Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Melted Rocks under the

Microscope: Microstructures
and Their Interpretation
Marian B. Holness1, Bernardo Cesare2 and Edward W. Sawyer3

1811-5209/11/0007-0247$2.50 DOI: 10.2113/gselements.7.4.247

R
ecognising the former presence of melt in rocks which have undergone MICROSTRUCTURES
cooling and exhumation over millions of years following regional meta- DUE TO MELTING
morphism commonly relies on the correct interpretation of grain-scale In the first instance, the identifica-
tion and interpretation of micro-
structures visible only under the microscope. The evolution of these structures
structures associated with partial
during prograde melting and, later, retrograde cooling can be understood melting involves comparison of
using concepts derived from experimental simulation and materials science. natural examples with experi-
mental simulations. Experiments
KEYWORDS : migmatite, microstructure, crystallization, dihedral angle, textural provide the opportunit y to
equilibration, mineral reactions, retrograde metamorphism constrain rock composition and to
control the pressure and tempera-
ture conditions, but a major draw-
back is that they can never be run
INTRODUCTION for sufficiently long periods to truly emulate geological
It can be relatively straightforward to deduce whether rocks events. However, we can melt rocks under laboratory condi-
have been subjected to partial melting. Field geologists tions to get an idea about what melting may look like,
look for high-temperature mineral assemblages and the at least on short timescales. We fi nd that melting always
presence of lenses and irregular patches of quartzofeld- initiates at the junctions between reactant grains and
spathic material (leucosomes; see Glossary on page 234) commonly forms fi lms of melt separating these reacting
with a bulk composition consistent with derivation from minerals (Acosta-Vigil et al. 2006).
an anatectic melt. Melting is commonly associated with
regional deformation which results in coalescence of these The next step towards understanding longer-duration
patches. The signature of the former presence of melt on events is to examine natural contexts in which rocks have
a smaller, microscopic scale is not always so obvious, and been melted and then cooled so quickly that the melt is
this signature is subject to modification during both the preserved as glass. The most extreme examples of this
melting event itself and the subsequent history of the rock. are pseudotachylites, formed during movement on fault
surfaces by frictional heating, but we are most interested
Spry (1969) took a significant step forward in understanding in examples in which deformation was less intense and
the development of metamorphic rocks by applying a the heat source was either igneous or radioactive heating
materials science approach to interpreting their micro- in over-thickened crust. Examples of these are found in
structures. At that time an anatectic origin for migmatites two, rather rare, environments. The fi rst is where rocks
was not widely accepted, and over the next two decades have been subjected to pyrometamorphism, defi ned as a
research on migmatite microstructures was motivated by short (10–1000 years) and very hot metamorphic event. The
the need to disprove the hypothesis that leucosomes were speed at which the rocks are brought up to their melting
subsolidus segregations and to fi nd evidence that they had point is generally matched by the speed at which they are
instead crystallized from melt (e.g. Vernon and Collins cooled back down again. Such conditions occur in the walls
1988). Subsequently it was shown that microstructures in of shallow magma conduits feeding major surface flows.
the material surrounding the leucosomes also record the Melt in the wall rocks is generated at the contacts between
presence of melt during the metamorphic peak (Sawyer reactant grains (FIG. 1A) to form parallel-sided fi lms that
1999). Recent interest in migmatite microstructures has thicken with time (Holness et al. 2005). The fi lms lose
taken another turn, this time focussed on what happened their continuity and parallelism when the melt proportion
inside the small former melt pockets. Research has now becomes sufficiently large for the remaining solid grains
revealed them to be a fascinating micro-world controlled to move relative to each other. For melting reactions that
by kinetic factors. involve a volume increase, overpressure creates a network
of melt-fi lled fractures (FIG. 1B) (Holness and Watt 2002).
This network doesn’t seem to provide a good pathway for
melt migration on short timescales, but it is possible that
1 Department of Earth Sciences, University of Cambridge it might play an important role during longer-lived events.
Downing Street, Cambridge CB2 3EQ, UK
E-mail: marian@esc.cam.ac.uk Other naturally quenched melted rocks are fragments,
2 Dipartimento di Geoscienze, Università di Padova or enclaves, of metasedimentary rock caught up in lava.
Via Gradenigo 6, I-35131 Padova, Italy Although metasedimentary xenoliths are comparatively
E-mail: bernardo.cesare@unipd.it common (Braun and Kriegsman 2001; Grapes and Li
3 Sciences de la Terre, Département des Sciences Appliquées 2010), the suite of metasedimentary enclaves erupted by
Université du Québec à Chicoutimi, Chicoutimi the volcano at El Hoyazo in southeastern Spain is very
Québec G7H 2B1, Canada unusual (FIG. 2A). They reached their metamorphic peak
E-mail: ewsawyer@uqac.ca

