 51(361): 431-435.
Pr

Droop, G. and Harte, B., 1995. The effect of Mn on the phase relations of medium-grade

pelites: constraints from natural assemblages on petrogenetic grid topology. Journal of


al

Petrology, 36(6): 1549-1578.


n

Droop, G.T.R., Clemens, J.D. and Dalrymple, D.J., 2003. Processes and conditions during
ur

contact anatexis, melt escape and restite formation: The Huntly Gabbro complex, NE
Jo

Scotland. Journal of Petrology, 44(6): 995-1029.

Duc-Tin, Q. and Keppler, H., 2015. Monazite and xenotime solubility in granitic melts and

the origin of the lanthanide tetrad effect. Contributions to Mineralogy and Petrology,

169(1): 8.

Ducea, M.N., Ganguly, J., Rosenberg, E.J., Patchett, P.J., Cheng, W. and Isachsen, C., 2003.

Sm-Nd dating of spatially controlled domains of garnet single crystals: A new method

of high-temperature thermochronology. Earth and Planetary Science Letters, 213(1-

2): 31-42.
Journal Pre-proof

Duesterhoeft, E. and de Capitani, C., 2013. THERIAK_D: An add‐on to implement

equilibrium computations in geodynamic models. Geochemistry, Geophysics,

Geosystems, 14(11): 4962-4967.

Dumond, G., Goncalves, P., Williams, M.L. and Jercinovic, M.J., 2015. Monazite as a

monitor of melting, garnet growth and feldspar recrystallization in continental lower

crust. Journal of Metamorphic Geology, 33(7): 735-762.

Eby, G.N., 1992. Chemical subdivision of the A-type granitoids: petrogenetic and tectonic

implications. Geology, 20(7): 641-644.

f
oo
Ellis, D. and Green, D., 1979. An experimental study of the effect of Ca upon garnet-

clinopyroxene Fe-Mg exchange equilibria. Contributions to Mineralogy and

Petrology, 71(1): 13-22.


pr
e-
Engi, M., 2017. Petrochronology based on REE-minerals: monazite, allanite, xenotime,
Pr

apatite. Reviews in Mineralogy and Geochemistry, 83(1): 365-418.

Eskola, P.E., 1920. The mineral facies of rocks.


al

Etheridge, M.A., Daczko, N.R., Chapman, T. and Stuart, C.A., 2020. Mechanisms of melt
n

extraction during lower crustal partial melting. Journal of Metamorphic Geology.


ur

Evans, K.A., Powell, R. and Holland, T.J.B., 2010. Internally consistent data for sulphur-
Jo

bearing phases and application to the construction of pseudosections for mafic

greenschist facies rocks in Na2O-CaO-K2O-FeO-MgO-Al2O3-SiO2-CO2-O-S-H2O.

Journal of Metamorphic Geology, 28(6): 667-687.

Evans, T.P., 2004. A method for calculating effective bulk composition modification due to

crystal fractionation in garnet-bearing schist: Implications for isopleth

thermobarometry. Journal of Metamorphic Geology, 22(6): 547-557.


Journal Pre-proof

Feisel, Y., White, R.W., Palin, R.M. and Johnson, T.E., 2018. New constraints on granulite

facies metamorphism and melt production in the Lewisian Complex, northwest

Scotland. Journal of Metamorphic Geology, 36(6): 799-819.

Feng, P., Wang, L., Brown, M., Johnson, T.E., Kylander-Clark, A. and Piccoli, P.M., 2021.

Partial melting of ultrahigh-pressure eclogite by omphacite-breakdown facilitates

exhumation of deeply-subducted crust. Earth and Planetary Science Letters.

Ferrero, S., O’Brien, P.J., Borghini, A., Wunder, B., Wälle, M., Günter, C. and Ziemann,

M.A., 2019. A treasure chest full of nanogranitoids: an archive to investigate crustal

f
oo
melting in the Bohemian Massif. Geological Society, London, Special Publications,

478(1): 13-38.
pr
Ferrero, S., Wunder, B., Walczak, K., O’Brien, P.J. and Ziemann, M.A., 2015. Preserved
e-
near ultrahigh-pressure melt from continental crust subducted to mantle depths.
Pr

Geology, 43(5): 447-450.

Ferrero, S., Wunder, B., Ziemann, M.A., Wälle, M. and O'Brien, P.J., 2016. Carbonatitic and
al

granitic melts produced under conditions of primary immiscibility during anatexis in


n

the lower crust. Earth and Planetary Science Letters, 454: 121-131.
ur

Ferry, J. and Watson, E., 2007. New thermodynamic models and revised calibrations for the
Jo

Ti-in-zircon and Zr-in-rutile thermometers. Contributions to Mineralogy and

Petrology, 154(4): 429-437.

Finger, F. and Clemens, J.D., 1995. Migmatization and "secondary" granitic magmas: effects

of emplacement and crystallization of "primary" granitoids in Southern Bohemia,

Austria. Contributions to Mineralogy and Petrology, 120(3-4): 311-326.

Fischer, S., Prave, A.R., Johnson, T.E., Cawood, P.A., Hawkesworht, C.J., Horstwood,

M.S.A. and EIMF, 2021. Using zircon in mafic migmatites to disentangle complex
Journal Pre-proof

high-grade gneiss terrains – Terrane spotting in the Lewisian complex, NW Scotland.

Precambrian Research, XX: xxx–xxx.

Fisher, C.M., Bauer, A.M., Luo, Y., Sarkar, C., Hanchar, J.M., Vervoort, J.D., Tapster, S.R.,

Horstwood, M. and Pearson, D.G., 2020. Laser ablation split-stream analysis of the

Sm-Nd and U-Pb isotope compositions of monazite, titanite, and apatite–

Improvements, potential reference materials, and application to the Archean Saglek

Block gneisses. Chemical Geology, 539: 119493.

Flament, N., Coltice, N. and Rey, P.F., 2008. A case for late-Archaean continental emergence

f
oo
from thermal evolution models and hypsometry. Earth and Planetary Science Letters,

275(3-4): 326-336.
pr
Flowerdew, M.J., Millar, I.L. and Vaughan, A.P.M., 2006. Potential of combined U-Pb
e-
geochronology and Hf isotope geochemical studies on zircon to aid sedimentary
Pr

provenance within Antarctica: Examples from West Antarctica. Terra Antartica

Reports(12): 57-64.
al

Flowers, R.M., Schmitt, A.K. and Grove, M., 2010. Decoupling of U–Pb dates from chemical
n

and crystallographic domains in granulite facies zircon. Chemical Geology, 270(1-4):


ur

20-30.
Jo

Foden, J., Sossi, P.A. and Wawryk, C.M., 2015. Fe isotopes and the contrasting petrogenesis

of A-, I-and S-type granite. Lithos, 212: 32-44.

Foley, S., 2008. A trace element perspective on Archean crust formation and on the presence

or absence of Archean subduction, Special Paper of the Geological Society of

America. Geological Society of America, pp. 31-50.

Foley, S.F., Buhre, S. and Jacob, D.E., 2003. Evolution of the Archaean crust by

delamination and shallow subduction. Nature, 421(6920): 249-252.


Journal Pre-proof

Fraser, G., Ellis, D. and Eggins, S., 1997. Zirconium abundance in granulite-facies minerals,

with implications for zircon geochronology in high-grade rocks. Geology, 25(7): 607-

610.

Frezzotti, M.L. and Ferrando, S., 2015. The chemical behavior of fluids released during deep

subduction based on fluid inclusions. American Mineralogist, 100(2-3): 352-377.

Frost, B.R., Barnes, C.G., Collins, W.J., Arculus, R.J., Ellis, D.J. and Frost, C.D., 2001. A

geochemical classification for granitic rocks. Journal of petrology, 42(11): 2033-

2048.

f
oo
Frost, B.R., Mavrogenes, J.A. and Tomkins, A.G., 2002. Partial melting of sulfide ore

deposits during medium-and high-grade metamorphism. The Canadian Mineralogist,

40(1): 1-18.
pr
e-
Fyfe, W.S., Turner, F.J. and Verhoogen, J., 1958. Metamorphic reactions and metamorphic
Pr

facies. Geological Society of America Memoirs, 73: 1-251.

Gallien, F., Mogessie, A., Hauzenberger, C., Bjerg, E., Delpino, S. and Castro de Machuca,
al

B., 2012. On the origin of multi‐layer coronas between olivine and plagioclase at the
n

gabbro–granulite transition, Valle Fértil–La Huerta Ranges, San Juan Province,


ur

Argentina. Journal of Metamorphic Geology, 30(3): 281-302.


Jo

Ganguly, J., 2008. Thermodynamics in earth and planetary sciences. Springer.

Ganzhorn, A.C., Labrousse, L., Prouteau, G., Leroy, C., Vrijmoed, J.C., Andersen, T.B. and

Arbaret, L., 2014. Structural, petrological and chemical analysis of syn-kinematic

migmatites: Insights from the Western Gneiss Region, Norway. Journal of

Metamorphic Geology, 32(6): 647-673.

Gao, L.E., Zeng, L. and Asimow, P.D., 2017a. Contrasting geochemical signatures of fluid-

absent versus fluid-fluxed melting of muscovite in metasedimentary sources: The

Himalayan leucogranites. Geology, 45(1): 39-42.


Journal Pre-proof

Gao, P., Zheng, Y.-F. and Zhao, Z.-F., 2016a. Distinction between S-type and peraluminous

I-type granites: Zircon versus whole-rock geochemistry. Lithos, 258: 77-91.

Gao, P., Zheng, Y.F. and Zhao, Z.F., 2016b. Experimental melts from crustal rocks: a

lithochemical constraint on granite petrogenesis. Lithos, 266: 133-157.

Gao, X.Y., Chen, Y.X. and Zhang, Q.Q., 2017b. Multiphase solid inclusions in ultrahigh-

pressure metamorphic rocks: A snapshot of anatectic melts during continental

collision. Journal of Asian Earth Sciences, 145: 192-204.

García-Arias, M., Corretgé, L.G. and Castro, A., 2012. Trace element behavior during partial

f
oo
melting of Iberian orthogneisses: An experimental study. Chemical Geology, 292: 1-

17.
pr
Gardien, V., Thompson, A.B., Grujic, D. and Ulmer, P., 1995. Experimental melting of
e-
biotite + plagioclase + quartz ± muscovite assemblages and implications for crustal
Pr

melting. Journal of Geophysical Research, 100(B8): 15,581-15,591.

Gardiner, N.J., Johnson, T.E., Kirkland, C.L. and Smithies, R.H., 2018. Melting controls on
al

the lutetium–hafnium evolution of Archaean crust. Precambrian Research, 305: 479-


n

488.
ur

Gardiner, N.J., Johnson, T.E., Kirkland, C.L. and Szilas, K., 2019a. Modelling the Hafnium-
Jo

Neodymium Evolution of Early Earth: A Study from West Greenland. Journal of

Petrology, 60(1): 177-197.

Gardiner, N.J., Kirkland, C.L., Hollis, J., Szilas, K., Steenfelt, A., Yakymchuk, C. and Heide-

Jørgensen, H., 2019b. Building Mesoarchaean crust upon Eoarchaean roots: the Akia

Terrane, West Greenland. Contributions to Mineralogy and Petrology, 174(3).

Gardiner, N.J., Kirkland, C.L., Hollis, J.A., Cawood, P.A., Nebel, O., Szilas, K. and

Yakymchuk, C., 2020. North Atlantic Craton architecture revealed by kimberlite-

hosted crustal zircons. Earth and Planetary Science Letters, 534: 116091.
Journal Pre-proof

Gervais, F. and Trapy, P.-H., 2021. Testing solution models for phase equilibrium (forward)

modeling of partial melting experiments. Contributions to Mineralogy and Petrology,

176(1): 1-18.

Gervasoni, F., Klemme, S., Rocha-Júnior, E.R. and Berndt, J., 2016. Zircon saturation in

silicate melts: a new and improved model for aluminous and alkaline melts.

Contributions to Mineralogy and Petrology, 171(3): 21.

Gerya, T.V. and Meilick, F., 2011. Geodynamic regimes of subduction under an active

margin: effects of rheological weakening by fluids and melts. Journal of Metamorphic

f
oo
Geology, 29(1): 7-31.

Ghent, E., 1976. Plagioclase–garnet–Al2SiO5–quartz: a potential geothermometer-

geobarometer. Am Mineral, 61: 710-714.


pr
e-
Ghiorso, M.S., Hirschmann, M.M., Reiners, P.W. and Kress, V.C., 2002. The pMELTS: A
Pr

revision of MELTS for improved calculation of phase relations and major element

partitioning related to partial melting of the mantle to 3 GPa. Geochemistry,


al

Geophysics, Geosystems, 3(5): 1-35.


n

Ghiorso, M.S. and Sack, R.O., 1995. Chemical mass transfer in magmatic processes IV. A
ur

revised and internally consistent thermodynamic model for the interpolation and
Jo

extrapolation of liquid-solid equilibria in magmatic systems at elevated temperatures

and pressures. Contributions to Mineralogy and Petrology, 119(2-3): 197-212.

Gibbs, J.W., 1879. On the equilibrium of heterogeneous substances.

Gibson, R., 2002. Impact‐induced melting of Archean granulites in the Vredefort Dome,

South Africa. I: anatexis of metapelitic granulites. Journal of Metamorphic Geology,

20(1): 57-70.

Gieré, R. and Sorensen, S.S., 2004. Allanite and other REE-rich epidote-group minerals.

Reviews in Mineralogy and Geochemistry, 56(1): 431-493.


Journal Pre-proof

Gilotti, J.A., 1993. Discovery of a medium-temperature eclogite province in the Caledonides

of North-East Greenland. Geology, 21(6): 523-526.

Glasson, K.J., Johnson, T.E., Kirkland, C.L., Gardiner, N.J., Clark, C., Blereau, E., Hartnady,

M.I.H., Spaggiari, C. and Smithies, H., 2019. A window into an ancient backarc? The

magmatic and metamorphic history of the Fraser Zone, Western Australia.

Precambrian Research, 323: 55-69.

Godet, A., Guilmette, C., Labrousse, L., Davis, D.W., Smit, M.A., Cutts, J.A., Vanier, M.A.,

Lafrance, I. and Charette, B., 2020a. Complete metamorphic cycle and long-lived

f
oo
anatexis in the c. 2.1 Ga Mistinibi Complex, Canada. Journal of Metamorphic

Geology, 38(3): 235-264.


pr
Godet, A., Guilmette, C., Labrousse, L., Smit, M.A., Davis, D.W., Raimondo, T., Vanier,
e-
M.A., Charette, B. and Lafrance, I., 2020b. Contrasting P-T-t paths reveal a
Pr

metamorphic discontinuity in the New Quebec Orogen: Insights into Paleoproterozoic

orogenic processes. Precambrian Research, 342.


al

Goldschmidt, V.M., 1954. Geochemistry, 78. LWW.


n

Gordon, S.M., Grove, M., Whitney, D.L., Schmitt, A.K. and Teyssier, C., 2009. Fluid-rock
ur

interaction in orogenic crust tracked by zircon depth profiling. Geology, 37(8): 735-
Jo

738.

Gordon, S.M., Whitney, D.L., Teyssier, C. and Fossen, H., 2013. U-Pb dates and trace-

element geochemistry of zircon from migmatite, Western Gneiss Region, Norway:

Significance for history of partial melting in continental subduction. Lithos, 170-171:

35-53.

Gottschalk, M., 1996. Internally consistent thermodynamic data for rock-forming minerals in

the system SiO2-TiO2-Al2O3-Fe2O3-CaO-MgO-FeO-K2O-Na2O-H2O-CO2. European

Journal of Mineralogy: 175-223.


Journal Pre-proof

Grant, J.A., 1985. Phase equilibria in partial melting of pelitic rocks, Migmatites. Springer,

pp. 86-144.

Grant, J.A., 2004. Liquid compositions from low-pressure experimental melting of pelitic

rock from Morton Pass, Wyoming, USA. Journal of Metamorphic Geology, 22(2):

65-78.

Green, E.C.R., White, R.W., Diener, J.F.A., Powell, R., Holland, T.J.B. and Palin, R.M.,

2016. Activity–composition relations for the calculation of partial melting equilibria

in metabasic rocks. Journal of Metamorphic Geology, 34(9): 845-869.

f
oo
Greenfield, J.E., Clarke, C.L., Bland, M. and Clark, D.J., 1996. In-situ migmatite and hybrid

diatexite at Mt Stafford, central Australia. Journal of Metamorphic Geology, 14(4):

413-426.
pr
e-
Gregory, C., Rubatto, D., Hermann, J., Berger, A. and Engi, M., 2012. Allanite behaviour
Pr

during incipient melting in the southern Central Alps. Geochimica et Cosmochimica

Acta, 84: 433-458.


al

Gualda, G.A., Bégué, F., Pamukcu, A.S. and Ghiorso, M.S., 2019. Rhyolite-MELTS vs
n

DERP—Newer Does not Make it Better: a Comment on ‘The Effect of Anorthite


ur

Content and Water on Quartz–Feldspar Cotectic Compositions in the Rhyolitic


Jo

System and Implications for Geobarometry’by Wilke et al.(2017; Journal of

Petrology, 58, 789–818). Journal of Petrology, 60(4): 855-864.

