You are on page 1of 50

Journal of Geodynamics 32 (2001) 115–164

www.elsevier.com/locate/jgeodyn

From microscope to mountain belt: 150 years of petrology


and its contribution to understanding geodynamics,
particularly the tectonics of orogens
M. Brown*
Laboratory for Crustal Petrology, Department of Geology, University of Maryland,
College Park, MD 20742, USA

Received 21 May 2000; received in revised form 20 March 2001; accepted 26 April 2001

Abstract
Thirty-five years ago the introduction of the plate tectonics paradigm led to a new understanding of oro-
geny. Subsequently, the development of advanced instruments for remote collection of information and for
analysis of elemental and isotopic composition of materials, and the increases in computing power have
enabled an unprecedented number of high-precision data about the Earth to be collected, analyzed, modelled
and displayed. Within this revolution in global tectonics, the metamorphic petrologist has developed methods
to unravel the depth, thermal, temporal and deformational history of orogens using detailed observations at
map, hand sample and thin-section scales in combination with elemental and isotope data, and using inverse
and forward modelling. Two exciting new directions in metamorphic petrology in relation to geodynamics
concern the kinship between earthquakes and metamorphic reactions in subduction zones, and the petrology
of the Earth’s mantle. Evidence of the changes in pressure (P) and temperature (T) in the Earth’s crust and
upper mantle during the break up, movement, and collision of pieces of the continental lithosphere is
sporadically recorded by the mineralogy and microstructures preserved in rocks exhumed to the surface.
Better calibration of phase equilibria, the use of internally-consistent thermodynamic data sets and the
development of techniques to retrieve close-to-peak P–T conditions from metamorphic rocks have yielded
more precise P–T data that enhance our ability to characterize the path followed by individual rocks in P–T
space. An improved ability to date segments of the P–T path, and to separate the length of time associated with
the prograde (increasing T) evolution from the age of close-to-peak P–T conditions has enabled better under-
standing of the rates and processes involved in lithosphere thickening. At the same time, better constraints on
the retrograde thermal history have contributed to our knowledge of the several tectonic processes that may
operate during exhumation, although these are less well understood. The expanding database of key
information, combined with predictions from modelling, has allowed the identification of characteristic P–

* Tel.: +1-301-405-4080; fax: +1-301-314-9661.


E-mail address: mbrown@geol.umd.edu

0264-3707/01/$ - see front matter # 2001 Elsevier Science Ltd. All rights reserved.
PII: S0264-3707(01)00018-7
116 M. Brown / Journal of Geodynamics 32 (2001) 115–164

T–t evolutions expected for rocks that have undergone distinct tectono-metamorphic histories. However,
relating structural events recorded by rocks to specific points along the P–T evolution remains problematic,
particularly regarding complex overprinting patterns of inclusion trails in porphyroblasts. These advances
have improved our understanding of the tectonic evolution of orogens. At the extreme of conditions for crustal
metamorphism are the recently discovered ultra-high pressure (UHP) and ultra-high temperature (UHT) facies
of metamorphism. Both are problematic given our limited knowledge of processes at these conditions, parti-
cularly the return of UHP rocks from peak-P conditions and the mechanism for extreme heat in the crust in
UHT metamorphism. The extreme depth inferred for metamorphism in some UHP terranes raises the issue
of whether theoretically plausible tectonic overpressures can be dynamically maintained to affect meta-
morphic reactions. If the pressure gradient recorded by UHP rocks is greater than lithostatic, the UHP
metamorphism may have occurred at depths shallower than currently believed. These studies have pro-
vided a reliable first-order framework for the comparison of rocks of ancient suture zones where the plate
tectonics situation is less certain. However, orogens are spatially and temporally extended nonlinear sys-
tems with feedback relations. Such complex systems generate apparently simple behavior by self-organi-
zation, and the influence of unique histories must be respected. # 2001 Elsevier Science Ltd. All rights
reserved.