E LEMENTS , V OL . 7, PP. 247–252 247 A UGUS T 2011


A B

(A) Parallel-sided films of brown glass develop on are the dark elongate patches formed of a fine-grained mineral
FIGURE 1 grain boundaries between the reactant phases quartz aggregate representing solidified former melt. They are connected
(Qtz, clear) and feldspar (Fsp, dusty) during pyrometamorphism in to each other by cracks (representative examples are arrowed),
the walls of a magma conduit that was active for 5 months. Mull, filled with fine-grained material solidified from the former melt and
western Scotland. (B) A partially melted muscovite schist, with the oriented diagonally across the image.
original foliation oriented horizontally. The reacted muscovite grains

at 15 to 20 km depth a few million years before they were energies associated with interfaces of all kinds, such as
erupted (Cesare et al. 2009a). This is consistent with their grain boundaries and fluid–solid interfaces) are reduced.
involvement in a melting event associated with the volca- These conditions are generally met in the deep crust.
nism. Their composition is exactly what we might expect Textural equilibrium is easily recognized in melt-free rocks
the residue to be after anatexis and extraction of up to by the uniform grain size and the smoothly curved grain
60 wt% of melt (although the extraction was incomplete boundaries, which meet at angles reflecting the relative
as the enclaves still contain abundant melt inclusions and magnitude of the different interfacial energies (FIG. 3A).
some interstitial melt). There is therefore the exciting possi-
bility that the enclaves may be fragments of the crustal The angle formed at the corners of melt-fi lled pores is
source of their host lava. Cesare et al. (1997) suggested that controlled by the relative magnitude of the fluid–solid
the microstructures in the enclaves were formed during a and grain-boundary energies and is known as the dihe-
rapid heating event caused by crustal thinning, followed dral angle (see FIG. 3B for an example of the dihedral angle
by a long (106 years) period at supersolidus temperatures. formed at a three-grain junction). Because the interfacial
After entrainment, the microstructures formed during this energy depends on the orientation of the crystal lattice,
melting event were quenched during eruption (Cesare et the equilibrium melt–solid dihedral angle depends on the
al. 1997; Acosta-Vigil et al. 2010). If this interpretation is relative orientations of the two grains forming the pore
correct, then the El Hoyazo enclaves provide us with a corner. This results in a range of equilibrium angles, typi-
missing link between pyrometamorphic aureoles evolving cally with a standard deviation of 10–15°(Holness 2006).
on a 101–103 -year timescale and migmatites formed during The dihedral angle controls melt connectivity in texturally
regional metamorphism on a 106 –107-year timescale. equilibrated materials: if the melt–solid dihedral angle is
The melt in the enclaves, now solidified to glass, has a less than 60° the melt forms a stable interconnected network
rather different distribution to that seen in the pyrometa- of channels along three-grain junctions (FIG. 3C), but if the
morphic aureoles or the more common type of metasedi- angle is greater than 60°, melt forms isolated pockets on
mentary xenoliths entrained in the shallow crust. It is four-grain junctions. This is of immense significance for
no longer confi ned to the sites of reaction but is present the properties of the melt-bearing rock as the mobility of
throughout the rock. Films of glass are present on some the melt and the strength of the partially melted material
grain boundaries (FIG. 2 B ), notably surrounding garnet are therefore directly related to the dihedral angle. The
grains, but glass also forms lenses parallel to the dominant dihedral angle for virtually all silicate mineral–melt pairs is
foliation in enclaves containing abundant biotite. <60° (Holness 2006); higher angles are seen for non-silicate
liquids, such as the Fe–Ni liquids relevant to core-forming
The El Hoyazo enclaves show us what melting might have events (e.g. Terasaki et al. 2005), and for some orientations
looked like during anatexis, but the preservation of these of highly anisotropic minerals like biotite (Laporte and
microstructures is extremely unusual. The great majority Watson 1995). Therefore in statically melted rocks, a silicate
of regional metamorphic rocks that underwent melting melt is generally completely interconnected by means of
cooled more slowly, and the long timescales involved in a stable network of channels along three-grain junctions,
the entire heating–cooling cycle complicate interpretation even when only a few percent of melt is present.
of melt-related microstructures. These long time periods
provide opportunities for modification of the microstruc- THE EFFECTS OF DEFORMATION
ture during and after solidification, driven by a combina-
tion of textural equilibration, reaction and deformation. Field evidence suggests that deformation is the rule rather
than the exception in regional metamorphic terrains. Melt
has an enormous effect on rock strength, even when only
STATIC MELT-BEARING ROCKS a few volume percent is present. On length scales greater
In the absence of deformation and if the temperature is than the grain size, melt-bearing rocks tend to behave in
permitted to change only slowly (reducing the rate of a ductile manner, so even at high strain rates it is unlikely
melt production or solidification), melt-bearing rocks will that they will break to form major fractures. Conversely
approach textural equilibrium as internal energies (i.e. the effect deformation has on grain-scale melt distribu-