Guernina, S. and Sawyer, E.W., 2003. Large-scale melt-depletion in granulite terranes: An

example from the Archean Ashuanipi subprovince of Quebec. Journal of

Metamorphic Geology, 21(2): 181-201.

Guilmette, C., Indares, A. and Hébert, R., 2011. High-pressure anatectic paragneisses from

the Namche Barwa, Eastern Himalayan Syntaxis: Textural evidence for partial

melting, phase equilibria modeling and tectonic implications. Lithos, 124(1-2): 66-81.
Journal Pre-proof

Hack, A.C. and Thompson, A.B., 2011. Density and viscosity of hydrous magmas and related

fluids and their role in subduction zone processes. Journal of Petrology, 52(7-8):

1333-1362.

Hack, A.C., Thompson, A.B. and Aerts, M., 2007. Phase relations involving hydrous silicate

melts, aqueous fluids, and minerals. Reviews in Mineralogy and Geochemistry, 65(1):

129-185.

Hacker, B., Kylander‐Clark, A. and Holder, R., 2019. REE partitioning between monazite

and garnet: Implications for petrochronology. Journal of Metamorphic Geology,

f
oo
37(2): 227-237.

Hall, D. and Kisters, A., 2012. The stabilization of self-organised leucogranite networks-
pr
Implications for melt segregation and far-field melt transfer in the continental crust.
e-
Earth and Planetary Science Letters, 355-356: 1-12.
Pr

Hall, J., 1805. III. Experiments on whinstone and lava. Earth and Environmental Science

Transactions of the Royal Society of Edinburgh, 5(1): 43-75.


al

Hamilton, D., Burnham, C.W. and Osborn, E., 1964. The solubility of water and effects of
n

oxygen fugacity and water content on crystallization in mafic magmas. Journal of


ur

Petrology, 5(1): 21-39.


Jo

Hammerli, J., Kemp, A.I. and Whitehouse, M.J., 2019. In situ trace element and Sm-Nd

isotope analysis of accessory minerals in an Eoarchean tonalitic gneiss from

Greenland: Implications for Hf and Nd isotope decoupling in Earth's ancient rocks.

Chemical Geology, 524: 394-405.

Hammerli, J., Kemp, A.I.S., Shimura, T., Vervoort, J.D., Eimf and Dunkley, D.J., 2018.

Generation of I-type granitic rocks by melting of heterogeneous lower crust. Geology,

46(10): 907-910.
Journal Pre-proof

Hammerli, J., Kemp, A.I.S. and Spandler, C., 2014. Neodymium isotope equilibration during

crustal metamorphism revealed by in situ microanalysis of REE-rich accessory

minerals. Earth and Planetary Science Letters, 392: 133-142.

Handy, M., Mulch, A., Rosenau, M. and Rosenberg, C., 2001a. The role of fault zones and

melts as agents of weakening, hardening and differentiation of the continental crust: a

synthesis. Geological Society, London, Special Publications, 186(1): 305-332.

Handy, M.R., Braun, J., Brown, M., Kukowski, N., Paterson, M.S., Schmid, S.M., Stöckhert,

B., Stüwe, K., Thompson, A.B. and Wosnitza, E., 2001b. Rheology and geodynamic

f
oo
modelling: The next step forward. International Journal of Earth Sciences, 90(1): 149-

156.
pr
Hans Wedepohl, K., 1995. The composition of the continental crust. Geochimica et
e-
Cosmochimica Acta, 59(7): 1217-1232.
Pr

Hansen, E., Johansson, L., Andersson, J., LaBarge, L., Harlov, D., Möller, C. and Vincent,

S., 2015. Partial melting in amphibolites in a deep section of the Sveconorwegian


al

Orogen, SW Sweden. Lithos, 236-237: 27-45.


n

Hanson, G.N., 1978. The application of trace elements to the petrogenesis of igneous rocks of
ur

granitic composition. Earth and planetary science letters, 38(1): 26-43.


Jo

Harley, S.L., 2008. Refining the P-T records of UHT crustal metamorphism. Journal of

Metamorphic Geology, 26(2): 125-154.

Harley, S.L., 2016. A matter of time: the importance of the duration of UHT metamorphism.

Journal of Mineralogical and Petrological Sciences, 111(2): 50-72.

Harley, S.L. and Kelly, N.M., 2007. Zircon: Tiny but timely. Elements, 3(1): 13-18.

Harley, S.L., Kelly, N.M. and Möller, A., 2007. Zircon behaviour and the thermal histories of

moutain chains. Elements, 3(1): 25-30.


Journal Pre-proof

Harley, S.L. and Nandakumar, V., 2014. Accessory mineral behaviour in granulite

migmatites: a case study from the Kerala Khondalite Belt, India. Journal of Petrology,

55(10): 1965-2002.

Harley, S.L. and Thompson, P., 2004. The influence of cordierite on melting and mineral-

melt equilibria in ultra-high-temperature metamorphism, Special Paper of the

Geological Society of America, pp. 87-98.

Harris, N., Ayres, M. and Massey, J., 1995. Geochemistry of granitic melts produced during

the incongruent melting of muscovite: implications for the extraction of Himalayan

f
oo
leucogranite magmas. Journal of Geophysical Research: Solid Earth, 100(B8): 15767-

15777.
pr
Harris, N., Vance, D. and Ayres, M., 2000. From sediment to granite: Timescales of anatexis
e-
in the upper crust. Chemical Geology, 162(2): 155-167.
Pr

Harrison, T.M. and Watson, E.B., 1983. Kinetics of zircon dissolution and zirconium

diffusion in granitic melts of variable water content. Contributions to Mineralogy and


al

Petrology, 84(1): 66-72.


n

Harrison, T.M. and Watson, E.B., 1984. The behavior of apatite during crustal anatexis:
ur

Equilibrium and kinetic considerations. Geochimica et Cosmochimica Acta, 48(7):


Jo

1467-1477.

Harte, B. and Hudson, N.F., 1979. Pelite facies series and the temperatures and pressures of

Dalradian metamorphism in E Scotland. Geological Society, London, Special

Publications, 8(1): 323-337.

Hartel, T. and Pattison, D., 1996. Genesis of the Kapuskasing (Ontario) migmatitic mafic

granulites by dehydration melting of amphibolite: the importance of quartz to reaction

progress. Journal of Metamorphic Geology, 14(5): 591-611.


Journal Pre-proof

Hasalová, P., Janoušek, V., Schulmann, K., Štípská, P. and Erban, V., 2008a. From

orthogneiss to migmatite: Geochemical assessment of the melt infiltration model in

the Gföhl Unit (Moldanubian Zone, Bohemian Massif). Lithos, 102(3-4): 508-537.

Hasalová, P., Štípská, P., Powell, R., Schulmann, K., Janoušek, V. and Lexa, O., 2008b.

Transforming mylonitic metagranite by open-system interactions during melt flow.

Journal of Metamorphic Geology, 26(1): 55-80.

Hawkesworth, C.J. and Kemp, A., 2006. Evolution of the continental crust. Nature,

443(7113): 811-817.

f
oo
Hay, D.C., Dempster, T.J., Lee, M.R. and Brown, D.J., 2010. Anatomy of a low temperature

zircon outgrowth. Contributions to Mineralogy and Petrology, 159(1): 81-92.


pr
Hayden, L.A., Watson, E.B. and Wark, D.A., 2008. A thermobarometer for sphene (titanite).
e-
Contributions to Mineralogy and Petrology, 155(4): 529-540.
Pr

Healy, B., Collins, W. and Richards, S., 2004. A hybrid origin for Lachlan S-type granites:

the Murrumbidgee Batholith example. Lithos, 78(1-2): 197-216.


al

Helgeson, H.C., HC, H., JM, D., HW, N. and DK, B., 1978. Summary and critique of the
n

thermodynamic properties of rock-forming minerals.


ur

Hensen, B., 1971. Theoretical phase relations involving cordierite and garnet in the system
Jo

MgO-FeO-Al 2 O 3-SiO 2. Contributions to Mineralogy and Petrology, 33(3): 191-

214.

Hermann, J., 2002. Allanite: Thorium and light rare earth element carrier in subducted crust.

Chemical Geology, 192(3-4): 289-306.

Hermann, J. and Rubatto, D., 2014. Subduction of continental crust to mantle depth:

geochemistry of ultrahigh-pressure rocks, Treatise on Geochemistry, 2nd Edition.

Elsevier.
Journal Pre-proof

Hermann, J., Spandler, C., Hack, A. and Korsakov, A.V., 2006. Aqueous fluids and hydrous

melts in high-pressure and ultra-high pressure rocks: Implications for element transfer

in subduction zones. Lithos, 92(3-4): 399-417.

Hermann, J., Zheng, Y.-F. and Rubatto, D., 2013. Deep fluids in subducted continental crust.

Elements, 9(4): 281-287.

Hernández-Montenegro, D., Andronicos, C.L., Zuluaga, C.A. and Aronoff, R.F., 2019.

Effects of melt loss, melt retention, and protolith composition on differentiation of

anatectic metapelites: A case study of the Wet Mountains, Colorado. Lithos, 344-345:

f
oo
425-439.

Hernández-Uribe, D. and Palin, R.M., 2019. A revised petrological model for subducted
pr
oceanic crust: Insights from phase equilibrium modelling. Journal of Metamorphic
e-
Geology, 37(6): 745-768.
Pr

Higashino, F., Rubatto, D., Kawakami, T., Bouvier, A.S. and Baumgartner, L.P., 2019.

Oxygen isotope speedometry in granulite facies garnet recording fluid/melt–rock


al

interaction (Sør Rondane Mountains, East Antarctica). Journal of metamorphic


n

geology, 37(7): 1037-1048.


ur

Hoffmann, J.E., Münker, C., Polat, A., Rosing, M.T. and Schulz, T., 2011. The origin of
Jo

decoupled Hf-Nd isotope compositions in Eoarchean rocks from southern West

Greenland. Geochimica et Cosmochimica Acta, 75(21): 6610-6628.

Hofmann, A.W., 1988. Chemical differentiation of the Earth: the relationship between

mantle, continental crust, and oceanic crust. Earth and Planetary Science Letters,

90(3): 297-314.

Holdaway, M., 2001. Recalibration of the GASP geobarometer in light of recent garnet and

plagioclase activity models and versions of the garnet-biotite geothermometer.

American Mineralogist, 86(10): 1117-1129.


Journal Pre-proof

Holder, R., Yakymchuk, C. and Viete, D., 2020. Accessory mineral Eu anomalies in

suprasolidus rocks: Beyond feldspar. Geochemistry, Geophysics, Geosystems, 21(8):

e2020GC009052.

Holk, G.J. and Taylor, H.P., Jr., 2000. Water as a petrologic catalyst driving 18o/16o

homogenization and anatexis of the middle crust in the metamorphic core complexes

of british columbia. International Geology Review, 42(2): 97-130.

Holland, T. and Powell, R., 1998a. An internally consistent thermodynamic data set for

phases of petrological interest. Journal of metamorphic Geology, 16(3): 309-343.

f
oo
Holland, T.B., Hudson, N.F.C., Powell, R. and Harte, B., 2013. Newthermodynamic models

and calculated phase equilibria in NCFMAS for basic and ultrabasic compositions
pr
through thetransition zone into the uppermost lower mantle. Journal of Petrology,
e-
54(9): 1901-1920.
Pr

Holland, T.J.B., Green, E.C.R. and Powell, R., 2018. Melting of peridotites through to

granites: A simple thermodynamic model in the system KNCFMASHTOCr. Journal


al

of Petrology, 59(5): 881-900.


n

Holland, T.J.B. and Powell, R., 1985. An internally consistent thermodynamic dataset with
ur

uncertainties and correlations: 2. Data and results. Journal of Metamorphic Geology,


Jo

3(4): 343-370.

Holland, T.J.B. and Powell, R., 1990. An enlarged and updated internally consistent

thermodynamic dataset with uncertainties and correlations: the system K2O–Na2O–

CaO–MgO–MnO–FeO–Fe2O3–Al2O3–TiO2–SiO2–C–H2–O2. Journal of Metamorphic

Geology, 8(1): 89-124.

Holland, T.J.B. and Powell, R., 1998b. An internally consistent thermodynamic data set for

phases of petrological interest. Journal of Metamorphic Geology, 16(3): 309-343.


Journal Pre-proof

Holland, T.J.B. and Powell, R., 2011. An improved and extended internally consistent

thermodynamic dataset for phases of petrological interest, involving a new equation

of state for solids. Journal of Metamorphic Geology, 29(3): 333-383.

Holness, M., 1999. Contact metamorphism and anatexis of Torridonian arkose by minor

intrusions of the Rum Igneous Complex, Inner Hebrides, Scotland. Geological

Magazine, 136(5): 527-542.

Holness, M.B., Cesare, B. and Sawyer, E.W., 2011. Melted rocks under the microscope:

Microstructures and their interpretation. Elements, 7(4): 247-252.

f
oo
Holness, M.B. and Sawyer, E.W., 2008. On the pseudomorphing of melt-filled pores during

the crystallization of migmatites. Journal of Petrology, 49(7): 1343-1363.


pr
Holtz, F. and Johannes, W., 1991. Genesis of peraluminous granites I. Experimental
e-
investigation of melt compositions at 3 and 5 kb and various H2O activities. Journal
Pr

of Petrology, 32(5): 935-958.

Holtz, F., Johannes, W., Tamic, N. and Behrens, H., 2001. Maximum and minimum water
al

contents of granitic melts generated in the crust: a reevaluation and implications.


n

Lithos, 56(1): 1-14.


ur

Holycross, M., Watson, E., Richter, F. and Villeneuve, J., 2018. Diffusive fractionation of Li
Jo

isotopes in wet, highly silicic melts. Geochemical Perspectives Letters, 6: 39-42.

Huang, G., Palin, R., Wang, D. and Guo, J., 2020. Open-system fractional melting of

Archean basalts: implications for tonalite–trondhjemite–granodiorite (TTG) magma

genesis. Contributions to Mineralogy and Petrology, 175(11).

Huang, W.-L. and Wyllie, P.J., 1975. Melting reactions in the system NaAlSi3O8-

KAlSi3O8-SiO2 to 35 kilobars, dry and with excess water. The Journal of Geology,

83(6): 737-748.
Journal Pre-proof

Hutton, J., 1788. X. Theory of the Earth; or an Investigation of the Laws observable in the

Composition, Dissolution, and Restoration of Land upon the Globe. Earth and

Environmental Science Transactions of The Royal Society of Edinburgh, 1(2): 209-

304.

Iles, K.A., Hergt, J.M. and Woodhead, J.D., 2018. Modelling isotopic responses to

disequilibrium melting in granitic systems. Journal of Petrology, 59(1): 87-113.

Indares, A., White, R.W. and Powell, R., 2008. Phase equilibria modelling of kyanite-bearing

anatectic paragneisses from the central Grenville Province. Journal of Metamorphic

f
oo
Geology, 26(8): 815-836.

Inger, S. and Harris, N., 1993. Geochemical constraints on leucogranite magmatism in the
pr
Langtang Valley, Nepal Himalaya. Journal of Petrology, 34(2): 345-368.
e-
Iwamori, H., McKenzie, D. and Takahashi, E., 1995. Melt generation by isentropic mantle
Pr

upwelling. Earth and Planetary Science Letters, 134(3-4): 253-266.

Jacob, J.-B., Moyen, J.-F., Fiannacca, P., Laurent, O., Bachmann, O., Janoušek, V., Farina, F.
al

and Villaros, A., 2021. Crustal melting vs. fractionation of basaltic magmas: Part 2,
n

Attempting to quantify mantle and crustal contributions in granitoids. Lithos: 106292.


ur

Jagoutz, O. and Klein, B., 2018. On the importance of crystallization-differentiation for the
Jo

generation of SiO2-rich melts and the compositional build-up of arc (and continental)

crust. American Journal of Science, 318(1): 29-63.

Jagoutz, O., Müntener, O., Schmidt, M.W. and Burg, J.-P., 2011. The roles of flux- and

decompression melting and their respective fractionation lines for continental crust

formation: Evidence from the Kohistan arc. Earth and Planetary Science Letters,

303(1-2): 25-36.

Jagoutz, O. and Schmidt, M.W., 2012. The formation and bulk composition of modern

juvenile continental crust: The Kohistan arc. Chemical Geology, 298: 79-96.
Journal Pre-proof

Jamieson, R.A., Beaumont, C., Fullsack, P. and Lee, B., 1998. Barrovian regional

metamorphism: Where’s the heat? Geological Society, London, Special Publications,

138(1): 23-51.

Johannes, W., 1985. The significance of experimental studies for the formation of

migmatites. Migmatites: 36-85.

Johannes, W. and Holtz, F., 1996. Petrogenesis and experimental petrology of granitic rocks,

22. Springer-Verlag Berlin Heidelberg.

Johnson, T. and Brown, M., 2004. Quantitative constraints on metamorphism in the

f
oo
Variscides of Southern Brittany—a complementary pseudosection approach. Journal

of Petrology, 45(6): 1237-1259.


pr
Johnson, T., Brown, M., Gibson, R. and Wing, B., 2004. Spinel-cordierite symplectites
e-
replacing andalusite: Evidence for melt-assisted diapirism in the Bushveld Complex,
Pr

South Africa. Journal of Metamorphic Geology, 22(6): 529-545.