1. Introduction

The scientist does not study nature because it is useful; he studies it because he delights in it, and
he delights in it because it is beautiful. If nature were not beautiful, it would not be worth
knowing, and if nature were not worth knowing, life would not be worth living. Of course I do
not here speak of that beauty that strikes the senses, the beauty of qualities and appearances;
not that I undervalue such beauty, far from it, but it has nothing to do with science; I mean that
profounder beauty which comes from the harmonious order of the parts, and which a pure
intelligence can grasp. Jules Henri Poincaré, translated from a passage in Science et méthode.

There is nothing more beautiful than a metamorphic rock seen in thin section under the pet-
rographic microscope, except perhaps the view of a mountain belt at dawn. These two inspiring
elements of geology are related by orogeny. Orogeny is the response of a geographically delimited
segment of Earth’s crust to Earth’s internal processes. It leads to alteration of the rock fabric
during deformation, metamorphism and magmatism, and results in the building of mountains.
The geographically delimited segment of the crust is referred to as an orogen. Much of the
Earth’s continental crust comprises deformed metamorphic rocks that formed during orogenesis.
On modern-day Earth, orogeny is a consequence of mantle convection and plate tectonics (e.g.
Davies, 1999), which reflects the self-organized behavior of a non-equilibrium system maintained
by exchange of energy with the outside world (Hodges, 1998). It is this complex dynamical system
that represents ‘‘ . . . the harmonious order of the parts . . .’’ to which Poincaré refers. In this paper
I illustrate how metamorphic petrology, one of the parts, has contributed to our understanding of
the whole process of orogeny.
Orogens are complex self-organized systems that enable efficient dissipation of energy, mani-
fested in feedback relations between deformation, metamorphism and anatexis (Brown and Solar,
1998a). How this feedback is recorded in orogenic crust is related to the properties of crustal rocks,
M. Brown / Journal of Geodynamics 32 (2001) 115–164 117

especially rheology, as well as plate interactions (e.g. Axen et al., 1998; Flesch et al., 2000, in press;
Vanderhaeghe and Teyssier, 2001). Rheology of orogenic crust varies with tectonic setting and with
evolution in time. Based on vertically averaged viscosities, orogens may be weak or strong, or they
may evolve from one to the other, particularly from weak to strong due to strain hardening.
However, the rheology will evolve through the prograde (increasing T) to retrograde (decreasing T)
metamorphic cycle (Brown and Solar, 2000). The rheology is weak if the crust melts and granite
magma is transferred to the upper crust (transiently weaker rheology locally in the upper crust,
evolving with cooling to stronger rheology for the whole orogenic crust), and with strain locali-
zation (Ellis et al., 2000). Weak orogens will absorb stresses and the effects will be concentrated
mainly in the orogen itself, whereas a strong orogen is associated with high stress levels in the
foreland, which causes deformation to affect a large part of the foreland, but not necessarily with
associated metamorphism or magmatism. Thus, coupling plays a critical role in controlling
deformation, subsidence, morphology and erosion in the orogen, the fore deep and the foreland.
Interactions among lithosphere plates drive orogenesis (e.g. Dewey and Bird, 1970; Dewey,
1975), within the context of which many interconnected processes may occur (e.g. Dewey, 1980),
including: variation in angle and rate of subduction; collisions; slab breakoff; lithosphere dela-
mination; partitioned flow; uplift; and exhumation. Collisions that occur during the movement of
the lithosphere over the Earth’s surface may involve plates of the same type, i.e. oceanic-oceanic
or continental-continental, or of different type, i.e. oceanic–continental. Thus, the large-scale
effect of continent break-up, drift, ocean closure and continent collision will be to generate a
series of interactions such that rocks of diverse origin, with different evolutions, may end up
juxtaposed, or interleaved in an orogen, or remnants of these may be preserved in an ancient
suture zone. The subduction of India beneath the Eurasian plate to form the Himalaya and the Tibe-
tan Plateau (Hodges, 2000) is an example of continental collision occurring today, which process is
widely inferred for major Phanerozoic orogens such as the Caledonian/Appalachian and Variscan
Orogens. Further back in the Earth’s history than the Phanerozoic, the tectonic regime cannot be not
well constrained due to the fragmentary nature of the evidence, although terranes of stable crust
were involved (e.g. Nutman et al., 1996; Komiya et al., 1999). Plate tectonics activity is inferred as
early as the late Archaean, based on evidence recorded in rocks as old as c. 2.7 Ga (Snoeyenbos et al.,
1995; Friend et al., 1996; Bleeker et al., 1999a,b), but the tectonic regime in the Archaean remains
obscure (e.g. Hamilton, 1998).
Subduction precedes collision, so that the degree of preheating during ocean closure is an impor-
tant variable in the thermal evolution of collision orogens (Jamieson et al., 1998). However, although
there is a clear involvement of mantle-derived magma in continental arc evolution, the role of litho-
sphere delamination, slab detachment and mantle magmatism in the evolution of collision zones
requires further study. For example, a recently proposed evolutionary model of the Himalayan–
Tibet system (Chemenda et al., 2000), based on modelling, geological and geophysical data, sup-
ports the deep subduction of continental crust, and shows that slab break-off, delamination and
tectonic underplating are fundamental events with drastic consequences on the evolution of the
convergent system. Additionally, as Vauchez et al. (1998) have shown, tectonic inheritance plays
a role, which causes orogenesis to be a trade-off between plate-boundary stresses, plate geometry
and the intrinsic properties of the plates (e.g. Flesch et al., 2000, in press).
It is necessary to know the thermal and mechanical consequences of particular types of inter-
action followed by individual samples of rock, especially the path in pressure (P)–temperature
118 M. Brown / Journal of Geodynamics 32 (2001) 115–164