E LEMENTS 248 A UGUS T 2011


A A

B
B

Enclaves of a partially melted crustal source entrained


FIGURE 2 by erupting lavas. (A) A 10 cm long metasedimentary
enclave enclosed in dacitic lava. (B) Photomicrograph of a partially
melted metasedimentary rock; a red tint plate has been used so
that the glass (representing the quenched melt) appears magenta.
The glass forms films and lenses between the solid grains. The
broad band of glass running subhorizontally across the image
contains cuspate, brown, devitrification structures.

tion can be profound. For interfacial energies to control


melt geometry the critical question is whether textural
equilibration can keep pace with deformation. Deformation
predominantly by diffusive processes is sufficiently slow to
permit textural equilibration, but when deformation rates Textural equilibrium. (A) An epidote-bearing quartzite
FIGURE 3 in textural equilibrium. The rare epidote grains are
are higher melt topology begins to be controlled more by labeled Ep. (B) The junction between two plagioclase grains (clear)
the deformation itself (e.g. Marchildon and Brown 2002). and an augite grain (brown). The interfacial energies act in the
plane of each interface, tangential to each boundary at their inter-
Observations of naturally deformed metamorphic rocks section, and pull away from the three-grain junction. They are
show that grain-scale melt distribution falls into two denoted γpp (the energy of the plagioclase–plagioclase grain
types. In some rocks melt forms pockets aligned parallel boundary) and γap (the energy of the augite–plagioclase grain
to foliation (e.g. Sawyer 2001; Marchildon and Brown 2002; boundary). In textural equilibrium, the three forces balance,
creating a characteristic angle, Θ: this is the plagioclase–plagio-
Guernina and Sawyer 2003). This is also the case in the El clase–augite dihedral angle. (C) A mould of the fluid-filled pore
Hoyazo enclaves (FIG. 2B). But in others it forms elongate network on three-grain junctions, visible after the dissolution of
pockets on grain boundaries aligned at a high angle to the solid. PHOTOGRAPH BY R. H. G ERMAN
the foliation (FIG. 4) (Rosenberg and Riller 2000; Závada
et al. 2007). In the latter case, melt may also occupy intra-
dihedral angle is zero (i.e. the interfacial energy of the grain
granular fractures. These differences in orientation seem to
boundary is greater than that of two fluid–solid interfaces),
depend on the magnitude of the stress and the amount of
which is not the case for silicate systems. Deformation-
melt present. Foliation-parallel melt pockets form at rela-
controlled, completely wetted grain boundaries are there-
tively low stresses and with small volumes of melt (<2%),
fore a transient feature.
and this melt is sucked into weak grain boundaries that
open during shear. As the stress is increased, or as the melt
fraction increases, grain boundaries at a high angle to the SOLIDIFICATION
foliation begin to dilate (Schulmann et al. 2008). The pyrometamorphic examples described above are those
in which the temperature dropped so fast after the meta-
If the temperature remains high after deformation has
morphic peak that the liquid phase did not have time to
stopped, the melt fi lms on grain boundaries will begin
crystallize. This does not happen in most geological envi-
to break up. This is because replacement of a dry grain
ronments, where cooling rates are very much slower. In the
boundary by a stable liquid fi lm is only possible if the
majority of slowly cooled rocks, any melt present crystal-