Johnson, T., Hudson, N. and Droop, G., 2001a. Partial melting in the Inzie Head gneisses: the
al

role of water and a petrogenetic grid in KFMASH applicable to anatectic pelitic


n

migmatites. Journal of Metamorphic Geology, 19(1): 99-118.


ur

Johnson, T.E., Brown, M., Gardiner, N.J., Kirkland, C.L. and Smithies, R.H., 2017. Earth's
Jo

first stable continents did not form by subduction. Nature, 543(7644): 239-242.

Johnson, T.E., Brown, M. and Solar, G.S., 2003a. Low-pressure subsolidus and suprasolidus

phase equilibria in the MnNCKFMASH system: Constraints on conditions of regional

metamorphism in western Maine, northern Appalachians. American Mineralogist,

88(4): 624-638.

Johnson, T.E., Brown, M. and White, R.W., 2010. Petrogenetic modelling of strongly

residual metapelitic xenoliths within the southern Platreef, Bushveld Complex, South

Africa. Journal of Metamorphic Geology, 28(3): 269-291.


Journal Pre-proof

Johnson, T.E., Clark, C., Taylor, R.J.M., Santosh, M. and Collins, A.S., 2015. Prograde and

retrograde growth of monazite in migmatites: An example from the Nagercoil Block,

southern India. Geoscience Frontiers, 6(3): 373-387.

Johnson, T.E., Fischer, S. and White, R.W., 2013. Field and petrographic evidence for partial

melting of TTG gneisses from the central region of the mainland Lewisian complex,

NW Scotland. Journal of the Geological Society, 170(2): 319-326.

Johnson, T.E., Fischer, S., White, R.W., Brown, M. and Rollinson, H.R., 2012. Archaean

intracrustal differentiation from partial melting of metagabbro-field and geochemical

f
oo
evidence from the central region of the Lewisian complex, NW Scotland. Journal of

Petrology, 53(10): 2115-2138.


pr
Johnson, T.E., Gardiner, N.J., Miljković, K., Spencer, C.J., Kirkland, C.L., Bland, P.A. and
e-
Smithies, H., 2018. An impact melt origin for Earth’s oldest known evolved rocks.
Pr

Nature Geoscience, 11(10): 795-799.

Johnson, T.E., Gibson, R.L., Brown, M., Buick, I.S. and Cartwright, I., 2003b. Partial melting
al

of metapelitic rocks beneath the Bushveld Complex, South Africa. Journal of


n

Petrology, 44(5): 789-813.


ur

Johnson, T.E., Hudson, N.F.C. and Droop, G.T.R., 2001b. Melt segregation structures within
Jo

the inzie head gneisses of the Northeastern Dalradian. Scottish Journal of Geology,

37(2): 59-72.

Johnson, T.E., Hudson, N.F.C. and Droop, G.T.R., 2001c. Partial melting in the Inzie Head

gneisses: The role of water and a petrogenetic grid in KFMASH applicable to

anatectic pelitic migmatites. Journal of Metamorphic Geology, 19(1): 99-118.

Johnson, T.E., Kirkland, C.L., Gardiner, N.J., Brown, M., Smithies, R.H. and Santosh, M.,

2019. Secular change in TTG compositions: Implications for the evolution of

Archaean geodynamics. Earth and Planetary Science Letters, 505: 65-75.


Journal Pre-proof

Johnson, T.E., White, R.W. and Brown, M., 2011. A year in the life of an aluminous

metapelite xenolith-The role of heating rates, reaction overstep, H2O retention and

melt loss. Lithos, 124(1-2): 132-143.

Johnson, T.E., White, R.W. and Powell, R., 2008. Partial melting of metagreywacke: A

calculated mineral equilibria study. Journal of Metamorphic Geology, 26(8): 837-853.

Jones, K.A. and Brown, M., 1990. High‐temperature ‘clockwise’P‐T paths and melting in the

development of regional migmatites: an example from southern Brittany, France.

Journal of Metamorphic Geology, 8(5): 551-578.

f
oo
Jung, S. and Hellebrand, E., 2006. Trace element fractionation during high-grade

metamorphism and crustal melting - Constraints from ion microprobe data of


pr
metapelitic, migmatitic and igneous garnets and implications for Sm-Nd garnet
e-
chronology. Lithos, 87(3-4): 193-213.
Pr

Kaczmarek, M.A., Müntener, O. and Rubatto, D., 2008. Trace element chemistry and U-Pb

dating of zircons from oceanic gabbros and their relationship with whole rock
al

composition (Lanzo, Italian Alps). Contributions to Mineralogy and Petrology,


n

155(3): 295-312.
ur

Kawakami, T., Yamaguchi, I., Miyake, A., Shibata, T., Maki, K., Yokoyama, T.D. and
Jo

Hirata, T., 2013. Behavior of zircon in the upper-amphibolite to granulite facies

schist/migmatite transition, Ryoke metamorphic belt, SW Japan: Constraints from the

melt inclusions in zircon. Contributions to Mineralogy and Petrology, 165(3): 575-

591.

Keller, C.B., Schoene, B., Barboni, M., Samperton, K.M. and Husson, J.M., 2015. Volcanic–

plutonic parity and the differentiation of the continental crust. Nature, 523(7560):

301-307.
Journal Pre-proof

Kelsey, D.E., Clark, C. and Hand, M., 2008. Thermobarometric modelling of zircon and

monazite growth in melt-bearing systems: Examples using model metapelitic and

metapsammitic granulites. Journal of Metamorphic Geology, 26(2): 199-212.

Kelsey, D.E. 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(3): 311-356.

Kelsey, D.E. and Powell, R., 2011. Progress in linking accessory mineral growth and

breakdown to major mineral evolution in metamorphic rocks: A thermodynamic

f
oo
approach in the Na 2O-CaO-K 2O-FeO-MgO-Al 2O 3-SiO 2-H 2O-TiO 2-ZrO 2

system. Journal of Metamorphic Geology, 29(1): 151-166.


pr
Kelsey, D.E., White, R.W. and Powell, R., 2005. Calculated phase equilibria in K2O-FeO-
e-
MgO-Al2 O3-SiO2-H2O for silica-undersaturated sapphirine-bearing mineral
Pr

assemblages. Journal of Metamorphic Geology, 23(4): 217-239.

Kemp, A.I.S., Hawkesworth, C.J., Foster, G.L., Paterson, B.A., Woodhead, J.D., Hergt, J.M.,
al

Gray, C.M. and Whitehouse, M.J., 2007. Magmatic and crustal differentiation history
n

of granitic rocks from Hf-O isotopes in zircon. Science, 315(5814): 980-983.


ur

Kenah, C. and Hollister, L.S., 1983. Anatexis in the Central Gneiss complex, British
Jo

Columbia. Migmatites, Melting and Metamorphism: 142-162.

Kendrick, J. and Yakymchuk, C., 2020. Garnet fractionation, progressive melt loss and bulk

composition variations in anatectic metabasites: Complications for interpreting the

geodynamic significance of TTGs. Geoscience Frontiers, 11(3): 745-763.

Keppler, H., 2017. Fluids and trace element transport in subduction zones. American

Mineralogist, 102(1): 5-20.

Kirkland, C.L., Yakymchuk, C., Gardiner, N.J., Szilas, K., Hollis, J., Olierook, H. and

Steenfelt, A., 2020. Titanite petrochronology linked to phase equilibrium modelling


Journal Pre-proof

constrains tectono-thermal events in the Akia Terrane, West Greenland. Chemical

Geology, 536.

Kirkland, C.L., Yakymchuk, C., Olierook, H.K.H., Hartnady, M.I.H., Gardiner, N.J., Moyen,

J.F., Hugh Smithies, R., Szilas, K. and Johnson, T.E., 2021. Theoretical versus

empirical secular change in zircon composition. Earth and Planetary Science Letters.

Klimm, K., Blundy, J.D. and Green, T.H., 2008. Trace element partitioning and accessory

phase saturation during H2O-saturated melting of basalt with implications for

subduction zone chemical fluxes. Journal of Petrology, 49(3): 523-553.

f
oo
Koblinger, B.M. and Pattison, D.R.M., 2017. Crystallization of heterogeneous pelitic

migmatites: Insights from thermodynamic modelling. Journal of Petrology, 58(2):

297-326.
pr
e-
Kohn, M.J., Corrie, S.L. and Markley, C., 2015. The fall and rise of metamorphic zircon.
Pr

American Mineralogist, 100(4): 897-908.

Korhonen, F.J., Brown, M., Clark, C. and Bhattacharya, S., 2013a. Osumilite-melt
al

interactions in ultrahigh temperature granulites: Phase equilibria modelling and


n

implications for the P-T-t evolution of the eastern ghats province, india. Journal of
ur

Metamorphic Geology, 31(8): 881-907.


Jo

Korhonen, F.J., Clark, C., Brown, M., Bhattacharya, S. and Taylor, R., 2013b. How long-

lived is ultrahigh temperature (UHT) metamorphism? Constraints from zircon and

monazite geochronology in the Eastern Ghats orogenic belt, India. Precambrian

Research, 234: 322-350.

Korhonen, F.J., Powell, R. and Stout, J.H., 2012. Stability of sapphirine + quartz in the

oxidized rocks of the Wilson Lake terrane, Labrador: Calculated equilibria in

NCKFMASHTO. Journal of Metamorphic Geology, 30(1): 21-36.


Journal Pre-proof

Korhonen, F.J., Saito, S., Brown, M. and Siddoway, C.S., 2010a. Modeling multiple melt

loss events in the evolution of an active continental margin. Lithos, 116(3-4): 230-

248.

Korhonen, F.J., Saito, S., Brown, M., Siddoway, C.S. and Day, J.M.D., 2010b. Multiple

generations of granite in the Fosdick Mountains, Marie Byrd Land, West Antarctica:

Implications for polyphase intracrustal differentiation in a continental margin setting.

Journal of Petrology, 51(3): 627-670.

Koukouvelas, I.K., Pe-Piper, G. and Piper, D.J., 2006. The relationship between length and

f
oo
width of plutons within the crustal-scale Cobequid Shear Zone, northern

Appalachians, Canada. International Journal of Earth Sciences, 95(6): 963-976.


pr
Kramers, J.D., 1988. An open-system fractional crystallization model for very early
e-
continental crust formation. Precambrian Research, 38(4): 281-295.
Pr

Kunz, B.E., Johnson, T.E., White, R.W. and Redler, C., 2014. Partial melting of metabasic

rocks in Val Strona di Omegna, Ivrea Zone, northern Italy. Lithos, 190: 1-12.
al

Kunz, B.E., Regis, D. and Engi, M., 2018. Zircon ages in granulite facies rocks: decoupling
n

from geochemistry above 850° C? Contributions to Mineralogy and Petrology,


ur

173(3): 26.
Jo

Kunz, B.E. and White, R.W., 2019. Phase equilibrium modelling of the amphibolite to

granulite facies transition in metabasic rocks (Ivrea Zone, NW Italy). Journal of

Metamorphic Geology, 37(7): 935-950.

Kusky, T.M. and Polat, A., 1999. Growth of granite-greenstone terranes at convergent

margins, and stabilization of Archean cratons. Tectonophysics, 305(1-3): 43-73.

Kylander-Clark, A.R., Hacker, B.R. and Cottle, J.M., 2013. Laser-ablation split-stream ICP

petrochronology. Chemical Geology, 345: 99-112.


Journal Pre-proof

Labrousse, L., Duretz, T. and Gerya, T., 2015. H2O-fluid-saturated melting of subducted

continental crust facilitates exhumation of ultrahigh-pressure rocks in continental

subduction zones. Earth and Planetary Science Letters, 428: 151-161.

Labrousse, L., Prouteau, G. and Ganzhorn, A.-C., 2011. Continental exhumation triggered by

partial melting at ultrahigh pressure. Geology, 39(12): 1171-1174.

Lanari, P. and Engi, M., 2017. Local bulk composition effects on metamorphic mineral

assemblages. Reviews in Mineralogy and Geochemistry, 83(1): 55-102.

Lanari, P., Riel, N., Guillot, S., Vidal, O., Schwartz, S., Pêcher, A. and Hattori, K.H., 2013.

f
oo
Deciphering high-pressure metamorphism in collisional context using microprobe

mapping methods: Application to the Stak eclogitic massif (northwest Himalaya).

Geology, 41(2): 111-114.


pr
e-
Lang, H.M. and Gilotti, J.A., 2007. Partial melting of metapelites at ultrahigh-pressure
Pr

conditions, Greenland Caledonides. Journal of Metamorphic Geology, 25(2): 129-

147.
al

Lang, H.M. and Gilotti, J.A., 2015. Modeling the exhumation path of partially melted
n

ultrahigh-pressure metapelites, North-East Greenland Caledonides. Lithos, 226: 131-


ur

146.
Jo

Laporte, D. and Watson, E.B., 1995. Experimental and theoretical constraints on melt

distribution in crustal sources: the effect of crystalline anisotropy on melt

interconnectivity. Chemical Geology, 124(3-4): 161-184.

Lappin, A. and Hollister, L., 1980. Partial melting in the Central gneiss complex near Prince

Rupert, British Columbia. American Journal of Science, 280(6): 518-545.

Lasalle, S. and Indares, A., 2014. Anatectic record and contrasting P-T paths of aluminous

gneisses from the central Grenville Province. Journal of Metamorphic Geology,

32(6): 627-646.
Journal Pre-proof

Le Breton, N.L. and Thompson, A.B., 1988. Fluid-absent (dehydration) melting of biotite in

metapelites in the early stages of crustal anatexis. Contributions to Mineralogy and

Petrology, 99(2): 226-237.

Lentz, D.R., 1999. Carbonatite genesis: a reexamination of the role of intrusion-related

pneumatolytic skarn processes in limestone melting. Geology, 27(4): 335-338.

Li, L., Lin, S., Li, J., He, J. and Ge, Y., 2018. Zircon U-Pb ages and Hf isotope compositions

of the Chencai migmatite, central Zhejiang Province, South China: Constraints on the

early Palaeozoic orogeny. Geological Magazine, 155(6): 1377-1393.

f
oo
Li, R.-Y., Ke, S., Li, S., Song, S., Wang, C. and Liu, C., 2020. Origins of two types of

Archean potassic granite constrained by Mg isotopes and statistical geochemistry:


pr
Implications for continental crustal evolution. Lithos: 105570.
e-
Liou, P. and Guo, J., 2019. Generation of Archaean TTG Gneisses Through Amphibole-
Pr

Dominated Fractionation. Journal of Geophysical Research: Solid Earth, 124(4):

3605-3619.
al

López, S. and Castro, A., 2001. Determination of the fluid-absent solidus and supersolidus
n

relationships of MORB-derived amphibolites in the range 4-14 kbar. American


ur

Mineralogist, 86(11-12): 1396-1403.


Jo

Luth, W.C., Jahns, R.H. and Tuttle, O.F., 1964. The granite system at pressures of 4 to 10

kilobars. Journal of Geophysical Research, 69(4): 759-773.

Maaløe, S. and Wyllie, P.J., 1975. Water content of a granite magma deduced from the

sequence of crystallization determined experimentally with water-undersaturated

conditions. Contributions to Mineralogy and Petrology, 52(3): 175-191.

Macera, P., Di Pisa, A. and Gasperini, D., 2011. Geochemical and Sr-Nd isotope disequilibria

during multi-stage anatexis in a metasedimentary Hercynian crust. European Journal

of Mineralogy, 23(2): 207-222.


Journal Pre-proof

Malaspina, N., Hermann, J. and Scambelluri, M., 2009. Fluid/mineral interaction in UHP

garnet peridotite. Lithos, 107(1-2): 38-52.

Malaspina, N., Hermann, J., Scambelluri, M. and Compagnoni, R., 2006. Polyphase

inclusions in garnet–orthopyroxenite (Dabie Shan, China) as monitors for

metasomatism and fluid-related trace element transfer in subduction zone peridotite.

Earth and Planetary Science Letters, 249(3-4): 173-187.

Malaspina, N., Scambelluri, M., Poli, S., Van Roermund, H. and Langenhorst, F., 2010. The

oxidation state of mantle wedge majoritic garnet websterites metasomatised by C-

f
oo
bearing subduction fluids. Earth and Planetary Science Letters, 298(3-4): 417-426.

Marchildon, N. and Brown, M., 2001. Melt segregation in late syn-tectonic anatectic
pr
migmatites: An example from the Onawa Contact Aureole, Maine, USA. Physics and
e-
Chemistry of the Earth, Part A: Solid Earth and Geodesy, 26(4-5): 225-229.
Pr

Marchildon, N. and Brown, M., 2002. Grain-scale melt distribution in two contact aureole

rocks: Implications for controls on melt localization and deformation. Journal of


al

Metamorphic Geology, 20(4): 381-396.


n

Marchildon, N. and Brown, M., 2003. Spatial distribution of melt-bearing structures in


ur

anatectic rocks from Southern Brittany, France: Implications for melt transfer at
Jo

grain- to orogen-scale. Tectonophysics, 364(3-4): 215-235.

Marmo, B.A., Clarke, G.L. and Powell, R., 2002. Fractionation of bulk rock composition due

to porphyroblast growth: Effects on eclogite facies mineral equilibria, Pam Peninsula,

New Caledonia. Journal of Metamorphic Geology, 20(1): 151-165.