(T)–time (t)–deformation (d) space, as well as the chances of preservation of evidence in rocks sub-
jected to the process, to reconstruct the sequence of events resulting in the exposed orogen. Field
relations, rock fabrics and microstructures, mineral assemblages and reaction relations, mineral and
fluid inclusion elemental composition and isotope ratio data provide an important, albeit often
incomplete, record of the rock history. Improved calibration of mineral equilibria and the develop-
ment of new techniques to retrieve close-to-peak P–T conditions from metamorphic rocks have
allowed refinement of both the evolution of metamorphic belts and the particular path followed by
individual rocks and units in P–T space (e.g. review by Spear, 1993). This refinement has been sig-
nificant in confirming the extreme P–T conditions associated with metamorphism during collision
orogeny, particularly UHP metamorphism in the Phanerozoic and UHT metamorphism in the
Precambrian. Our improved ability to date segments of the P–T evolution, and to separate the
length of time associated with the prograde evolution from the age of the close-to-peak P–T
conditions, has enabled better understanding of the rates and processes involved in thickening the
crust. Similarly, better constraints on the thermal history during exhumation have contributed to
our ability to separate several tectonic processes that operate post-peak metamorphism.
Fluid inclusions in minerals preserve direct evidence for the presence and composition of
ancient fluids that have long since left the sample. Fluid inclusions are tiny quantities of liquid,
vapor, or mixtures of these phases, from submicroscopic up to several hundred micrometers in
diameter, trapped in isolated voids, that may be observed under a microscope in transmitted light
once the rocks are cut into thin slices and polished. Observations reveal that samples of the
lithosphere of 0.001 m3 volume commonly contain billions of fluid inclusions. It is of historical
interest to note that the basic principles of the study of fluid inclusions are as old as modern
petrography, having been introduced by Sorby (1858).
Detailed petrography permits discrimination between inclusions that formed during initial mineral
growth (primary inclusions) from those formed some time after mineral growth (secondary inclu-
sions). Although most inclusions contain only one phase at the time of entrapment (silicate liquid or
volatile phase), during cooling, that phase may unmix to form a vapor bubble and daughter crystals
(Sorby, 1858; Roedder, 1984). Various sources of evidence suggest that many fluid inclusions pre-
serve the chemical and physical properties of the original parent fluids from which they formed.
Fluid inclusions are therefore considered samples of the volatile phase associated with meta-
morphism, deformation and exhumation, and chemical analysis of inclusions provides information
on the composition and density of this phase. These data enable us to constrain further both the
evolutionary history of specific rock samples and the fluid regime at points along that evolution.
Although the fluid regime attending metamorphism is an important element in understanding
metamorphic processes, a more detailed review is beyond the scope of this paper (but see Touret,
2001).
Numerical experiments of the thermo-mechanical evolution of the crust during orogeny have
also advanced as computer power has increased (e.g. Thompson, 1981; Huerta et al., 1998, 1999;
Pfiffner et al., 2000). In geodynamical models to investigate the lithosphere-scale evolution of
orogens, Beaumont and colleagues (e.g. Ellis and Beaumont, 1999) distinguish between cold col-
lision orogens (e.g. European Alps, New Zealand Alps; e.g. Beaumont et al., 1996; Ellis et al.,
1999, 2000) and collision orogens with inherited weakness (Appalachians, Variscides; e.g. Ellis et
al., 1998; Henk et al., 2000). The second group may evolve to orogens with weak thickened crust
and plateau development (Variscides, Himalayas/Tibet Plateau; e.g. England and Houseman,
M. Brown / Journal of Geodynamics 32 (2001) 115–164 119