E LEMENTS 249 A UGUS T 2011


Backscattered electron image of a mylonite composed Neosome in a migmatite from Australia, showing
FIGURE 4 predominantly of K-feldspar (pale grey), with grains of
FIGURE 6 crystal faces of microcline (Kfs) and (pseudomorphed)
quartz and albite(dark grey) on grain boundaries perpendicular to the cordierite (Crd). These facetted grains crystallized from a melt. The
lamination (shown by the double-ended arrow). FROM Z ÁVADA ET AL. base of the photo corresponds to 10 mm. R EPRODUCED FROM VERNON
(2007), REPRODUCED/MODIFIED BY PERMISSION OF AMERICAN GEOPHYSICAL UNION (2004, FIG. 4.80), WITH PERMISSION FROM C AMBRIDGE U NIVERSITY PRESS

lizes. The rate at which this occurs, the amount of H2O THE IMPORTANCE OF PORE SIZE
present and, critically, the size of the melt pockets control It has been known for a long time – particularly in the engi-
what the melt looks like after it has solidified. In some cases neering community where, for example, it is very impor-
the melt never actually gets to crystallize during cooling; tant to stop ice and halite crystallizing within cement
instead it is consumed by reactions with the surrounding structures – that the temperature at which crystallization
solid assemblage. These retrograde reactions are particu- occurs in confined spaces depends on the size of that space.
larly common in hydrous melts, and the growth of biotite This is because the thermodynamics of solidification is
at the expense of garnet and hydrous melt is a typical dependent on the energy of the interface between the
example (FIG. 5). Retrograde reactions modify and may growing crystal and its host liquid. This energy increases
erase microstructural and chemical information about the as the curvature of the interface becomes higher (i.e. as
melt-bearing stage and the peak metamorphic conditions. the crystal becomes smaller). This process is analogous
The simplest understanding of solidification can be gained to Ostwald ripening, whereby larger particles grow at the
from studying relatively large pockets of former melt, such expense of smaller ones. What it means in practice is that
as are now represented by layers and patches of leucosome the degree of supersaturation required for crystal growth
in migmatites. These crystallize a progressive sequence of into a small pore is greater than that required for growth
minerals that can be inferred from the relevant phase into a larger pore. This is easily seen in melt inclusions in
diagram. The early-formed minerals tend to crystallize as phenocrysts from extrusive igneous rocks. While the larger
euhedral grains bounded by growth faces. Normal or oscil- inclusions are crystalline, the smaller ones may be glassy
latory zoning may be present in plagioclase (Vernon 2011). despite having cooled at exactly the same rate. Crystals
The later, lower-temperature minerals fi ll in the gaps, like could not nucleate and grow in the smallest inclusions due
cement in a sedimentary rock (FIG. 6). This picture is very to the inhibiting effect of the small pore size. This effect
much what we would expect from igneous rocks, which in may be important even on much longer timescales. Cesare
essence is what these layers and patches of leucosome in et al. (2009b) recently discovered tiny droplets of glass of
migmatites are. However, solidification in the smaller pores broadly granitic composition preserved within refractory
may look very different indeed, and the reason for this is minerals in a granulite from the Kerala Khondalite Belt in
the effect surface curvature has in determining the degree India. These rocks have undergone a metamorphic cycle
of supersaturation required for crystal growth. lasting about 107 years. Inclusions larger than about 15 μm
have crystallized to “nanogranite”, a fine-grained aggregate
of quartz, feldspar and biotite (FIG. 7) but, astonishingly,
the smallest inclusions remain glassy. Further discoveries
of glassy inclusions in migmatites from other terranes
(e.g. Ronda, Spain) suggest they may be quite common
and could provide an exciting opportunity to determine
the original composition of the anatectic melt.
A major microstructural consequence of the effect of pore
size on inhibiting solidification is a change in crystallization
order. This is most easily seen in relatively simple systems,
such as quartzofeldspathic migmatites, in which the liquid
is saturated in two or three phases (quartz and one or two
feldspars). In a narrow pore bounded by quartz grains, the
quartz component of the liquid can crystallize by over-
growth on the walls; no new quartz grains need to be nucle-
ated, so there is no kinetic barrier. The remaining liquid
becomes increasingly saturated in the feldspar component
until feldspar begins to nucleate and grow. Instead of the
Evidence of retrograde reaction in a granulite from simultaneous crystallization of minerals along a cotectic
FIGURE 5 Antarctica. The central garnet grain that grew during predicted by equilibrium thermodynamics, the result is
the prograde melting reaction (Grt) was partially replaced by a sequential crystallization, producing a microstructure
fine-grained intergrowth of biotite (Bt) and feldspar (Fsp) during in which the fi nal melt pockets are pseudomorphed by
solidification.