Martin, H., 1994. The Archean Grey Gneisses and the Genesis of Continental Crust,

Developments in Precambrian Geology, pp. 205-259.


Journal Pre-proof

Martin, H., Smithies, R.H., Rapp, R., Moyen, J.F. and Champion, D., 2005. An overview of

adakite, tonalite-trondhjemiten-granodiorite (TTG), and sanukitoid: Relationships and

some implications for crustal evolution. Lithos, 79(1-2 SPEC. ISS.): 1-24.

Massonne, H.-J., 2003. A comparison of the evolution of diamondiferous quartz-rich rocks

from the Saxonian Erzgebirge and the Kokchetav Massif: are so-called

diamondiferous gneisses magmatic rocks? Earth and Planetary Science Letters,

216(3): 347-364.

Massonne, H.-J. and Fockenberg, T., 2012. Melting of metasedimentary rocks at ultrahigh

f
oo
pressure—Insights from experiments and thermodynamic calculations. Lithosphere,

4(4): 269-285.
pr
Mayne, M.J., Moyen, J.F., Stevens, G. and Kaislaniemi, L., 2016. Rcrust: a tool for
e-
calculating path-dependent open system processes and application to melt loss.
Pr

Journal of Metamorphic Geology, 34(7): 663-682.

Mayne, M.J., Stevens, G. and Moyen, J.F., 2020a. A phase equilibrium investigation of
al

selected source controls on the composition of melt batches generated by sequential


n

melting of an average metapelite, Geological Society Special Publication, pp. 223-


ur

241.
Jo

Mayne, M.J., Stevens, G., Moyen, J.F. and Johnson, T., 2020b. Performing process-oriented

investigations involving mass transfer using rcrust: A new phase equilibrium

modelling tool, Geological Society Special Publication, pp. 209-221.

McCaffrey, K. and Petford, N., 1997. Are granitic intrusions scale invariant? Journal of the

Geological Society, 154(1): 1-4.

McDonough, W.F. and Sun, S.-S., 1995. The composition of the Earth. Chemical geology,

120(3): 223-253.
Journal Pre-proof

McKenzie, D. and Bickle, M.J., 1988. The Volume and Composition of Melt Generated by

Extension of the Lithosphere. Journal of Petrology, 29(3): 625-679.

McLellan, E.L., 1988. Migmatite structures in the Central Gneiss Complex, Boca de Quadra,

Alaska. Journal of Metamorphic Geology, 6(4): 517-542.

Mehnert, K., 1968. Migmatites and The Origin of Granitic Rocks. Elsevier, Amsterdam,

393p.

Milord, I., Sawyer, E.W. and Brown, M., 2001. Formation of diatexite migmatite and granite

magma during anatexis of semi-pelitic metasedimentary rocks: An example from St.

f
oo
Malo, France. Journal of Petrology, 42(3): 487-505.

Miron, G.D., Wagner, T., Kulik, D.A. and Heinrich, C.A., 2016. Internally consistent
pr
thermodynamic data for aqueous species in the system Na–K–Al–Si–O–H–Cl.
e-
Geochimica et Cosmochimica Acta, 187: 41-78.
Pr

Mitchell, R.J., Johnson, T.E., Clark, C., Gupta, S., Brown, M., Harley, S.L. and Taylor, R.,

2019. Neoproterozoic evolution and Cambrian reworking of ultrahigh temperature


al

granulites in the Eastern Ghats Province, India. Journal of Metamorphic Geology,


n

37(7): 977-1006.
ur

Mitchell, R.J., Johnson, T.E., Evans, K., Gupta, S. and Clark, C., 2020. Controls on the scales
Jo

of equilibrium during granulite facies metamorphism. Journal of Metamorphic

Geology, 39: 519-540.

Miyashiro, A., 1961. Evolution of metamorphic belts. Journal of petrology, 2(3): 277-311.

Mogk, D.W., 1992. Ductile shearing and migmatization at mid‐crustal levels in an Archaean

high‐grade gneiss belt, northern Gallatin Range, Montana, USA. Journal of

Metamorphic Geology, 10(3): 427-438.


Journal Pre-proof

Montel, J.-M. and Vielzeuf, D., 1997. Partial melting of metagreywackes, Part II.

Compositions of minerals and melts. Contributions to Mineralogy and Petrology,

128(2-3): 176-196.

Morfin, S., Sawyer, E.W. and Bandyayera, D., 2013. Large volumes of anatectic melt

retained in granulite facies migmatites: An injection complex in northern Quebec.

Lithos, 168-169: 200-218.

Morfin, S., Sawyer, E.W. and Bandyayera, D., 2014. The geochemical signature of a felsic

injection complex in the continental crust: Opinaca Subprovince, Quebec. Lithos,

f
oo
196-197: 339-355.

Moyen, J.-F., 2020a. Archean granitoids: classification, petrology, geochemistry and origin.
pr
Geological Society, London, Special Publications, 489(1): 15-49.
e-
Moyen, J.-F., Janoušek, V., Laurent, O., Bachmann, O., Jacob, J.-B., Farina, F., Fiannacca, P.
Pr

and Villaros, A., 2021. Crustal melting vs. fractionation of basaltic magmas: Part 1,

The bipolar disorder of granite petrogenetic models. Lithos: 106291.


al

Moyen, J.F., 2011. The composite Archaean grey gneisses: Petrological significance, and
n

evidence for a non-unique tectonic setting for Archaean crustal growth. Lithos, 123(1-
ur

4): 21-36.
Jo

Moyen, J.F., 2020b. Granites and crustal heat budget, Geological Society Special Publication.

Geological Society of London, pp. 77-100.

Moyen, J.F. and Martin, H., 2012. Forty years of TTG research. Lithos, 148: 312-336.

Moyen, J.F., Martin, H., Jayananda, M. and Auvray, B., 2003. Late Archaean granites: A

typology based on the Dharwar Craton (India). Precambrian Research, 127(1-3): 103-

123.
Journal Pre-proof

Moyen, J.F. and Stevens, G., 2006. Experimental constraints on TTG petrogenesis:

Implications for archean geodynamics. In: K.C. Condie, K. Benn and J.C. Mareschal

(Editors), Geophysical Monograph Series. Blackwell Publishing Ltd, pp. 149-175.

Mungall, J.E., Dingwell, D. and Chaussidon, M., 1999. Chemical diffusivities of 18 trace

elements in granitoid melts. Geochimica et Cosmochimica Acta, 63(17): 2599-2610.

Mysen, B., 2014. Water-melt interaction in hydrous magmatic systems at high temperature

and pressure. Progress in Earth and Planetary Science, 1(1): 4.

Mysen, B.O., 1978. Limits of solution of trace elements in minerals according to Henry's

f
oo
law: review of experimental data. Geochimica et Cosmochimica Acta, 42(6): 871-

885.
pr
Nabelek, P.I., 1999. Trace element distribution among rock-forming minerals in Black Hills
e-
migmatites, South Dakota: A case for solid-state equilibrium. American Mineralogist,
Pr

84(9): 1256-1269.

Nabelek, P.I., Whittington, A.G. and Hofmeister, A.M., 2010. Strain heating as a mechanism
al

for partial melting and ultrahigh temperature metamorphism in convergent orogens:


n

Implications of temperature‐dependent thermal diffusivity and rheology. Journal of


ur

Geophysical Research: Solid Earth, 115(B12).


Jo

Nair, R. and Chacko, T., 2002. Fluid-absent melting of high-grade semi-pelites: P-T

constraints on orthopyroxene formation and implications for granulite genesis.

Journal of Petrology, 43(11): 2121-2142.

Naumov, G.B., Ryzhenko, B.N. and Khodakovskii, I.L.v., 1974. Handbook of

thermodynamic data.

Nebel, O., Capitanio, F., Moyen, J.-F., Weinberg, R., Clos, F., Nebel-Jacobsen, Y. and

Cawood, P., 2018. When crust comes of age: on the chemical evolution of Archaean,

felsic continental crust by crustal drip tectonics. Philosophical Transactions of the


Journal Pre-proof

Royal Society A: Mathematical, Physical and Engineering Sciences, 376(2132):

20180103.

Nehring, F., 2012. Fluid-controlled melting of granulites and TTG-amphibolite associations

in the Iisalmi Complex, Central Finland, Special Paper of the Geological Survey of

Finland, pp. 244-254.

Nelson, K.D., Zhao, W., Brown, L., Kuo, J., Che, J., Liu, X., Klemperer, S., Makovsky, Y.,

Meissner, R. and Mechie, J., 1996. Partially molten middle crust beneath southern

Tibet: synthesis of project INDEPTH results. science, 274(5293): 1684-1688.

f
oo
Nemchin, A.A., Giannini, L.M., Bodorkos, S. and Oliver, N.H.S., 2001. Ostwald ripening as

a possible mechanism for zircon overgrowth formation during anatexis: Theoretical


pr
constraints, a numerical model, and its application to pelitic migmatites of the
e-
tickalara metamorphics, Northwestern Australia. Geochimica et Cosmochimica Acta,
Pr

65(16): 2771-2788.

Nyman, M.W., Pattison, D.R.M. and Ghent, E.D., 1995. Melt extraction during formation of
al

K-feldspar + sillimanite migmatites, west of revelstoke, British Columbia. Journal of


n

Petrology, 36(2): 351-372.


ur

O'Neill, H.S.C. and Wood, B., 1979. An experimental study of Fe-Mg partitioning between
Jo

garnet and olivine and its calibration as a geothermometer. Contributions to

Mineralogy and Petrology, 70(1): 59-70.

Palin, R.M., Weller, O.M., Waters, D.J. and Dyck, B., 2016a. Quantifying geological

uncertainty in metamorphic phase equilibria modelling; a Monte Carlo assessment

and implications for tectonic interpretations. Geoscience Frontiers, 7(4): 591-607.

Palin, R.M., White, R.W. and Green, E.C.R., 2016b. Partial melting of metabasic rocks and

the generation of tonalitic–trondhjemitic–granodioritic (TTG) crust in the Archaean:

Constraints from phase equilibrium modelling. Precambrian Research, 287: 73-90.


Journal Pre-proof

Palin, R.M., White, R.W., Green, E.C.R., Diener, J.F.A., Powell, R. and Holland, T.J.B.,

2016c. High-grade metamorphism and partial melting of basic and intermediate rocks.

Journal of Metamorphic Geology, 34(9): 871-892.

Pan, Y. and Fleet, M.E., 1996. Rare earth element mobility during prograde granulite facies

metamorphism: significance of fluorine. Contributions to Mineralogy and Petrology,

123(3): 251-262.

Parrish, R.R., 1990. U–Pb dating of monazite and its application to geological problems.

Canadian Journal of Earth Sciences, 27(11): 1431-1450.

f
oo
Patiño Douce, A.E., 1999. What do experiments tell us about the relative contributions of

crust and mantle to the origin of granitic magmas?, Geological Society Special

Publication, pp. 55-75.


pr
e-
Patiño Douce, A.E., 2005. Vapor-absent melting of tonalite at 15-32 kbar. Journal of
Pr

Petrology, 46(2): 275-290.

Patiño Douce, A.E., 1997. Generation of metaluminous A-type granites by low-pressure


al

melting of calc-alkaline granitoids. Geology, 25(8): 743-746.


n

Patiño Douce, A.E. and Beard, J.S., 1995. Dehydration-melting of biotite gneiss and quartz
ur

amphibolite from 3 to 15 kbar. Journal of Petrology, 36(3): 707-738.


Jo

Patiño Douce, A.E. and Beard, J.S., 1996. Effects of P, f(O2) and Mg/Fe ratio on dehydration

melting of model metagreywackes. Journal of Petrology, 37(5): 999-1024.

Patiño Douce, A.E. and Harris, N., 1998. Experimental constraints on Himalayan anatexis.

Journal of Petrology, 39(4): 689-710.

Patiño Douce, A.E. and Johnston, A.D., 1991. Phase equilibria and melt productivity in the

pelitic system: implications for the origin of peraluminous granitoids and aluminous

granulites. Contributions to Mineralogy and Petrology, 107(2): 202-218.


Journal Pre-proof

Pattison, D. and Harte, B., 1988. Evolution of structurally contrasting anatectic migmatites in

the 3‐kbar Ballachulish aureole, Scotland. Journal of Metamorphic Geology, 6(4):

475-494.

Pattison, D.R.M. and Harte, B., 1984. A petrogenetic grid for pelites in the Ballachulish and

other Scottish thermal aureoles. Journal of the Geological Society, 142(1): 7-28.

Pattison, D.R.M. and Harte, B., 1997. The geology and evolution of the Ballachulish Igneous

Complex and Aureole. Scottish Journal of Geology, 33(1): 1-29.

Pattison, D.R.M., Spear, F.S., Debuhr, C.L., Cheney, J.T. and Guidotti, C.V., 2002.

f
oo
Thermodynamic modelling of the reaction muscovite + cordierite → Al2SiO5 + biotite

+ quartz + H2O: Constraints from natural assemblages and implications for the
pr
metapelitic petrogenetic grid. Journal of Metamorphic Geology, 20(1): 99-118.
e-
Pearce, M., White, A. and Gazley, M., 2015. TCInvestigator: automated calculation of
Pr

mineral mode and composition contours for thermocalc pseudosections. Journal of

Metamorphic Geology, 33(4): 413-425.


al

Pichavant, M., Montel, J.-M. and Richard, L.R., 1992. Apatite solubility in peraluminous
n

liquids: Experimental data and an extension of the Harrison-Watson model.


ur

Geochimica et Cosmochimica Acta, 56(10): 3855-3861.


Jo

Pickering, J.M. and Johnston, D.A., 1998. Fluid-absent melting behavior of a two-mica

metapelite: experimental constraints on the origin of Black Hills granite. Journal of

Petrology, 39(10): 1787-1804.

Polat, A., Wang, L. and Appel, P.W.U., 2014. A review of structural patterns and melting

processes in the Archean craton of West Greenland: Evidence for crustal growth at

convergent plate margins as opposed to non-uniformitarian models. Tectonophysics,

662: 67-94.
Journal Pre-proof

Pourteau, A., Doucet, L.S., Blereau, E.R., Volante, S., Johnson, T.E., Collins, W.J., Li, Z.-X.

and Champion, D.C., 2020. TTG generation by fluid-fluxed crustal melting: Direct

evidence from the Proterozoic Georgetown Inlier, NE Australia. Earth and Planetary

Science Letters, 550: 116548.

Powell, R., 1978. Equilibrium thermodynamics in petrology: an introduction. Not Avail.

Powell, R., Evans, K.A., Green, E.C. and White, R.W., 2019. The truth and beauty of

chemical potentials. Journal of Metamorphic Geology, 37(7): 1007-1019.

Powell, R., Guiraud, M. and White, R.W., 2005. Truth and beauty in metamorphic phase-

f
oo
equilibria: conjugate variables and phase diagrams. The Canadian Mineralogist,

43(1): 21-33.
pr
Powell, R., Holland, T. and Worley, B., 1998. Calculating phase diagrams involving solid
e-
solutions via non-linear equations, with examples using THERMOCALC. Journal of
Pr

Metamorphic Geology, 16(4): 577-588.

Powell, R. and Holland, T.J.B., 1988. An internally consistent dataset with uncertainties and
al

correlations: 3. Applications to geobarometry, worked examples and a computer


n

program. Journal of Metamorphic Geology, 6(2): 173-204.


ur

Powell, R. and Holland, T.J.B., 2008. On thermobarometry. Journal of Metamorphic


Jo

Geology, 26(2): 155-179.

Powell, R., White, R., Green, E., Holland, T. and Diener, J., 2014. On parameterizing

thermodynamic descriptions of minerals for petrological calculations. Journal of

Metamorphic Geology, 32(3): 245-260.

Prent, A.M., Beinlich, A., Raimondo, T., Kirkland, C.L., Evans, N.J. and Putnis, A., 2020.

Apatite and monazite: An effective duo to unravel superimposed fluid-flow and

deformation events in reactivated shear zones. Lithos, 376: 105752.


Journal Pre-proof

Pressley, R.A. and Brown, M., 1999. The Phillips pluton, Maine, USA: Evidence of

heterogeneous crustal sources and implications for granite ascent and emplacement

mechanisms in convergent orogens. Lithos, 46(3): 335-366.

Pyle, J.M. and Spear, F.S., 1999. Yttrium zoning in garnet: coupling of major and accessory

phases during metamorphic reactions. Geological Materials Research, 1(6): 1-49.

Pyle, J.M. and Spear, F.S., 2003. Four generations of accessory-phase growth in low-pressure

migmatites from SW New Hampshire. American Mineralogist, 88(2-3): 338-351.

Pyle, J.M., Spear, F.S., Rudnick, R.L. and McDonough, W.F., 2001. Monazite–xenotime–

f
oo
garnet equilibrium in metapelites and a new monazite–garnet thermometer. Journal of

Petrology, 42(11): 2083-2107.


pr
Rapp, R.P., Ryerson, F.J. and Miller, C.F., 1987. Experimental evidence bearing on the
e-
stability of monazite during crustal anaatexis. Geophysical Research Letters, 14(3):
Pr

307-310.