1989; Henk, 1995, 1997, 1999, 2000; Royden et al., 1997; Clark and Royden, 2000). Further, a
change in mantle dynamics may lead to decoupling and delamination, slab breakoff, or subduction
zone roll back (e.g. Houseman et al., 1981; Royden, 1993b,c; Vissers, et al., 1995; Brouwer, 2000).
These processes have also been investigated using analogue models (Pinet and Cobbold, 1992;
Pubellier and Cobbold, 1996; Chemenda et al., 1995, 1996, 2000). An important question, therefore,
is do any of these tectonic settings produce unique metamorphic styles?
In addressing the issue of metamorphic evolution and orogeny, we are concerned with two
scales of interactions. At the global scale, we are concerned with similarities in styles of meta-
morphism and relationship to plate tectonic setting, which relationship reflects global processes.
At the regional scale, we are concerned with differences from terrane to terrane that enable us to
distinguish the particular combination of tectonic events that caused the metamorphism of each
terrane, and which inform us about plate tectonic interactions at the regional scale. The evolu-
tionary history at each of these two scales of interactions can be informed by detailed petrologic
studies at the outcrop, hand sample and thin-section scales; thus, the microscopic studies of the
petrologist contribute to our understanding of mountain belts.

2. Historical perspective

During the last few decades of the 19th and the first few decades of the 20th centuries, the
subject of metamorphism was addressed from two points of view. For convenience, these two
approaches may be distinguished as descriptive and genetic. The standard descriptive work,
published just after the turn of the century, is that of Grubenmann (1906). The descriptive
method is rooted in the conception of metamorphic rocks as a distinct group defined by certain
characters. In contrast, the genetic method starts from the idea of metamorphism as a certain
class of changes, which affect rocks of any kind and alter their characters. In its origin, it follows
the Huttonian doctrine as developed by Lyell (1860, pp. 591–593), to whom the word ‘meta-
morphism’ is due. In spite of this logical starting-point, no serious advance was made in following
the genetic method until the development of physical chemistry and the successful application of
the experimental method to problems of petrology.
The first serious attempt to discuss the process of metamorphism in relation to first principles
was made by Van Hise (1904), although the scope of his monograph is far wider than our modern
understanding of metamorphism, including subjects such as weathering, sedimentation and ore
deposits. In his anniversary address to the Geological Society of London, Harker (1919) sug-
gested that physical chemistry and experimental petrology would provide the foundation upon
which a complete theory of metamorphism would be built in the future, and the textbooks by
Ramberg (1952) and Turner (1968) represent attempts to achieve such a complete theory.
We have been aware of the intimate relationship between metamorphism and deformation since
the middle 19th century, when early studies of strained objects in slates enabled better under-
standing of the origin of slatey cleavage (e.g. Sharpe, 1847, 1849; Sorby, 1853, 1856; Haughton,
1856; Phillips, 1857). Ironically, it was not until a century later, with the seminal papers of Zwart
(1960, 1962) that recording the chronological succession of episodes of deformation and mineral
growth became a routine part of petrology. Early work stimulated by Zwart’s papers was sum-
marized by Spry (1969). More recently, Passchier and Trouw (1996) have synthesized our
120 M. Brown / Journal of Geodynamics 32 (2001) 115–164