E LEMENTS 250 A UGUS T 2011


primarily concentrated in thick fi lms on grain boundaries,
a “string of beads” texture may result (FIG. 10). At fi rst sight
this might be confused with the textures formed in the
early stages of static recrystallization of highly deformed
rocks, but the presence of other indicators of melting (such
as larger-scale segregations or peritectic minerals) should
be diagnostic.

THE PATH TO THE SURFACE


While pyrometamorphic aureoles give clues about the
onset of melting and enclaves provide a window into the
metamorphic peak, we also need to know what occurs after
the rock has solidified and is cooling down; we need to
know how the melting story ends. This is because most of
the anatectic rocks we collect have been exhumed slowly,
in some cases after further heating and/or deformation
events, and this later history will also have left its imprint
“Nanogranite” inclusion inside a garnet in granulite on the microstructures.
FIGURE 7 from the Kerala Khondalite Belt (India). Kfs = potassic
feldspar, Pl = plagioclase, Bt = biotite, Qtz = quartz. R EPRODUCED FROM Understanding these microstructures is only possible once
CESARE ET AL. (2009B), WITH PERMISSION FROM THE G EOLOGICAL SOCIETY OF we know how to recognise the remains of melt left behind
A MERICA after the bulk of it was extracted to form intrusions higher
in the crust. What we see seems to depend again on the
amount of H 2O present because of its role in enhancing
the mineral with the greatest difficulty nucleating. This
diffusion. If there is sufficient H 2O and the rocks stay
process shows up very clearly in migmatites in which the
hot enough for extensive diffusion, the microstructures
pore structure is fi lled by single plagioclase grains (FIG. 8),
change, driven by the reduction in interfacial energy. The
commonly with a shape highly reminiscent of melt pockets
rock attempts to minimize the energy associated with grain
in experimental charges (Sawyer 1999). The dihedral angles
boundaries by decreasing surface area – this leads to a
at the corners of these cuspate grains are generally much
general increase in grain size and a straightening of highly
lower than those expected for solid grains, pointing to
curved and irregular grain boundaries. Highly cuspate
inheritance of the shape of the melt-fi lled pore by the pseu-
pseudomorphs of melt at three-grain junctions (e.g. FIG. 8)
domorphing mineral.
become more rounded as the inherited dihedral angle
increases towards the angle for solid-state equilibrium,
THE IMPORTANCE OF DIFFUSION which is generally in the range 110–140˚ (FIG. 11). Because
Microstructures resulting from reactions that yield several textural equilibration and recrystallization can wipe out
product minerals are highly dependent on diffusion rates. any microstructural record of a previous melting event,
If H 2O is available, reactions tend to go to completion. one would anticipate that the microstructures resulting
In contrast, reactions in drier rocks are generally incom- from solidification are more likely to be retained in dry
plete, and such rocks contain microstructures that result rocks, and in particular dry rocks which have not been
from diffusion-limited control; coronae, intergrowths and significantly deformed on the retrograde path. If the micro-
symplectites are a common consequence of this (FIG. 5). structures have been obliterated, recognising the former
The most common type of intergrowth in solidifying presence of melt becomes extremely difficult, and melt
crustal rocks is that between quartz and alkali feldspar, inclusions in peritectic relicts may be the only evidence
known as granophyre (FIG. 9), which is common in the preserved.
relatively rapidly cooled shallow crust.
If the cooling rate is sufficiently slow, the balance between
the imperative to crystallize and the constraints imposed
by diffusion means that it becomes possible to nucleate
and grow individual grains from the melt. If the melt was