Rapp, R.P. and Watson, E.B., 1986. Monazite solubility and dissolution kinetics: implications
al

for the thorium and light rare earth chemistry of felsic magmas. Contributions to
n

Mineralogy and Petrology, 94(3): 304-316.


ur

Rapp, R.P. and Watson, E.B., 1995. Dehydration melting of metabasalt at 8-32 kbar:
Jo

Implications for continental growth and crust-mantle recycling. Journal of Petrology,

36(4): 891-931.

Rapp, R.P., Watson, E.B. and Miller, C.F., 1991. Partial melting of amphibolite/eclogite and

the origin of Archean trondhjemites and tonalites. Precambrian Research, 51(1-4): 1-

25.

Read, H.H., 1952. Metamorphism and migmatisation in the Ythan Valley, Aberdeenshire.

Transactions of the Edinburgh Geological Society, 15(1): 265-279.


Journal Pre-proof

Read, H.H., 1957. The granite controversy: geological addresses illustrating the evolution of

a disputant. T. Murby.

Renner, J., Evans, B. and Hirth, G., 2000. On the rheologically critical melt fraction. Earth

and Planetary Science Letters, 181(4): 585-594.

Riesco, M., Stüwe, K. and Reche, J., 2005. Formation of corundum in metapelites around

ultramafic bodies. An example from the Saualpe region, Eastern Alps. Mineralogy

and Petrology, 83(1-2): 1-25.

f
oo
deposit, Duluth Complex, Minnesota, USA. Lithos, 21(2): 143-159.

Robertson, J. and Wyllie, P., 1971a. Experimental studies on rocks from the Deboullie stock,
pr
northern Maine, including melting relations in the water-deficient environment. The
e-
Journal of Geology, 79(5): 549-571.
Pr

Robertson, J. and Wyllie, P., 1971b. Rock-water systems, with special reference to the water-

deficient region. American Journal of Science, 271(3): 252-277.


al

Rollinson, H. and Martin, H., 2005. Geodynamic controls on adakite, TTG and sanukitoid
n

genesis: Implications for models of crust formation. Lithos, 79(1-2 SPEC. ISS.): ix-
ur

xii.
Jo

Rosenberg, C. and Handy, M., 2005. Experimental deformation of partially melted granite

revisited: implications for the continental crust. Journal of metamorphic Geology,

23(1): 19-28.

Rubatto, D., 2002. Zircon trace element geochemistry: Partitioning with garnet and the link

between U-Pb ages and metamorphism. Chemical Geology, 184(1-2): 123-138.

Rubatto, D., 2017. Zircon: the metamorphic mineral. Reviews in mineralogy and

geochemistry, 83(1): 261-295.


Journal Pre-proof

Rubatto, D., Chakraborty, S. and Dasgupta, S., 2013. Timescales of crustal melting in the

Higher Himalayan Crystallines (Sikkim, Eastern Himalaya) inferred from trace

element-constrained monazite and zircon chronology. Contributions to Mineralogy

and Petrology, 165(2): 349-372.

Rubatto, D. and Hermann, J., 2003. Zircon formation during fluid circulation in eclogites

(Monviso, Western Alps): Implications for Zr and Hf budget in subduction zones.

Geochimica et Cosmochimica Acta, 67(12): 2173-2187.

Rubatto, D. and Hermann, J., 2007. Experimental zircon/melt and zircon/garnet trace element

f
oo
partitioning and implications for the geochronology of crustal rocks. Chemical

Geology, 241(1-2): 38-61.


pr
Rubatto, D., Hermann, J., Berger, A. and Engi, M., 2009. Protracted fluid-induced melting
e-
during Barrovian metamorphism in the Central Alps. Contributions to Mineralogy and
Pr

Petrology, 158(6): 703-722.

Rubatto, D., Hermann, J. and Buick, I.S., 2006. Temperature and bulk composition control on
al

the growth of monazite and zircon during low-pressure anatexis (Mount Stafford,
n

Central Australia). Journal of Petrology, 47(10): 1973-1996.


ur

Rubie, D.C. and Brearley, A.J., 1990. A model for rates of disequilibrium melting during
Jo

metamorphism, High-temperature metamorphism and crustal anatexis. Springer, pp.

57-86.

Rudnick, R. and Gao, S., 2003. Composition of the continental crust. The crust, 3: 1-64.

Rudnick, R.L. and Fountain, D.M., 1995. Nature and composition of the continental crust: a

lower crustal perspective. Reviews of geophysics, 33(3): 267-309.

Rushmer, T., 1991. Partial melting of two amphibolites: contrasting experimental results

under fluid-absent conditions. Contributions to Mineralogy and Petrology, 107(1): 41-

59.
Journal Pre-proof

Saukko, A., Ahläng, C., Nikkilä, K., Soesoo, A. and Eklund, O., 2020. Double Power-Law in

Leucosome Width Distribution: Implications for Recognizing Melt Movement in

Migmatites. Front. Earth Sci, 8: 591871.

Sawyer, E., 2014a. The inception and growth of leucosomes: microstructure at the start of

melt segregation in migmatites. Journal of Metamorphic Geology, 32(7): 695-712.

Sawyer, E., 2020. Petrogenesis of secondary diatexites and the melt budget for crustal

reworking. Journal of Petrology.

Sawyer, E.W., 1987. The role of partial melting and fractional crystallization in determining

f
oo
discordant migmatite leucosome compositions. Journal of Petrology, 28(3): 445-473.

Sawyer, E.W., 1991. Disequilibrium melting and the rate of melt-residuum separation during
pr
migmatization of mafic rocks from the Grenville front, Quebec. Journal of Petrology,
e-
32(4): 701-738.
Pr

Sawyer, E.W., 1998. Formation and evolution of granite magmas during crustal reworking:

the significance of diatexites. Journal of Petrology, 39(6): 1147-1167.


al

Sawyer, E.W., 1999. Criteria for the recognition of partial melting. Physics and Chemistry of
n

the Earth, Part A: Solid Earth and Geodesy, 24(3): 269-279.


ur

Sawyer, E.W., 2001. Melt segregation in the continental crust: Distribution and movement of
Jo

melt in anatectic rocks. Journal of Metamorphic Geology, 19(3): 291-309.

Sawyer, E.W., 2008. Atlas of migmatites, 9. NRC Research press.

Sawyer, E.W., 2010. Migmatites formed by water-fluxed partial melting of a

leucogranodiorite protolith: Microstructures in the residual rocks and source of the

fluid. Lithos, 116(3-4): 273-286.

Sawyer, E.W., 2014b. The inception and growth of leucosomes: Microstructure at the start of

melt segregation in migmatites. Journal of Metamorphic Geology, 32(7): 695-712.


Journal Pre-proof

Sawyer, E.W. and Barnes, S.J., 1988. Temporal and compositional differences between

subsolidus and anatectic migmatite leucosomes from the Quetico metasedimentary

belt, Canada. Journal of Metamorphic Geology, 6(4): 437-450.

Sawyer, E.W., Cesare, B. and Brown, M., 2011. When the continental crust melts. Elements,

7(4): 229-234.

Sawyer, E.W. and Robin, P.Y.F., 1986. The subsolidus segregation of layer‐parallel quartz‐

feldspar veins in greenschist to upper amphibolite facies metasediments. Journal of

Metamorphic Geology, 4(3): 237-260.

f
oo
Saxena, S.K., Chatterjee, N., Fei, Y. and Shen, G., 2012. Thermodynamic data on oxides and

silicates: an assessed data set based on thermochemistry and high pressure phase
pr
equilibrium. Springer Science & Business Media.
e-
Scambelluri, M., Pettke, T. and Van Roermund, H., 2008. Majoritic garnets monitor deep
Pr

subduction fluid flow and mantle dynamics. Geology, 36(1): 59-62.

Scambelluri, M. and Philippot, P., 2001. Deep fluids in subduction zones. Lithos, 55(1-4):
al

213-227.
n

Schaltegger, U., Fanning, C.M., Günther, D., Maurin, J.C., Schulmann, K. and Gebauer, D.,
ur

1999. Growth, annealing and recrystallization of zircon and preservation of monazite


Jo

in high-grade metamorphism: Conventional and in-situ U-Pb isotope,

cathodoluminescence and microchemical evidence. Contributions to Mineralogy and

Petrology, 134(2-3): 186-201.

Scherer, E.E., Cameron, K.L. and Blichert-Toft, J., 2000. Lu–Hf garnet geochronology:

closure temperature relative to the Sm–Nd system and the effects of trace mineral

inclusions. Geochimica et Cosmochimica Acta, 64(19): 3413-3432.

Schorn, S., Diener, J.F., Powell, R. and Stüwe, K., 2018. Thermal buffering in the orogenic

crust. Geology, 46(7): 643-646.


Journal Pre-proof

Schreinemakers, F., 1915. In-, Mono-, and Divariant Equilibria: Koninkl. Akad.

Wetenschappen te Amsterdam Proc., English ed, 18: 28.

Schubert, G. and Reymer, A., 1985. Continental volume and freeboard through geological

time. Nature, 316(6026): 336-339.

Schulmann, K., Martelat, J.E., Ulrich, S., Lexa, O., Štípská, P. and Becker, J.K., 2008.

Evolution of microstructure and melt topology in partially molten granitic mylonite:

Implications for rheology of felsic middle crust. Journal of Geophysical Research:

Solid Earth, 113(B10).

f
oo
Schumacher, J.C., 1991. Empirical ferric iron corrections: necessity, assumptions, and effects

on selected geothermobarometers. Mineralogical Magazine, 55(378): 3-18.


pr
Schwindinger, M. and Weinberg, R.F., 2017. A felsic MASH zone of crustal magmas —
e-
Feedback between granite magma intrusion and in situ crustal anatexis. Lithos, 284-
Pr

285: 109-121.

Schwindinger, M., Weinberg, R.F. and Clos, F., 2019. Wet or dry? The difficulty of
al

identifying the presence of water during crustal melting. Journal of Metamorphic


n

Geology, 37(3): 339-358.


ur

Schwindinger, M., Weinberg, R.F. and White, R.W., 2020. The fate of accessory minerals
Jo

and key trace elements during anatexis and magma extraction. Journal of Petrology.

Semprich, J., Moreno, J.A. and Oliveira, E.P., 2015. Phase equilibria and trace element

modeling of Archean sanukitoid melts. Precambrian Research, 269: 122-138.

Sen, C. and Dunn, T., 1994. Dehydration melting of a basaltic composition amphibolite at 1.5

and 2.0 GPa: implications for the origin of adakites. Contributions to Mineralogy and

Petrology, 117(4): 394-409.

Shand, S., 1927. On the relations between silica, alumina, and the bases in eruptive rocks,

considered as a means of classification. Geological Magazine, 64(10): 446-449.


Journal Pre-proof

Shand, S.J., 1943. Eruptive rocks: their genesis, composition, and classification, with a

chapter on meteorites. J. Wiley & sons, Incorporated.

Shaw, D.M., 1956. Geochemistry of pelitic rocks. Part III: Major elements and general

geochemistry. Geological Society of America Bulletin, 67(7): 919-934.

Shaw, D.M., 1970. Trace element fractionation during anatexis. Geochimica et

Cosmochimica Acta, 34(2): 237-243.

Shrestha, S., Larson, K.P., Duesterhoeft, E., Soret, M. and Cottle, J.M., 2019.

Thermodynamic modelling of phosphate minerals and its implications for the

f
oo
development of PTt histories: A case study in garnet-monazite bearing metapelites.

Lithos, 334: 141-160.


pr
Singh, J. and Johannes, W., 1996a. Dehydration melting of tonalites. Part I. Beginning of
e-
melting. Contributions to Mineralogy and Petrology, 125(1): 16-25.
Pr

Singh, J. and Johannes, W., 1996b. Dehydration melting of tonalites. Part II. Composition of

melts and solids. Contributions to Mineralogy and Petrology, 125(1): 26-44.


al

Sisson, T., Ratajeski, K., Hankins, W. and Glazner, A.F., 2005. Voluminous granitic magmas
n

from common basaltic sources. Contributions to Mineralogy and Petrology, 148(6):


ur

635-661.
Jo

Sizova, E., Gerya, T. and Brown, M., 2012. Exhumation mechanisms of melt‐bearing

ultrahigh pressure crustal rocks during collision of spontaneously moving plates.

Journal of Metamorphic Geology, 30(9): 927-955.

Skora, S. and Blundy, J., 2010. High-pressure hydrous phase relations of radiolarian clay and

implications for the involvement of subducted sediment in arc magmatism. Journal of

Petrology, 51(11): 2211-2243.


Journal Pre-proof

Slagstad, T., Jamieson, R.A. and Culshaw, N.G., 2005. Formation, crystallization, and

migration of melt in the mid-orogenic crust: Muskoka domain migmatites, Grenville

Province, Ontario. Journal of Petrology, 46(5): 893-919.

Smit, M.A., Scherer, E.E. and Mezger, K., 2013. Lu–Hf and Sm–Nd garnet geochronology:

Chronometric closure and implications for dating petrological processes. Earth and

Planetary Science Letters, 381: 222-233.

Smithies, R., Yongjun, L., Kirkland, C., Johnson, T., Mole, D., Martin, L., KHeejin, J.,

Wingate, M. and Johnson, S., 2021. Oxygen isotopes in zircon trace the origins of

f
oo
Earth’s earliest continental crust. Nature, XX: xxx–xxx.

Smithies, R.H., 2000. The Archaean tonalite-trondhjemite-granodiorite (TTG) series is not an


pr
analogue of Cenozoic adakite. Earth and Planetary Science Letters, 182(1): 115-125.
e-
Smithies, R.H., Champion, D.C. and Van Kranendonk, M.J., 2009. Formation of
Pr

Paleoarchean continental crust through infracrustal melting of enriched basalt. Earth

and Planetary Science Letters, 281(3-4): 298-306.


al

Smithies, R.H., Lu, Y., Johnson, T.E., Kirkland, C.L., Cassidy, K.F., Champion, D.C., Mole,
n

D.R., Zibra, I., Gessner, K. and Sapkota, J., 2019. No evidence for high-pressure
ur

melting of Earth’s crust in the Archean. Nature communications, 10(1): 1-12.


Jo

Smye, A.J., Roberts, N.M., Condon, D.J., Horstwood, M.S. and Parrish, R.R., 2014.

Characterising the U–Th–Pb systematics of allanite by ID and LA-ICPMS:

Implications for geochronology. Geochimica et Cosmochimica Acta, 135: 1-28.

Soesoo, A. and Bons, P.D., 2015. From Migmatites to Plutons: Power Law Relationships in

the Evolution of Magmatic Bodies. Pure and Applied Geophysics, 172(7): 1787-1801.

Solar, G.S. and Brown, M., 2001a. Deformation partitioning during transpression in response

to Early Devonian oblique convergence, Northern Appalachian orogen, USA. Journal

of Structural Geology, 23(6-7): 1043-1065.


Journal Pre-proof

Solar, G.S. and Brown, M., 2001b. Petrogenesis of migmatites in Maine, USA: Possible

source of perluminous leucogranite in Plutons? Journal of Petrology, 42(4): 789-823.

Spear, F.S., 1995. Metamorphic phase equilibria and pressure–temperature–time paths.

Monograph. Mineralogical Society of America, Washington DC, 799 pp.

Spear, F.S. and Kohn, M.J., 1996. Trace element zoning in garnet as a monitor of crustal

melting. Geology, 24(12): 1099-1102.

Spear, F.S., Kohn, M.J. and Cheney, J.T., 1999. P-T paths from anatectic pelites.

Contributions to Mineralogy and Petrology, 134(1): 17-32.

f
oo
Spear, F.S. and Menard, T., 1989. Program GIBBS; a generalized Gibbs method algorithm.

American Mineralogist, 74(7-8): 942-943.


pr
Spear, F.S. and Pyle, J.M., 2002. Apatite, monazite, and xenotime in metamorphic rocks.
e-
Reviews in Mineralogy and Geochemistry, 48(1): 293-335.
Pr

Spear, F.S. and Pyle, J.M., 2010. Theoretical modeling of monazite growth in a low-Ca

metapelite. Chemical Geology, 273(1-2): 111-119.


al

Spicer, E.M., Stevens, G. and Buick, I.S., 2004. The low-pressure partial-melting behaviour
n

of natural boron-bearing metapelites from the Mt. Stafford area, central Australia.
ur

Contributions to Mineralogy and Petrology, 148(2): 160-179.


Jo

Stepanov, A.S., Hermann, J., Korsakov, A.V. and Rubatto, D., 2014. Geochemistry of

ultrahigh-pressure anatexis: fractionation of elements in the Kokchetav gneisses

during melting at diamond-facies conditions. Contributions to Mineralogy and

Petrology, 167(5): 1002.

Stepanov, A.S., Hermann, J., Rubatto, D., Korsakov, A.V. and Danyushevsky, L.V., 2016.

Melting history of an ultrahigh-pressure paragneiss revealed by multiphase solid

inclusions in garnet, Kokchetav massif, Kazakhstan. Journal of Petrology, 57(8):

1531-1554.
Journal Pre-proof

Stepanov, A.S., Hermann, J., Rubatto, D. and Rapp, R.P., 2012. Experimental study of

monazite/melt partitioning with implications for the REE, Th and U geochemistry of

crustal rocks. Chemical Geology, 300-301: 200-220.