understanding of micro-tectonics at the end of the last century, reflecting twenty-five years of
explosive growth in data and understanding. Indeed, in the metamorphic realm, the intimate rela-
tionship between deformation and metamorphism has made the specialties of petrology and struc-
tural geology essentially seamless (e.g. Vernon, 1976; Williams, 1994; Johnson and Vernon, 1995;
Hodges, 1998; Solar and Brown, 1999).
In 1907, Leith introduced the concept of a ‘metamorphic cycle’. Contemporaneous classic studies
in the late 19th and early 20th centuries (Barrow, 1893, 1912; Goldschmidt, 1911; Eskola, 1915, 1920;
Tilley, 1924a, 1924b, 1925) led to the idea that index minerals or mineral assemblages in metamorphic
rocks of a given bulk composition are a direct consequence of the intensive variables during meta-
morphism. A scheme was developed whereby a set of mineral assemblages, from rocks that cover the
full range of possible bulk compositions and that formed under the same limited range of intensive
variables, defines a metamorphic facies (e.g. Turner, 1968). The important reaction boundaries
that limit the different metamorphic facies have become increasingly better defined as data from
experimentally determined phase equilibria have accumulated and consequently, the meta-
morphic facies scheme allows a reliable first-order estimate of intensive variables based on field
studies or petrography alone (Fig. 1).
The observation that the variation in temperature from low to high grade in metamorphic belts
could be expressed by a series of metamorphic facies or by a metamorphic facies series appeared
in a landmark paper by Miyashiro (1961). This work subdivided metamorphic belts into one of five
facies series types, which, in order of increasing pressure are: (1) andalusite-sillimanite type (Fig. 1,
low P/T); (2) low-pressure intermediate group; (3) kyanite–sillimanite type (Fig. 1, int. P/T); (4)
high-pressure intermediate group; and, (5) jadeite–glaucophane type (Fig. 1, high P/T). Each of
these metamorphic facies series types was subsequently (Miyashiro, 1973) attributed to a specific
average geothermal gradient.
In addition to this fundamental observation of the relationship between geothermal gradient
and rock type, Miyashiro (1961) also noted that, in Japan and other parts of the Circum-Pacific
region, metamorphic belts of low-pressure type (1 or 2) and high-pressure type (4 or 5) occurred
side-by-side as pairs. The low-P/T type ‘inner’ belt on the continentward side, accompanied by the
intrusion of a large amount of granitic rocks, shows the facies sequence greenschist > amphibolite
>granulite whereas the ‘outer’ belt on the oceanward side, always of high-P/T type, exhibits either
the facies sequence blueschist>eclogite or zeolite>prehnite-pumpellyite>pumpellyite-actinolite
>greenschist>epidote-amphibolite.
The advent of the plate tectonic theory led to ideas of the temporal change of P–T conditions
during the evolution of metamorphic belts and to the development of tectonic models for the
processes of regional metamorphism (e.g. Oxburgh and Turcotte, 1974; England and Richardson,
1977). It became clear that geothermal gradients are transient in a dynamically evolving setting,
due to processes such as subduction, crustal thickening due to collision and removal of over-
burden by erosion or tectonics, and advection of heat in rising magma. Thus, supracrustal rocks
taken to depth, metamorphosed, and brought up to become exposed at the surface of the earth by
erosion or extension, experience changes of pressure and temperature, i.e. a single metamorphic
belt evolves across a range of transient geotherms rather than following a single geotherm (e.g.
England and Richardson, 1977; Fig. 2).
For this reason, the line in P–T space defined by a facies series at Earth’s surface is not the
equivalent of a geotherm. This line is actually the locus of (usually) peak metamorphic conditions
M. Brown / Journal of Geodynamics 32 (2001) 115–164 121

Fig. 1. P–T field of crustal metamorphism to show the eight principal metamorphic facies (Zeo, zeolite facies; PP,
prehnite–pumpellyite facies; EA, epidote–amphibolite facies). Boundaries between adjacent facies are transition zones
of considerable P–T width. The shaded bands represent the three principal facies series. Also shown as thin lines are
the phase boundaries between kyanite (high P, low to high T), sillimanite (high T, low to high P) and andalusite (low
P, intermediate T).