A partially melted quartz–plagioclase rock (also Partially melted quartzofeldspathic gneiss from the
FIGURE 8 containing some opaque Fe oxides) from the aureole
FIGURE 9 aureole of the Rum Igneous Complex, western
of the Duluth Igneous Complex, Minnesota, USA. The minor plagio- Scotland. The early-formed feldspar has grown as large euhedral
clase (now all brown and turbid) pseudomorphs an original textur- grains, while the later feldspar forms a complex intergrowth with
ally equilibrated melt-filled porosity. Pl = plagioclase, Qtz = quartz quartz (granophyre).

E LEMENTS 251 A UGUS T 2011


“String of beads” texture formed by crystallization at A highly cuspate plagioclase grain (centre) in a granu-
FIGURE 10 sufficiently slow rates to permit nucleation of indi- FIGURE 11 lite from the Nemiscau subprovince, Canada. This
vidual grains from the original melt film on boundaries between grain is inferred to have originally had low dihedral angles against
quartz (Qtz) and feldspar (Fsp) grains. The continuous, solidified the surrounding orthopyroxene (Opx), most likely inherited from an
former melt films are shown by the arrows. original melt-filled pore. However, the plagioclase developed
bulbous ends as the dihedral angle increased during approach to
solid-state textural equilibrium.

ACKNOWLEDGMENTS
and processes. We acknowledge helpful and constructive
We are grateful to those who have collaborated with us on reviews by Michael Brown, Hap McSween, Ondrej Lexa
studies of partially melted rocks; they have contributed and Ron Vernon.
greatly to our understanding of the underlying controls

REFERENCES Guernina S, Sawyer EW (2003) Large- Schulmann K, Martelat J-E, Ulrich S,