Stern, C.R. and Wyllie, P.J., 1981. Phase relationships of I-type granite with H2O to 35

kilobars: the Dinkey Lakes biotite-granite from the Sierra Nevada batholith. Journal

of Geophysical Research, 86(B11): 10412-10422.

Stevens, G. and Clemens, J., 1993. Fluid-absent melting and the roles of fluids in the

lithosphere: a slanted summary? Chemical Geology, 108(1-4): 1-17.

f
oo
Stevens, G., Clemens, J.D. and Droop, G.T., 1997. Melt production during granulite-facies

anatexis: experimental data from “primitive” metasedimentary protoliths.


pr
Contributions to Mineralogy and Petrology, 128(4): 352-370.
e-
Stevens, G., Prinz, S. and Rozendaal, A., 2005. Partial melting of the assemblage sphalerite+
Pr

galena+ pyrrhotite+ chalcopyrite+ sulfur: implications for high-grade metamorphosed

massive sulfide deposits. Economic Geology, 100(4): 781-786.


al

Stevens, G., Villaros, A. and Moyen, J.-F., 2007. Selective peritectic garnet entrainment as
n

the origin of geochemical diversity in S-type granites. Geology, 35(1): 9-12.


ur

Štípská, P., Hasalová, P., Powell, R., Závada, P., Schulmann, K., Racek, M., Aguilar, C. and
Jo

Chopin, F., 2019. The effect of melt infiltration on metagranitic rocks: The Snieznik

dome, Bohemian Massif. Journal of Petrology, 60(3): 591-618.

Stuart, C.A., Piazolo, S. and Daczko, N.R., 2018. The recognition of former melt flux

through high‐strain zones. Journal of Metamorphic Geology, 36(8): 1049-1069.

Stuck, T.J. and Diener, J.F.A., 2018. Mineral equilibria constraints on open-system melting in

metamafic compositions. Journal of Metamorphic Geology, 36(2): 255-281.

Stüwe, K., 1995. Thermal buffering effects at the solidus. Implications for the equilibration

of partially melted metamorphic rocks. Tectonophysics, 248(1-2): 39-51.


Journal Pre-proof

Sun, S.s. and McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts:

implications for mantle composition and processes. Geological Society, London,

Special Publications, 42(1): 313-345.

Sylvester, P.J., 1998. Post-collisional strongly peraluminous granites. Lithos, 45(1-4): 29-44.

Tafur, L.A. and Diener, J.F.A., 2020. Mineral equilibrium constraints on the feasibility of

diffusive H2O-fluxed melting in the continental crust. Journal of Metamorphic

Geology.

Tanner, D.C., 1999. The scale-invariant nature of migmatite from the Oberpfalz, NE Bavaria

f
oo
and its significance for melt transport. Tectonophysics, 302(3-4): 297-305.

Taylor Jr, H.P., 1978. Oxygen and hydrogen isotope studies of plutonic granitic rocks. Earth
pr
and Planetary Science Letters, 38(1): 177-210.
e-
Taylor, R., Clark, C., Harley, S., Kylander‐Clark, A., Hacker, B. and Kinny, P., 2017.
Pr

Interpreting granulite facies events through rare earth element partitioning arrays.

Journal of Metamorphic Geology, 35(7): 759-775.


al

Taylor, R., Harley, S., Hinton, R., Elphick, S., Clark, C. and Kelly, N., 2015a. Experimental
n

determination of REE partition coefficients between zircon, garnet and melt: a key to
ur

understanding high‐T crustal processes. Journal of Metamorphic Geology, 33(3): 231-


Jo

248.

Taylor, R.J., Kirkland, C.L. and Clark, C., 2016. Accessories after the facts: Constraining the

timing, duration and conditions of high-temperature metamorphic processes. Lithos,

264: 239-257.

Taylor, R.J.M., Harley, S.L., Hinton, R.W., Elphick, S., Clark, C. and Kelly, N.M., 2015b.

Experimental determination of REE partition coefficients between zircon, garnet and

melt: A key to understanding high-T crustal processes. Journal of Metamorphic

Geology, 33(3): 231-248.


Journal Pre-proof

Taylor, S.R. and McLennan, S.M., 1985. The continental crust: its composition and

evolution.

Teyssier, C. and Whitney, D.L., 2002. Gneiss domes and orogeny. Geology, 30(12): 1139-

1142.

Thompson, A.B., 1982. Dehydration melting of pelitic rocks and the generation of H 2 O-

undersaturated granitic liquids. American Journal of Science, 282(10): 1567-1595.

Thompson Jr, J.B., 1957. The graphical analysis of mineral assemblages in pelitic schists.

American Mineralogist: Journal of Earth and Planetary Materials, 42(11-12): 842-

f
oo
858.

Tinkham, D.K. and Ghent, E.D., 2005a. Estimating P-T conditions of garnet growth with
pr
isochemical phase-diagram sections and the problem of effective bulk-composition.
e-
Canadian Mineralogist, 43(1): 35-50.
Pr

Tinkham, D.K. and Ghent, E.D., 2005b. XRMapAnal: A program for analysis of quantitative

X-ray maps. American Mineralogist, 90(4): 737-744.


al

Tollari, N., Toplis, M. and Barnes, S.-J., 2006. Predicting phosphate saturation in silicate
n

magmas: an experimental study of the effects of melt composition and temperature.


ur

Geochimica et Cosmochimica Acta, 70(6): 1518-1536.


Jo

Tomkins, A.G., Pattison, D.R.M. and Frost, B.R., 2007. On the initiation of metamorphic

sulfide anatexis. Journal of Petrology, 48(3): 511-535.

Turner, S. and Rushmer, T., 2009. Similarities between mantle-derived A-type granites and

voluminous rhyolites in continental flood basalt provinces. Earth and Environmental

Science Transactions of the Royal Society of Edinburgh, 100(1-2): 51-60.

Tuttle, O.F. and Bowen, N.L., 1958. Origin of granite in the light of experimental studies in

the system NaAlSi3O8-KAlSi3O8-SiO 2-H2O, Memoir of the Geological Society of

America, pp. 1-145.


Journal Pre-proof

Unsworth, M., Jones, A.G., Wei, W., Marquis, G., Gokarn, S. and Spratt, J., 2005. Crustal

rheology of the Himalaya and Southern Tibet inferred from magnetotelluric data.

Nature, 438(7064): 78-81.

van der Laan, S.R. and Wyllie, P.J., 1992. Constraints on Archean trondhjemite genesis from

hydrous crystallization experiments on Nuk Gneiss at 10-17 Kbar. The Journal of

Geology, 100(1): 57-68.

Vance, D. and Mahar, E., 1998. Pressure-temperature paths from P-T pseudosections and

zoned garnets: Potential, limitations and examples from the Zanskar Himalaya, NW

f
oo
India. Contributions to Mineralogy and Petrology, 132(3): 225-245.

Vavra, G., Schmid, R. and Gebauer, D., 1999. Internal morphology, habit and U-Th-Pb
pr
microanalysis of amphibolite-to-granulite facies zircons: Geochronology of the Ivrea
e-
Zone (Southern Alps). Contributions to Mineralogy and Petrology, 134(4): 380-404.
Pr

Vernon, R.H., 2018. A practical guide to rock microstructure. Cambridge university press.

Vervoort, J.D. and Kemp, A.I., 2016. Clarifying the zircon Hf isotope record of crust–mantle
al

evolution. Chemical Geology, 425: 65-75.


n

Vielzeuf, D. and Holloway, J.R., 1988. Experimental determination of the fluid-absent


ur

melting relations in the pelitic system. Contributions to Mineralogy and Petrology,


Jo

98(3): 257-276.

Vielzeuf, D. and Montel, J.M., 1994. Partial melting of metagreywackes. Part I. Fluid-absent

experiments and phase relationships. Contributions to Mineralogy and Petrology,

117(4): 375-393.

Vielzeuf, D. and Schmidt, M., 2001. Melting relations in hydrous systems revisited:

application to metapelites, metagreywackes and metabasalts. Contributions to

Mineralogy and Petrology, 141(3): 251.


Journal Pre-proof

Vigneresse, J.L., 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(6): 1579-1600.

Villaros, A., Stevens, G., Moyen, J.F. and Buick, I.S., 2009. The trace element compositions

of S-type granites: Evidence for disequilibrium melting and accessory phase

entrainment in the source. Contributions to Mineralogy and Petrology, 158(4): 543-

561.

von Seckendorff, V. and O'Neill, H.S.C., 1993. An experimental study of Fe-Mg partitioning

f
oo
between olivine and orthopyroxene at 1173, 1273 and 1423 K and 1.6 GPa.

Contributions to Mineralogy and Petrology, 113(2): 196-207.


pr
Wang, L., Wang, S.J., Brown, M., Zhang, J.F., Feng, P. and Jin, Z.M., 2018. On the survival
e-
of intergranular coesite in UHP eclogite. Journal of Metamorphic Geology, 36(2):
Pr

173-194.

Wang, S., Wang, L., Brown, M. and Feng, P., 2016. Multi-stage barite crystallization in
al

partially melted UHP eclogite from the Sulu belt, China. American Mineralogist,
n

101(3): 564-579.
ur

Wang, S.-J., Wang, L., Brown, M., Piccoli, P.M., Johnson, T.E., Feng, P., Deng, H.,
Jo

Kitajima, K. and Huang, Y., 2017. Fluid generation and evolution during exhumation

of deeply subducted UHP continental crust: Petrogenesis of composite granite–quartz

veins in the Sulu belt, China. Journal of Metamorphic Geology, 35(6): 601-629.

Wang, S.J., Wang, L., Brown, M., Johnson, T.E., Piccoli, P.M., Feng, P. and Wang, Z.L.,

2020. Petrogenesis of leucosome sheets in migmatitic UHP eclogites—Evolution

from silicate-rich supercritical fluid to hydrous melt. Lithos, 360-361.

Wang, W.R., Dunkley, E., Clarke, G.L. and Daczko, N.R., 2014. The evolution of zircon

during low-P partial melting of metapelitic rocks: Theoretical predictions and a case
Journal Pre-proof

study from Mt Stafford, central Australia. Journal of Metamorphic Geology, 32(8):

791-808.

Ward, R.F. and Werner, S.L., 1964. Analysis of variance of migmatite composition II:

Comparison of two areas. Science, 143(3610): 1032-1033.

Waters, D., 1988. Partial melting and the formation of granulite facies assemblages in

Namaqualand, South Africa. Journal of Metamorphic Geology, 6(4): 387-404.

Watkins, J., Clemens, J. and Treloar, P., 2007. Archaean TTGs as sources of younger granitic

magmas: melting of sodic metatonalites at 0.6–1.2 GPa. Contributions to Mineralogy

f
oo
and Petrology, 154(1): 91-110.

Watson, E.B., 1996. Dissolution, growth and survival of zircons during crustal fusion:
pr
Kinetic principals, geological models and implications for isotopic inheritance.
e-
Transactions of the Royal Society of Edinburgh, Earth Sciences, 87(1-2): 43-56.
Pr

Watson, E.B. and Harrison, T.M., 1983. Zircon saturation revisited: temperature and

composition effects in a variety of crustal magma types. Earth and Planetary Science
al

Letters, 64(2): 295-304.


n

Watson, E.B. and Harrison, T.M., 1984. Accessory minerals and the geochemical evolution
ur

of crustal magmatic systems: a summary and prospectus of experimental approaches.


Jo

Physics of the Earth and Planetary Interiors, 35(1-3): 19-30.

Watt, G.R., Burns, I.M. and Graham, G.A., 1996. Chemical characteristics of migmatites:

Accessory phase distribution and evidence for fast melt segregation rates.

Contributions to Mineralogy and Petrology, 125(1): 100-111.

Watt, G.R. and Harley, S.L., 1993. Accessory phase controls on the geochemistry of crustal

melts and restites produced during water-undersaturated partial melting. Contributions

to Mineralogy and Petrology, 114(4): 550-566.


Journal Pre-proof

Weinberg, R.F. and Hasalová, P., 2015. Water-fluxed melting of the continental crust: A

review. Lithos, 212-215: 158-188.

Weinberg, R.F., Hasalová, P., Ward, L. and Fanning, C.M., 2013. Interaction between

deformation and magma extraction in migmatites: Examples from Kangaroo Island,

South Australia. Bulletin of the Geological Society of America, 125(7-8): 1282-1300.

Weinberg, R.F. and Regenauer-Lieb, K., 2010. Ductile fractures and magma migration from

source. Geology, 38(4): 363-366.

Weinberg, R.F., Veveakis, E. and Regenauer-Lieb, K., 2015. Compaction-driven melt

f
oo
segregation in migmatites. Geology, 43(6): 471-474.

Weinberg, R.F., Wolfram, L.C., Nebel, O., Hasalová, P., Závada, P., Kylander-Clark, A.R.C.
pr
and Becchio, R., 2020. Decoupled U-Pb date and chemical zonation of monazite in
e-
migmatites: The case for disturbance of isotopic systematics by coupled dissolution-
Pr

reprecipitation. Geochimica et Cosmochimica Acta, 269: 398-412.

Weller, O.M., Jackson, S., Miller, W.G., St‐Onge, M.R. and Rayner, N., 2020. Quantitative
al

elemental mapping of granulite‐facies monazite: Textural insights and implications


n

for petrochronology. Journal of Metamorphic Geology, 38(8): 853-880.


ur

Whalen, J.B., Currie, K.L. and Chappell, B.W., 1987. A-type granites: geochemical
Jo

characteristics, discrimination and petrogenesis. Contributions to mineralogy and

petrology, 95(4): 407-419.

Wheller, C.J. and Powell, R., 2014. A new thermodynamic model for sapphirine: Calculated

phase equilibria in K2O-FeO-MgO-Al2O3-SiO2-H2O-TiO2-Fe2O3. Journal of

Metamorphic Geology, 32(3): 287-299.

White, A. and Chappell, B., 1988. Some supracrustal (S-type) granites of the Lachlan Fold

Belt. Earth and Environmental Science Transactions of The Royal Society of

Edinburgh, 79(2-3): 169-181.


Journal Pre-proof

White, R. and Powell, R., 2002a. Melt loss and the preservation of granulite facies mineral

assemblages. Journal of Metamorphic Geology, 20(7): 621-632.

White, R., Powell, R. and Halpin, J., 2004. Spatially‐focussed melt formation in aluminous

metapelites from Broken Hill, Australia. Journal of Metamorphic Geology, 22(9):

825-845.

White, R., Powell, R., Holland, T. and Worley, B., 2000. The effect of TiO2 and Fe2O3 on

metapelitic assemblages at greenschist and amphibolite facies conditions: mineral

equilibria calculations in the system K2O-FeO-MgO-Al2O3-SiO2-H2O-TiO2-Fe2O3.

f
oo
Journal of Metamorphic Geology, 18(5): 497-512.

White, R.W., Palin, R.M. and Green, E.C.R., 2017. High-grade metamorphism and partial
pr
melting in Archean composite grey gneiss complexes. Journal of Metamorphic
e-
Geology, 35(2): 181-195.
Pr

White, R.W., Pomroy, N.E. and Powell, R., 2005. An in situ metatexite-diatexite transition in

upper amphibolite facies rocks from Broken Hill, Australia. Journal of Metamorphic
al

Geology, 23(7): 579-602.


n

White, R.W. and Powell, R., 2002b. Melt loss and the preservation of granulite facies mineral
ur

assemblages. Journal of Metamorphic Geology, 20(7): 621-632.


Jo

White, R.W. and Powell, R., 2010. Retrograde melt-residue interaction and the formation of

near-anhydrous leucosomes in migmatites. Journal of Metamorphic Geology, 28(6):

579-597.

White, R.W. and Powell, R., 2011. On the interpretation of retrograde reaction textures in

granulite facies rocks. Journal of Metamorphic Geology, 29(1): 131-149.

White, R.W., Powell, R. and Baldwin, J.A., 2008. Calculated phase equilibria involving

chemical potentials to investigate the textural evolution of metamorphic rocks.

Journal of Metamorphic Geology, 26(2): 181-198.


Journal Pre-proof

White, R.W., Powell, R. and Holland, T.J.B., 2001. Calculation of partial melting equilibria

in the system Na2O–CaO–K2O–FeO–MgO–Al2O3–SiO2–H2O (NCKFMASH).

Journal of Metamorphic Geology, 19(2): 139-153.

White, R.W., Powell, R. and Holland, T.J.B., 2007. Progress relating to calculation of partial

melting equilibria for metapelites. Journal of Metamorphic Geology, 25(5): 511-527.

White, R.W., Powell, R., Holland, T.J.B., Johnson, T.E. and Green, E.C.R., 2014a. New

mineral activity-composition relations for thermodynamic calculations in metapelitic

systems. Journal of Metamorphic Geology, 32(3): 261-286.

f
oo
White, R.W., Powell, R. and Johnson, T.E., 2014b. The effect of Mn on mineral stability in

metapelites revisited: New a-x relations for manganese-bearing minerals. Journal of

Metamorphic Geology, 32(8): 809-828.


pr
e-
White, R.W., Powell, R. and Phillips, G.N., 2003. A mineral equilibria study of the
Pr

hydrothermal alteration in mafic greenschist facies rocks at Kalgoorlie, Western

Australia. Journal of Metamorphic Geology, 21(5): 455-468.


al

White, R.W., Stevens, G. and Johnson, T.E., 2011. Is the crucible reproducible? Reconciling
n

melting experiments with thermodynamic calculations. Elements, 7(4): 241-246.


ur

Whitney, D.L. and Evans, B.W., 2010. Abbreviations for names of rock-forming minerals.
Jo

American mineralogist, 95(1): 185-187.