recorded by individual rock units at different times on individual transient geotherms. There are
various terms used in the literature for this line, including the ‘metamorphic geotherm’ (Oxburgh
and Turcotte, 1974; England and Richardson, 1977), the ‘P–T array’ (Thompson and England,
1984) and the ‘metamorphic field gradient’ (Spear et al., 1984). In these days of interdisciplinary
science, the P–T array is understandable to all and is to be preferred. The array of P–T conditions
at which maximum T was reached on each of a set of P–T–t loops, with each loop corresponding
122 M. Brown / Journal of Geodynamics 32 (2001) 115–164

Fig. 2. Geothermal gradients (dashed lines) in continental crust at times of 20, 40, 60 and 80 Ma after tectonic thick-
ening modelled by instantaneous over-thrusting. Erosion of the rock column above ‘A’ lowers P whilst heating is in
progress (arrows) so that maximum T is reached at 33 Ma. The array of Tmax—P(Tmax) points achieved by rocks that
subsequently reach the surface is shown as the ‘metamorphic geotherm’ (solid line), which is labeled with the times at
which the rocks appear at the surface. Reproduced from England and Richardson (1977, Fig. 1) with permission of the
Geological Society, London.

to a different initial depth, is the ‘piezothermic array’ (Richardson and England, 1979; Thompson
and England, 1984). The P–T and piezothermic arrays may coincide, but they need not do so,
and care must be taken to ensure that there are no metamorphic discontinuities in the sequence
(Mortimer, 2000).
The trajectory in P–T space followed by a particular volume of the metamorphic belt during
some interval of time is the P–T–t path. Various types of petrographic information may be
recorded along segments of the P–T–t path (Spear et al., 1984; Fig. 3), and using preserved por-
tions of this record, the petrologist attempts to reconstruct the evolutionary history (e.g. Spear
and Selverstone, 1983). Metamorphism might involve overprinting of one facies series type on an
earlier and different type. This may occur either as part of continuous evolution or as two discrete
superimposed events [e.g. southern Brittany (Brown, 1983; Jones and Brown, 1990; Brown and
Dallmeyer, 1996), the Ivrea Zone (Barboza et al., 1999; Barboza and Bergantz, 2000) and the
Alboran Domain (Platt et al., 1998; Soto and Platt, 1999)].
The formation of paired metamorphic belts was thought to be a result of under-thrusting of an
oceanic plate beneath an island arc or continental margin (Ernst, 1971a; Oxburgh and Turcotte,
1971; Miyashiro, 1973; Uyeda and Miyashiro, 1974). The inner metamorphic belt was interpreted
M. Brown / Journal of Geodynamics 32 (2001) 115–164 123

Fig. 3. Generalized P–T diagram showing relationship between P–T paths followed by individual samples (1, 2 and 3)
and metamorphic field gradient. Also shown are types of information that can be used to construct different parts of a
P–T loop for an individual sample, as follows: realm of entrapment of mineral inclusions in poikiloblastic host grains
(A); peak P–T conditions from thermobarometry (B); retrograde path from reset mineral equilibria and reaction rela-
tions (C); and realm of entrapment of fluid inclusions (D). Reproduced from Spear et al. (1984, Fig. 1) with permission
of the Geological Society of America.

as being developed in the arc because of heat advected through the crust by sub-crustal magma-
tism, while the outer metamorphic belt was interpreted to be developed in the trench because of a
lower geothermal gradient caused by the subduction process.
The plate tectonic paradigm has the consequence that oceanic island arcs, ocean plateaus and
continents may collide by the process of subduction. In suture belts formed by such collisions, the
associated metamorphism is likely to be complex. An example of such a suture belt is the Cale-
donian/Appalachian Orogen. It includes the type example (the Grampian Highlands of Scotland)
of the ‘‘. . . standard type of facies series’’ (Miyashiro, 1961), characterized by the stability of
kyanite at lower grade and of sillimanite at higher grade. In contrast, the Variscides of Europe
represent a problem, having been regarded as characterized by widely developed metamorphism
of the andalusite-sillimanite type associated with voluminous granitic magmatism, but without a
contemporaneous parallel high-P/T type metamorphic belt (Miyashiro, 1961). It is now clear that
the Variscides have a long and complex history of subduction and collision, and preserves evi-
dence of early high-P metamorphism as well as late high-T metamorphism (Brown and O’Brien,
1997; Franke et al., 2000).
Tibet might be an appropriate modern analog for some ancient low-P/T type metamorphic
belts. A variety of geophysical data have been accumulated from Tibet, on the basis of which it
has been argued that there is a melt-bearing middle layer capped at 15 km depth by a horizon of
124 M. Brown / Journal of Geodynamics 32 (2001) 115–164