scale melt-depletion in granulite Lexa O, Štípská P, Becker JK (2008)
Acosta-Vigil A, London D, Morgan GB terranes: an example from the Archean Evolution of microstructure and melt
(2006) Experiments on the kinetics Ashuanipi Subprovince of Quebec. topology in partially molten granitic
of partial melting of a leucogranite at Journal of Metamorphic Geology 21: mylonite: Implications for rheology
200 MPa H 2O and 690–800°C: compo- 181-201 of felsic middle crust. Journal of
sitional variability of melts during the Geophysical Research 113: B10406,
onset of H 2O-saturated crustal anatexis. Holness MB (2006) Melt-solid dihedral doi:10.029/2007JB005508
Contributions to Mineralogy and angles of common minerals in natural
Petrology 151: 539-557 rocks. Journal of Petrology 47: 791-800 Spry A (1969) Metamorphic Textures.
Pergamon Press, Oxford, 350 pp
Acosta-Vigil A, Buick I, Hermann J, Cesare Holness MB, Watt GR (2002) The aureole
B, Rubatto D, London D, Morgan GB VI of the Traigh Bhàn na Sgùrra sill, Isle of Terasaki H, Frost DJ, Rubie DC,
(2010) Mechanisms of crustal anatexis: Mull: Reaction-driven micro-cracking Langenhorst F (2005) The effect of
a geochemical study of partially melted during pyrometamorphism. Journal of oxygen and sulphur on the dihedral
metapelitic enclaves and host dacite, SE Petrology 43: 511-534 angle between Fe–O–S melt and silicate
Spain. Journal of Petrology 51: 785-821 minerals at high pressure: Implications
Holness MB, Dane K, Sides R, Richardson for Martian core formation. Earth and
Braun I, Kriegsman LM (2001) Partial C, Caddick M (2005) Melting and Planetary Science Letters 232: 379-392
melting in crustal xenoliths and melt segregation in the aureole of the
anatectic migmatites: a comparison. Glenmore plug, Ardnamurchan. Journal Vernon RH (2004) A Practical Guide
Physics and Chemistry of the Earth A of Metamorphic Geology 23: 29-43 to Rock Microstructure. Cambridge
26: 261-266 University Press, Cambridge, 594 pp
Laporte D, Watson EB (1995)
Cesare B, Salvioli-Mariani E, Venturelli Experimental and theoretical Vernon RH (2011) Microstructures of
G (1997) Crustal anatexis and melt constraints on melt distribution in melt-bearing metamorphic rocks. In:
extraction during deformation in the crustal sources: the effect of crystalline van Reenen DD, Kramers JD, McCourt
restitic xenoliths at El Joyazo (SE Spain). anisotropy on melt interconnectivity. S, Perchuk LL (eds) Origin and
Mineralogical Magazine 61: 15-27 Chemical Geology 124: 161-184 Evolution of Precambrian High-Grade
Gneiss Terrains, with Special Emphasis
Cesare B, Rubatto D, Gómez-Pugnaire Marchildon N, Brown M (2002) Grain- on the Limpopo Complex of Southern
MT (2009a) Do extrusion ages reflect scale melt distribution in two contact Africa. Geological Society of America
magma generation processes at depth? aureole rocks: implications for controls Memoir 207, pp 1-12
An example from the Neogene Volcanic on melt localization and deformation.
Province of SE Spain. Contributions to Journal of Metamorphic Geology 20: Vernon RH, Collins WJ (1988) Igneous
Mineralogy and Petrology 157: 267-279 381-396 microstructures in migmatites. Geology
16: 1126-1129
Cesare B, Ferrero S, Salvioli-Mariani Rosenberg CL, Riller U (2000) Partial-
E, Pedron D, Cavallo A (2009b) melt topology in statically and dynami- Závada P, Schulmann K, Konopásek
“Nanogranite” and glassy inclusions: cally recrystallized granite. Geology J, Ulrich S, Lexa O (2007) Extreme
The anatectic melt in migmatites and 28: 7-10 ductility of feldspar aggregates—melt-
granulites. Geology 37: 627-630 enhanced grain boundary sliding and
Sawyer EW (1999) Criteria for the recog- creep failure: Rheological implica-
Grapes R, Li X-P (2010) Disequilibrium nition of partial melting. Physics and tions for felsic lower crust. Journal of
thermal breakdown of staurolite: Chemistry of the Earth A 24: 269-279 Geophysical Research 112: B10210,
A natural example. European Journal doi:10.1029/2006JB004820
of Mineralogy 22: 147-157 Sawyer EW (2001) Melt segregation in
the continental crust: distribution and
movement of melt in anatectic rocks.
Journal of Metamorphic Geology 19:
291-309

E LEMENTS 252 A UGUS T 2011

You might also like