Whitney, J.A., 1988. The origin of granite: The role and source of water in the evolution of

granitic magmas. Bulletin of the Geological Society of America, 100(12): 1886-1897.

Williams, I.S., 2001. Response of detrital and zircon and monazite, and their U-Pb isotopic

systems, to regional metamorphism and host-rock partial melting, Cooma Complex,

Southeastern Australia. Australian Journal of Earth Sciences, 48(4): 557-580.


Journal Pre-proof

Williams, M.A., Kelsey, D.E., Baggs, T., Hand, M. and Alessio, K.L., 2018. Thorium

distribution in the crust: Outcrop and grain-scale perspectives. Lithos, 320-321: 222-

235.

Williams, M.L., Jercinovic, M.J. and Hetherington, C.J., 2007. Microprobe monazite

geochronology: understanding geologic processes by integrating composition and

chronology. Annu. Rev. Earth Planet. Sci., 35: 137-175.

Windley, B.F., 1996. The evolving continents. Oceanographic Literature Review, 8(43): 785.

Windley, B.F., Kusky, T. and Polat, A., 2020. Onset of plate tectonics by the Eoarchean.

f
oo
Precambrian Research.

Winkler, H.G.F., 1957. Experimentelle Gesteinsmetamorphose-I Hydrothermale


pr
Metamorphose karbonatfreier Tone. Geochimica et Cosmochimica Acta, 13(1): 42-
e-
54,IN3-IN4,55-69.
Pr

Winkler, H.G.F. and Von Platen, H., 1958. Experimentelle Gesteinsmetamorphose-II.

Bildung von anatektischen granitischen Schmelzen bei der Metamorphose von NaCl-
al

führenden kalkfreien Tonen. Geochimica et Cosmochimica Acta, 15(1-2): 91-112.


n

Winkler, H.G.F. and von Platen, H., 1960. Experimentelle gesteinsmetamorphose-III.


ur

Anatektische ultrametamorphose kalkhaltiger tone. Geochimica et Cosmochimica


Jo

Acta, 18(3-4): 294-316.

Winkler, H.G.F. and von Platen, H., 1961. Experimentelle Gesteinsmetamorphose-IV.

Bildung anatektischer Schmelzen aus metamorphisierten Grauwacken. Geochimica et

Cosmochimica Acta, 24(1-2): 48-69.

Wise, D.U., 1974. Continental margins, freeboard and the volumes of continents and oceans

through time, The geology of continental margins. Springer, pp. 45-58.


Journal Pre-proof

Wolf, M.B. and London, D., 1994. Apatite dissolution into peraluminous haplogranitic melts:

An experimental study of solubilities and mechanisms. Geochimica et Cosmochimica

Acta, 58(19): 4127-4145.

Wolf, M.B. and London, D., 1995. Incongruent dissolution of REE- and Sr-rich apatite in

peraluminous granitic liquids: differential apatite, monazite, and xenotime solubilities

during anatexis. American Mineralogist, 80(7-8): 765-775.

Wolf, M.B. and Wyllie, P.J., 1994. Dehydration-melting of amphibolite at 10 kbar: the

effects of temperature and time. Contributions to Mineralogy and Petrology, 115(4):

f
oo
369-383.

Wolf, M.B. and Wyllie, P.J., 1995. Liquid segregation parameters from amphibolite
pr
dehydration melting experiments. Journal of Geophysical Research, 100(B8): 15,611-
e-
15,621.
Pr

Wood, B.J. and Fraser, D.G., 1976. Elementary thermodynamics for geologists. Oxford

University Press, USA.


al

Wu, F.Y., Li, X.H., Zheng, Y.F. and Gao, S., 2007. Lu-Hf isotopic systematics and thier
n

applications in petrology. Acta Petrologica Sinica, 23(2): 185-220.


ur

Wyart, J. and Sabatier, G., 1959. Transformation des sédiments pélitiques à 800 C sous une
Jo

pression d'eau de 1 800 bars et granitisation. Bulletin de Minéralogie, 82(4): 201-210.

Wyllie, P. and Tuttle, O., 1964. Experimental investigation of silicate systems containing two

volatile components; Part 3, The effects of SO_3, P_2O_5, HCl, and Li_2O, in

addition to H_2O, on the melting temperatures of albite and granite. American journal

of Science, 262(7): 930-939.

Wyllie, P.J., 1971. Experimental limits for melting in the earth's crust and upper mantle. The

structure and physical properties of the Earth's crust, 14: 279-301.

Wyllie, P.J. and Tuttle, O.F., 1959. Synthetic carbonatite magma. Nature, 183(4663): 770.
Journal Pre-proof

Wyllie, P.J. and Tuttle, O.F., 1961. Experimental investigation of silicate systems containing

two volatile components; Part 2, The effects of NH 3 and HF, in addition to H 2 O on

the melting temperatures of albite and granite. American Journal of Science, 259(2):

128-143.

Wyllie, P.J. and Wolf, M.B., 1993. Amphibolite dehydration-melting: Sorting out the solidus,

Geological Society Special Publication, pp. 405-416.

Xiang, H., 2020. GeoPS: an interactive visualization tool for thermodynamic modeling of

phase equilibria.

f
oo
Xiang, H. and Connolly, J.A., GeoPS: an interactive visual computing tool for

thermodynamic modeling of phase equilibria. Journal of Metamorphic Geology.


pr
Yakymchuk, C., 2017a. Applying phase equilibria modelling to metamorphic and geological
e-
processes: Recent developments and future potential. Geoscience Canada, 44(1): 27-
Pr

46.

Yakymchuk, C., 2017b. Behaviour of apatite during partial melting of metapelites and
al

consequences for prograde suprasolidus monazite growth. Lithos, 274-275: 412-426.


n

Yakymchuk, C., 2019. On granites. Journal of the Geological Society of India, 94(1): 9-22.
ur

Yakymchuk, C., 2021. Migmatites. In: D. Alderton and S.A. Elias (Editors), Encyclopedia of
Jo

Geology (Second Edition). Academic Press, Oxford, pp. 492-501.

Yakymchuk, C. and Acosta-Vigil, A., 2019. Geochemistry of phosphorus and the behavior of

apatite during crustal anatexis: Insights from melt inclusions and nanogranitoids.

American Mineralogist, 104(12): 1765-1780.

Yakymchuk, C. and Brown, M., 2014a. Behaviour of zircon and monazite during crustal

melting. Journal of the Geological Society, 171(4): 465-479.

Yakymchuk, C. and Brown, M., 2014b. Consequences of open-system melting in tectonics.

Journal of the Geological Society, 171(1): 21-40.


Journal Pre-proof

Yakymchuk, C. and Brown, M., 2019. Divergent behaviour of Th and U during anatexis:

Implications for the thermal evolution of orogenic crust. Journal of Metamorphic

Geology, 37(7): 899-916.

Yakymchuk, C., Brown, M., Clark, C., Korhonen, F.J., Piccoli, P.M., Siddoway, C.S.,

Taylor, R.J.M. and Vervoort, J.D., 2015. Decoding polyphase migmatites using

geochronology and phase equilibria modelling. Journal of Metamorphic Geology,

33(2): 203-230.

Yakymchuk, C., Brown, M., Ivanic, T.J. and Korhonen, F.J., 2013. Leucosome distribution in

f
oo
migmatitic paragneisses and orthogneisses: A record of self-organized melt migration

and entrapment in a heterogeneous partially-molten crust. Tectonophysics, 603: 136-

154.
pr
e-
Yakymchuk, C., Clark, C. and White, R.W., 2017. Phase relations, reaction sequences and
Pr

petrochronology. Reviews in Mineralogy and Geochemistry, 83(1): 13-53.

Yakymchuk, C., Kirkland, C.L. and Clark, C., 2018. Th/U ratios in metamorphic zircon.
al

Journal of Metamorphic Geology, 36(6): 715-737.


n

Yakymchuk, C., Kirkland, C.L., Hollis, J.A., Kendrick, J., Gardiner, N.J. and Szilas, K.,
ur

2020. Mesoarchean partial melting of mafic crust and tonalite production during high-
Jo

T–low-P stagnant tectonism, Akia Terrane, West Greenland. Precambrian Research,

339.

Yakymchuk, C., Rehm, A., Liao, Z. and Cottle, J.M., 2019a. Petrochronology of oxidized

granulites from southern Peru. Journal of Metamorphic Geology, 37(6): 839-862.

Yakymchuk, C., Zhao, W., Wan, Y., Lin, S. and Longstaffe, F.J., 2019b. Fluid-present

anatexis of Neoarchean tonalite and amphibolite in the Western Shandong Province.

Lithos, 326-327: 110-124.


Journal Pre-proof

Yardley, B.W. and Valley, J.W., 1997. The petrologic case for a dry lower crust. Journal of

Geophysical Research: Solid Earth, 102(B6): 12173-12185.

Yardley, B.W.D. and Barber, J.P., 1991. Melting reactions in the Connemara Schists: the role

of water infiltration in the formation of amphibolite facies migmatites. American

Mineralogist, 76(5-6): 848-856.

Yoder, H.S. and Tilley, C.E., 1962. Origin of basalt magmas: An experimental study of

natural and synthetic rock systems. Journal of Petrology, 3(3): 342-532.

Yuan, X., Sobolev, S.V., Kind, R., Oncken, O., Bock, G., Asch, G., Schurr, B., Graeber, F.,

f
oo
Rudloff, A. and Hanka, W., 2000. Subduction and collision processes in the Central

Andes constrained by converted seismic phases. Nature, 408(6815): 958-961.


pr
Zen, E.-A., 1966. Construction of pressure-temperature diagrams for multicomponent
e-
systems after the method of Schreinemakers: A geometric approach, 1225. US
Pr

Government Printing Office.

Zeng, L., Saleeby, J.B. and Ducea, M., 2005. Geochemical characteristics of crustal anatexis
al

during the formation of migmatite at the Southern Sierra Nevada, California.


n

Contributions to Mineralogy and Petrology, 150(4): 386-402.


ur

Zhao, Y., Li, N., Jiang, Y. and Niu, H., 2017. Petrogenesis of the Late Archean (∼2.5 Ga)
Jo

Na- and K-rich granitoids in the Zhongtiao-Wangwu region and its tectonic

significance for the crustal evolution of the North China Craton. Precambrian

Research, 303: 590-603.

Zheng, Y.-F., 2012. Metamorphic chemical geodynamics in continental subduction zones.

Chemical Geology, 328: 5-48.

Zheng, Y.-F., Gao, X.-Y., Chen, R.-X. and Gao, T., 2011. Zr-in-rutile thermometry of

eclogite in the Dabie orogen: Constraints on rutile growth during continental

subduction-zone metamorphism. Journal of Asian Earth Sciences, 40(2): 427-451.


Journal Pre-proof

Zheng, Y.-F. and Hermann, J., 2014. Geochemistry of continental subduction-zone fluids.

Earth, Planets and Space, 66(1): 93.

Zheng, Y.F., Chen, R.X. and Zhao, Z.F., 2009. Chemical geodynamics of continental

subduction-zone metamorphism: Insights from studies of the Chinese Continental

Scientific Drilling (CCSD) core samples. Tectonophysics, 475(2): 327-358.

Zhong, X., Petley-Ragan, A.J., Incel, S.H., Dabrowski, M., Andersen, N.H. and Jamtveit, B.,

2021. Lower crustal earthquake associated with highly pressurized frictional melts.

Nature Geoscience: 1-7.

f
oo
Zuluaga, C.A., Stowell, H.H. and Tinkham, D.K., 2005. The effect of zoned garnet on

metapelite pseudosection topology and calculated metamorphic P-T paths. American

Mineralogist, 90(10): 1619-1628.


pr
e-
Pr

Figure captions
n al

Fig. 1. In (a), the coloured data points show P–T conditions for 564 localities from the global
ur

age, temperature and pressure dataset for secular change in metamorphism (Brown and
Jo

Johnson, 2019a). The data are grouped into three types of metamorphism: low T/P (blue),

intermediate T/P (orange) and high T/P (red). Superimposed on the data in (a) are the

metamorphic facies. At pressures below coesite stability the facies boundaries are transitional

to indicate the control of bulk composition on the change in mineral assemblages from one

facies to another. Low-to-moderate pressure facies are: L = low-grade metamorphism,

including the zeolite facies; Gs = greenschist facies; A = amphibolite facies; and, G =

granulite facies (including ultrahigh temperature metamorphism at T > 900 °C). High-to-

ultrahigh pressure facies are: Bs = blueschist facies; Ec = eclogite facies; and, UHP =
Journal Pre-proof

ultrahigh pressure metamorphism, which is the part of the eclogite facies at pressures where

coesite rather than quartz is stable. Progressive metamorphism across a sequence of facies is

a metamorphic facies series. In (b), we show the range of thermobaric ratios (T/P) defined by

each type of metamorphism: low T/P, < 375 °C/GPa; intermediate T/P, 375 to 775 °C/GPa;

and, high T/P, > 775 °C/GPa. In (c), we plot T/P vs age to show the evolution of bimodality

in the Paleoproterozoic and the dominance of lower thermobaric ratios since the

Neoproterozoic. Contours, which are for the dataset as a whole, represent the proportion of

the data inside 1 sd, 2 sd and 3 sd, respectively.

f
oo
Fig. 2. Evidence of partial melting at outcrop and nomenclature of migmatites. (a). Fluid-
pr
present partial melting of amphibolite where the melt has migrated to form leucosome-
e-
dominated patches leaving behind coarse-grained amphibole-dominated areas of residue
Pr

within a medium-grained amphibolite protolith. (b). Incongruent partial melting in which

peritectic garnet, a solid product of the melting reaction, is located within the leucosome
al

veins, suggesting that the locations of garnet nucleation control the concentration of melt at
n

these sites. (c). Stromatic metatexite migmatite at the top becomes disrupted across a one-
ur

meter transition to schlieric diatexite migmatite. Note the paleosome within the transition
Jo

zone and the scholle of paleosome at the bottom-right within the diatexite. (d). Outcrop

comprising areas of patch metatexite migmatite and areas where the patches have merged to

form a dilatant metatexite migmatite. (e). Stromatic metatexite migmatite; the thin dark-

coloured mica-rich layers at the edges of the leucosomes are sometimes called mafic

selvedges. (f). Net-structured metatexite migmatite; this structure is common in

homogeneous protoliths. Migmatite nomenclature follows Sawyer (2008).


Journal Pre-proof

Fig. 3. Microstructural evidence of partial melting and melt crystallisation in thin section. (a).

Backscattered electron image to show H2O-saturated partial melting of muscovite (Ms) +

quartz. An early stage of melting is shown in which a layer of former melt (L) developed

along the muscovite–quartz grain boundary with the localised nucleation of mullite (M).

Small grains of biotite (high contrast) are also present. Reproduced from D.C. Rubie and A.J.

Brearley (1990, A Model for rates of disequilibrium melting during metamorphism. In:

Ashworth, J.R. and Brown, M., 1990, High‐temperature Metamorphism and Crustal

Anatexis. Unwin Hyman, pp. 57–86) with permission. (b). Optical image to show the

f
oo
occurrence of K-feldspar as thin interstitial cuspate films with low dihedral angles along

grain boundaries of phengite and quartz (cross-polarized light). (c). Cordierite-rich mantle
pr
around large garnet (at bottom). The compacted microstructure of the cordierite suggests melt
e-
drainage, but the cuspate interstitial quartz around the small garnet represents a pseudomorph
Pr

of a trapped pocket of melt. Note the low dihedral angles of the quartz against the

surrounding cordierite.
n al

Fig. 4. Microstructural evidence of incongruent partial melting in thin section of eclogite. (a)
ur

and (c). Plane polarized light; (b) and (d). Crossed polarized light with 1 λ plate. Networks of
Jo

optically-continuous plagioclase pseudomorph former grain boundary films of melt (pale

orange in (b) and (d)). Although the eclogite contains both phengite and zoisite, melting was

predominantly by omphacite breakdown with only a minor contribution from these two

minerals plus quartz, yielding omphacite with a lower concentration of the Na–Al pyroxene

(jadeite) component than the large grains of Na–Al-rich pyroxene (omphacite) and melt

(Feng et al., 2021). Microstructural evidence of this reaction is seen on the left and at the

bottom right in (a) and (b), and in the centre of (c) and (d), where large omphacite grains

(lavender or blue in (b) and blue in (d)) are partially replaced by finer-grained omphacite and
Journal Pre-proof

optically-continuous plagioclase. Subhedral hornblende has crystallized from the melt (e.g.

lower left in (d)). A thin layer of plagioclase (blue) pseudomorphs melt along an omphacite–

garnet grain boundary in the upper left in (c) and (d), where the omphacite is seen to have a

ragged grain boundary against plagioclase but the garnet (purple in (d)) retains a straight

grain boundary (white oval in (c)), consistent with no contribution from garnet to the melting

reaction. The scale bar is 500 m wide.

f
Fig. 5. Leucogranite forming a sill and dike network in stromatic metatexite migmatite at

oo
Maigetter Peak (height 480m) in the Fosdick Mountains of West Antarctica (76°26’38”S,

146°30’00”W). The view is to the southeast and was taken from the air (Twin Otter wing tip
pr
in upper right). From the aerial perspective and also on close examination in outcrops,
e-
intersecting sills and dikes do not appear to truncate or displace each other. Furthermore, the
Pr

sills and dikes of leucogranite crosscut foliation but may be either continuous with or

discordant to garnet and cordierite-bearing leucosomes in the migmatites. Thus, during


al

anatexis this level in the crust was both a source of melt and a melt transfer zone.
n
ur

Fig. 6. P–T diagram showing the stability of the aluminosilicate minerals andalusite, kyanite
Jo

and sillimanite in the one-component Al2SiO5 system. The various points and paths are

discussed in the text.