laccolithic granites (Nelson et al., 1996; Alsdorf et al., 1998; Alsdorf and Nelson, 1999). Although
the crust is thick under Tibet, there is little erosion at the surface. As a result, if the collision
between the Indian continent and Asia slows to a stop, a likely response may be extension and
preservation of the middle crust as a migmatite belt.
Within a decade of the introduction of the plate tectonics paradigm, new theories of continental
growth focused on the idea that the crust of the continents had grown through the accretion of dis-
crete allochthonous fragments of oceanic and continental material at active plate margins. The ter-
rane concept was developed by workers in the North American Cordillera, and the term ‘‘terrane’’
was introduced by Irwin (1972) to refer ‘‘. . .to an association of geologic features, such as strati-
graphic formations, intrusive rocks, mineral deposits, and tectonic history, some or all of which lend
a distinguishing character to a particular tract of rocks and differ from those of an adjacent ter-
rane.’’ Eventually, the definition has come to emphasize the fault-bounded nature of terranes, and
to allow for the idea of composite terranes formed by pre-accretionary amalgamation.
The fact that petro-tectonic assemblages, such as volcanic arcs, fore-arc basins and subduction
complexes, may be extensive along a continental margin does not demonstrate that these assem-
blages formed in their present position and caution must be exercised in applying plate tectonic
models that assume that tectonic domains presently juxtaposed are genetically related (Brown,
1998a, b). Another feature of terrane tectonics in general, and transpressive orogeny in particular,
is the speed of events, particularly the short time from peak metamorphism to exhumation as
recorded by mineral thermo-chronometry (e.g. Brown, 1998b; Solar et al., 1998; Oliver et al.,
2000).
Referring to the Variscides of Europe, the traditional view of the tectonics, based on limited
orogen parallel translation of tectono-metamorphic terranes, may have constrained our ability to
explain in detail the complexity and the contrasts observed along the length of the belt. This issue
has been raised recently by Shelley and Bossière (2000), who argue that the Variscan belt of Iberia
and France is a collage of disparate terranes developed during a massive dextral transpressive
deformation as Laurentia slid along the boundary of Gondwana, and as Laurentia, Baltica, and
East Avalonia rotated clockwise to form Pangea. In this model, the Ibero-Armorican arc was
formed as the mobile dextral shear belts were deformed about the rigid Iberian basement block.

3. Methods in petrology

Science is built up with facts, as a house is with stones. But a collection of facts is no more a
science than a heap of stones is a house. Jules Henri Poincaré, translated from a passage in La
Science et l’hypothèse.

We must extract various types of information from the metamorphic rocks exposed at Earth’s
surface if we are to understand the depth-temperature-time-deformation evolution of orogens.
Primary data include observations made in the field and at hand-sample scale, and from exam-
ination of thin sections in transmitted and incident light, and using electron microscopy. These
data provide information about changes that have occurred in rocks, and by inference about the
history of the crust or mantle. The microstructure (shape, size, orientation and distribution of
M. Brown / Journal of Geodynamics 32 (2001) 115–164 125