Fig. 7. Hypothetical isobaric diagrams illustrating phase equilibria in the binary system A–B

in which no phase shows solid solution. (a) simple system with no intermediate phases; (b)

system with an intermediate phase AB that melts congruently; (c) system with an

intermediate phase AB that melts incongruently. See text for discussion. A more detailed
Journal Pre-proof

treatment of these diagrams and of phase equilibria in more complex ternary and quaternary

systems is provided by (Cox et al., 1979).

Fig. 8. Qualitative phase relations in the KASH system. (a) Compatibility triangle in the K–

A–H system projected from quartz (S) showing the compositions of the phases considered.

The only phase showing solid solution is melt that contains variable H2O contents. (b–f)

Sequential development of a Schreinemakers’ bundle showing the stable (solid red) and

metastable (dotted red) segments of univariant equilibria around the invariant point (IP) at

f
oo
which all six phases coexist. Also shown each panel is the compatibility diagram appropriate

for that univariant equilibria, from which the reaction can be deduced. See text for more

details.
pr
e-
Pr

Fig. 9. Calculated phase diagrams based on the experimental starting composition of Taylor

et al. (2015b) in the NKFMASH system to illustrate the features of, and logic behind,
al

pseudosections. (a) P–T pseudosection. The red dotted lines are for a constant abundance
n

(mode isopleths in mol.% ~vol.%) of biotite. The inset shows the changing variance (F) of
ur

the four fields within the black dashed box; (b) isobaric (P = 0.9 GPa) T–X pseudosection, in
Jo

which X shows XMg varying from 0.1 to 0.9. See text for more details. The white dashed lines

in (a) and (b) are equivalent. (c) Modebox diagram showing the changing abundance (as

mol.% ~vol.%) along the isobaric heating path between A and B shown in (a). See the main

text or more details.

Fig. 10. Modebox diagrams illustrating the contrasting behaviour during open vs

conditionally closed system melting of a model metapelite (a) and metagreywacke (b)

(Yakymchuk and Brown, 2014b). See the text for more details.
Journal Pre-proof

Fig. 11. Field relations of migmatised metabasic (a–d) and metatonalitic (e and f) rocks from

the Lewisian Gneiss Complex, NW Scotland, illustrating the generation, segregation and

migration of TTG-like melts/magmas. (a) Metatexite containing irregular patch of peritectic

Cpx-bearing leucosome; (b) Stromatic metatexite comprising discontinuous stroma of

leucosome that interconnect with coarser-grained patches, some containing retrogressed

peritectic Cpx grains; (c) A peritectic Cpx-bearing sheet (bottom) feeds into a trondhjemite

sheet (top). The Cpx-bearing sheet is associated with a diffuse mafic selvedge and

f
oo
interconnects with thin veins of leucosome within the melanosome. (d) Large peritectic Cpx

grains within a leucosome channel are surrounded by a thin rind of hornblende, suggesting
pr
limited retrograde reaction with melt. For more information see Johnson et al. (2012) and
e-
Feisel et al. (2018). (e) and (f) Stromatic leucosomes in petrographic continuity with a
Pr

coarse-grained cross-cutting leucosome within a shear band. The cross-cutting leucosome

contains abundant peritectic pyroxene consistent with fluid-absent incongruent melting


al

reactions consuming hornblende.


n
ur

Fig. 12. Calculated phase relations for the average Coucal basalt in the 10-component Na2O–
Jo

CaO–K2O–FeO–MgO–Al2O3–SiO2–H2O–TiO2–O (NCKFMASHTO) chemical system. (a)

P–T pseudosection using an Fe3+/Fe of 0.1 and an H2O content just sufficient to saturate the

solidus at 1.0 GPa. The red dotted lines show melt proportions (F) as mol.% (~vol.%). The

thick white dashed line shows the a 900 °C/GPa thermobaric gradient and the thinner dashed

lines the 700 and 1100 °C/GPa thermobaric gradients discussed in detail in the main text (see

also Fig. 14). (b) Simplified version of (a) emphasising some key features. At low melt

fractions (<10%) the melts are granitic (high K) in composition. At higher melt fractions, the

yellow, green and blue fields show the stability of low-, medium- and high-P TTGs
Journal Pre-proof

according to (Moyen, 2011), which are inferred to record the presence or absence of garnet

and rutile. The white dashed lines are linear geotherms (°C/GPa), with the 900 °C/GPa

thermobaric ratio emboldened. Modified after Johnson et al. (2017).

Fig. 13. Calculated modebox diagrams showing the changing abundance of phases with

heating along the 700 °C/GPa (a), 900 °C/GPa (b) and 1100 °C/GPa (c) thermobaric ratios,

which correspond to relatively low, moderate and high geotherms, respectively. (d)

Calculated major oxide composition of melt produced from the average Coucal basalt along

f
oo
the 900 °C/GPa thermobaric gradient at various melt fractions (5–40%) normalised to the

average Paleoarchaean East Pilbara TTG. The modelled melt compositions at a melt fraction
pr
of 25% fit within the 2 envelope on the average Paleoarchaean East Pilbara TTG data,
e-
which is indicated by the grey shaded region. Also shown are the global average
Pr

compositions of TTGs and potassic granites (Moyen, 2011). Modified from Johnson et al.

(2017).
n al

Fig. 14. P–T pseudosection from 0.35 to 1.35 GPa and 640–1000 °C based on the hydrated
ur

compositions of an average: (a) MORB (with plagioclase in excess; (Sun and McDonough,
Jo

1989)) and; (b) Eoarchean basalt from Isua, Greenland (with hornblende in excess;

(Hoffmann et al., 2011)). Both modelled compositions contain sufficient H2O to minimally

saturate the wet solidus at 1.3 GPa. The thin red dashed lines in (b) show melt proportions (as

mol.% ~vol.%). Calculations use the end-member thermodynamic data of (Holland and

Powell, 2011) and activity–composition models of Green et al. (2016) and White et al.

(2014a). For more details see Gardiner et al. (2018), Gardiner et al. (2019a) and the main

text.
Journal Pre-proof

Fig. 15. (a) P–T pseudosection from 0.15 to 1.0 GPa and 640–900 °C and (b) isobaric (P =

0.8 GPa) T–X pseudosection for a hornblende-gabbro sample 87S35a used in the experiments

of Sisson et al. (2005), with hornblende and plagioclase in excess. In (a), the bulk

composition contains sufficient H2O to minimally saturate the wet solidus at 0.9 GPa. In (b),

the x-axis illustrates the influence of adding H2O (up to a total of 6 wt%) at temperatures of

650–800°C at a pressure equivalent to the base of thickened arc crust. The thin red dashed

lines in (b) show melt proportions (as mol.% ~vol.%). Calculations use the end-member

thermodynamic data of (Holland and Powell, 2011) and activity–composition models of

f
oo
Green et al. (2016) and White et al. (2014a). For more details see Collins et al. (2020b) and

the main text.


pr
e-
Fig. 16. Simplified A(K)FM diagram projected from K-feldspar showing the garnet–
Pr

cordierite tie line that is generally considered to separate aluminous from subaluminous

metapelites. Various average compositions of shales, metapelites and greywackes are also
al

shown. S56, Shaw (1956); A91, Ague (1991); C93, Condie (1993); YB14, Yakymchuk and
n

Brown (2014b); NASC, North Atlantic Shale Composite; PAAS, average post-Archean
ur

Australian shale.
Jo

Fig. 17. Petrogenetic grid in the K2O–FeO–MgO–Al2O3–SiO2–H2O (KFMASH) system

showing suprasolidus univariant reactions. Modified from White et al. (2014a). Simplified

A(K)FM diagrams showing co-existing mineral assemblages are modified from Spear et al.

(1999). The numbered reactions are discussed in the text.

Fig. 18. Suprasolidus phase relationships for an average amphibolite-facies metapelite

composition (Ague, 1991). (a) P–T pseudosection (isochemical phase diagram) in the MnO–
Journal Pre-proof

Na2O–CaO–K2O–FeO–MgO–Al2O3–SiO2–H2O–TiO2–O (MnNCKFMASHTO) system

(modified from Yakymchuk (2017a)). All fields include liquid and ilmenite. For clarity, not

all of the small multivariant fields are labelled, but are provided in Yakymchuk (2017a). The

H2O content was chosen so that the bulk composition does not contain any free H2O at the

solidus at 0.7 GPa. The numbers in circles indicate reactions equivalent to the numbered

KFMASH equilibria given in the text. (b) Simplified P–T diagram showing the stability

fields of key minerals. (c) Simplified P–T diagram showing isopleths for the proportion of

melt in the system. Note that the H2O content was adjusted for calculations relevant to this

f
oo
diagram so that there is no free H2O at 0.6 GPa. Consequently, quartz is stabilized to slightly

higher temperatures than in (a) and spinel stability is slightly decreased.


pr
e-
Fig. 19. Modebox diagrams showing the changing abundance (as mol.% ~vol.%) of phases
Pr

along an isobaric (P = 0.6 GPa) heating path for an average metapelite (a) and

metagreywacke (b). See text for more details. Note that the H2O content was adjusted such
al

that there is no free H2O at 0.6 GPa, which is different than the scenario of no free H2O at 0.7
n

GPa for the P–T diagrams in Fig. 19a (metapelite) and Fig. 21a (greywacke). Consequently,
ur

the stability fields are slightly different between diagrams for the same rock type. For the
Jo

metapelite, quartz is stabilized to slightly higher temperatures than in Fig. 19a and spinel

stability is slightly decreased. For the greywacke, the differences are negligible.

Fig. 20. Suprasolidus phase relationships for an average passive margin greywacke

(Yakymchuk and Brown, 2014b). (a) P–T pseudosection (isochemical phase diagram) in the

MnO-Na2O-CaO-K2O-FeO-MgO-Al2O3-SiO2-H2O-TiO2-O (MnNCKFMASHTO) system

(modified from (Yakymchuk, 2017a). All fields include quartz, liquid and ilmenite. The H2O

content of the bulk composition was chosen such that there is no free H2O at the solidus at
Journal Pre-proof

0.7 GPa. The numbers in circles indicate reactions equivalent to the numbered KFMASH

equilibria given in the text. (b) Simplified P–T diagram showing the stability fields of key

minerals. (c) Temperature–proportion (modebox) diagram showing the changes in predicted

mineral proportions during isobaric heating at 0.6 GPa. Note that the H2O content was

adjusted in this diagram so that there is no free H2O at 0.6 GPa. (d) One-oxide normalized

molar percentage (approximately equivalent to vol.%) of melt.

Fig. 21. Modelled melt compositions from a metapelite (a–c) and greywacke (d–f).

f
oo
Concentrations of SiO2 in melt are generally predicted to increase with heating (a, d) except

after quartz exhaustion in the metapelite (a). A/CNK (molar Al2O3 / [Na2O + K2O + CaO])
pr
increases with heating. The inflection in the predicted A/CNK values coincides with the
e-
exhaustion of sillimanite. The Fe# (molar Fe / [Fe + Mg]) of the melt decreases with
Pr

temperature until biotite exhaustion in both rock types (c, f) and then increases.
al

Fig. 22. Range of melt compositions from 0.4–1.2 GPa and the solidus to 950C for an
n

average amphibolite-facies metapelite and an average passive margin greywacke. The fields
ur

enclose all of the melt compositions for each rock type over these P–T conditions. (a) A/NK
Jo

versus A/CNK diagram to differentiate peralkaline, metaluminous and peraluminous melt

compositions (Shand, 1927). Both modelled compositions are predicted to generate

peraluminous melt and strongly peraluminous (Clarke, 2019) melt compositions. (b)

CaO/Na2O versus A/CNK diagram showing the predicted melt compositions (fields). The

horizontal dashed line at CaO/Na2O = 0.3 is commonly used to differentiate melt generated

from metapelitic (<0.3) and metapsammitic (>0.3) sources (Sylvester, 1998). Note the

overlap between the two sources and the inconsistency with the commonly used boundary for

separating metapelite- and metapsammite-derived melt compositions.


Journal Pre-proof

Fig. 23. Simplified P–T phase diagram for rock-buffered silicate mineral–H2O systems. The

thick red curve between the two black lines is the low P wet solidus for compositions

between haplogranite (no CaO; lower-T line, S1) and Ca-bearing granite (CaO of 1.93 wt%,

plagioclase An20; higher-T line, S2), respectively, which terminates at the second critical

endpoints SCE 1 (haplogranite) and SCE 2 (Ca-bearing granite) with the critical lines (CL 1

and CL 2) extending to lower P and higher T (Hack et al., 2007). The critical lines separate

the phase assemblage field of silicate minerals plus supercritical fluid (SCF) at high P from

f
oo
that of silicate minerals plus hydrous melt and aqueous fluid (M + F) at low P. The solid

curve extending to high P from SCE 2 is the isopleth of 30 wt% solute in the fluid phase,
pr
which separates the P–T field of aqueous fluid (at low T) from that of supercritical fluid (at
e-
high T) in equilibrium with silicate minerals (purple dashed lines on either side of 30 are
Pr

isopleths of solute content in fluid (at lower T) and supercritical fluid (at higher T),

respectively). Also shown is the wet solidus for basalt (S 3), which terminates at the second
al

critical endpoint SCE 3 with the critical line CL 3 separating the phase assemblage field of
n

silicate minerals plus supercritical fluid (SCF) at high P from that of silicate minerals plus
ur

hydrous melt and aqueous fluid (M + F) at low P. Solid green lines show the high-P stability
Jo

limit of albite, and the transformation of quartz to coesite and graphite to diamond.

Fig. 24. Modelled stability of zircon and monazite in an average metapelite (a, b) and an

average passive margin greywacke (c, d) during closed-system melting. The diagram was

constructed by extracting mineral/melt modes and compositions over a grid every 25 °C and

0.05 GPa and contouring the diagram at this resolution based on the solubility of zircon and

monazite, using the solubility models of Boehnke et al. (2013) and Stepanov et al. (2012).

Solid black lines show the amount of zircon and monazite remaining in the residue relative to
Journal Pre-proof

the amount at the solidus (dashed black line) for a scenario with whole-rock concentrations of

150 ppm Zr and 150 ppm LREE. The solid blue lines show the limit of monazite and zircon

stability for the specified concentrations of Zr and the LREE. The background colours show

the stability of the major minerals for the metapelite (see Fig. 19) and the greywacke (see Fig.

21). For all scenarios, the amounts of zircon and monazite are expected to decrease with

increasing temperature. For more details see Yakymchuk and Brown (2014a) and

Yakymchuk et al. (2017).

f
oo
Fig. 25. Concentrations of trace elements in accessory minerals as a function of pressure and

temperature in an average amphibolite-facies metapelite (Ague, 1991). (a) Chondrite-


pr
normalized Gd/Yb ratio of zircon in equilibrium with the major phase assemblage using the
e-
partition coefficients of the REE in zircon from the Taylor et al. (2015b) lattice-strain model
Pr

and the zircon solubility expression of Boehnke et al. (2013). The coloured fields are the

same as in Fig. 19b. (b) Modelled concentration of Y2O3 in monazite as a function of


al

pressure and temperature using the partition coefficients determined from the C3308
n

experimental run of Stepanov et al. (2012) combined with their monazite solubility
ur

expression. (c) Modelled proportions of zircon and monazite (relative to the amount at the
Jo

solidus) and of garnet for an isobaric heating path at 0.6 GPa. The proportions of zircon and

monazite decrease non-linearly with temperature whereas the quantity of garnet increases. (d)

Modelled concentrations of REE in zircon along a 0.6 GPa isobaric heating path at 25 °C

intervals. (e) Modelled concentrations of REE in monazite along a 0.6 GPa isobaric heating

path at 25 °C intervals. Line colours in (d) and (e) represent the amount of garnet in the

system at the modelled P–T condition. For all models, partition coefficients of ilmenite,

rutile, apatite, plagioclase and biotite were taken from Bédard (2006b); cordierite, K-feldspar,

quartz, sillimanite and kyanite were assumed to host negligible REE and Y. The bulk
Journal Pre-proof

composition of the metapelite was assumed to be that of the average Proterozoic shale of

Condie (1993) except for Zr, which was assumed to be 150 ppm. All chondrite values are

from McDonough and Sun (1995). Apatite solubility was modelled with the expression of

Pichavant et al. (1992).

f
oo
pr
e-
Pr
nal
ur
Jo
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12
Figure 13
Figure 14
Figure 15
Figure 16
Figure 17
Figure 18
Figure 19
Figure 20
Figure 21
Figure 22
Figure 23
Figure 24
Figure 25

You might also like