grains) of a rock reveals information about mineral nucleation and grown, and mineral transfor-
mations. Other information obtained from microscopy includes the peak mineral assemblage, the
previous mineral assemblage if relicts are preserved within porphyroblasts or in reaction relations,
evidence of early deformation deduced from inclusion fabrics in porphyroblasts, deformation in
relation to growth of the peak mineral assemblage from matrix fabric – porphyroblast relations,
and retrograde reaction history. These observations are supplemented by element and isotope
compositional information on whole-rock samples, minerals and fluid inclusions.
The advent of large internally consistent thermodynamic data sets (Helgeson, et al., 1978;
Holland and Powell, 1985, 1990, 1998; Berman, 1988) and the associated development of multi-
reaction techniques (e.g. Powell, 1985; Powell and Holland, 1988, 1994; Berman, 1991; Gordon,
1992) have revolutionized thermobarometry over the past two decades. Nonetheless, uncertain-
ties on calculated pressures and temperatures remain large, generally more than 1.5 kbar and
 50  C. However, the contribution of systematic uncertainties to thermobarometry can be
minimized if we are concerned with differences in P–T rather than absolute P–T (Spear, 1989;
Worley and Powell, 2000). Further, these large internally consistent thermodynamic data sets
allow the construction of realistic P–T grids and pseudo-sections for specific bulk compositions in
model systems (Powell et al., 1998; Spear, 1999), and permit the modes and mineral assemblages
to be tracked through P–T grids, providing quantitative models for comparison with field
observation.
A word of caution is appropriate at this point. Most petrological studies assume that the stress
state of the earth is near lithostatic, and that barometric results can be translated directly to a
depth of metamorphism. However, in studies of rocks from orogens, this assumption may not be
valid. In addition to the question of ‘tectonic overpressure’ (discussed below under UHP meta-
morphism), it has been argued that even small deviatoric stresses under some circumstances can
be significant in the interpretation of metamorphic parageneses (Stüwe and Sandiford, 1994;
Tenczer et al., 2001). Stüwe and Sandiford (1994) have shown that the crust can support devia-
toric stresses of up to 100 MPa, and that localized decompression can result in different
apparent P–T–t paths with different volumes of rock recording ‘clockwise’ or ‘anticlockwise’
evolution. More recently, Tenczer et al. (2001) have argued that the common occurrence in
pressure shadows around porphyroblasts of the products of metamorphic reactions that were
crossed during decompression might reflect local rather than lithostatic decompression. Thus,
care should be taken in the interpretation of barometric results derived from mineral assemblages
in pressure shadows associated with porphyroblasts. Also, Kriegsman and Hensen (1998) have
argued that barometry based on the assumption of quartz saturation may be inappropriate in
melt-bearing rocks. In such rocks, the melt may be silica-undersaturated, which affects the stabi-
lity of important granulite facies mineral assemblages, particularly those involving spinel
(Kriegsman and Hensen, 1998). As with all things, the devil in thermobarometry is in the details!
Complementary temperature-time information can be obtained using isotopic thermo-chron-
ometers that have closure temperatures ranging from granulite facies conditions (U–Pb zircon
and monazite) to 100  C (fission track annealing in apatite). To link the P–T–t information to
thin-section and outcrop observations of deformation we have to incorporate specific evidence
reflecting the deformation events (d) into the P–T–t path information. Although valuable infor-
mation on the relative timing and kinematics of natural deformations potentially can be retrieved
from relationships between porphyroblasts, fabrics in the surrounding matrix and any inclusion
126 M. Brown / Journal of Geodynamics 32 (2001) 115–164

fabrics, many issues remain about the interpretation of inclusion fabrics in porphyroblasts (e.g.
Bell et al., 1992; Passchier et al., 1992; Johnson, 1999).
Finally, the relationship between styles of metamorphism and tectonic processes is addressed using
both inverse and forward modelling of possible evolutionary histories (e.g. Thompson, 1981;
Thompson and England, 1984; Peacock et al., 1994; Jamieson et al., 1998; Huerta et al., 1999).

3.1. Comments on geochronology

The rates and duration of metamorphic heating and granite generation and emplacement in
orogens are not well constrained, and determining the age of peak P–T conditions of meta-
morphism remains elusive, although it can be achieved (e.g. Mezger et al., 1989). Establishing rates
and time scales of processes such as heating has only recently come within our grasp (e.g. Vance,
1995), and time scales of magmatic processes remain poorly constrained, although faster rather
than slower seems to be the emerging consensus (Cruden, 1998; Thompson, 1999; Harris et al.,
2000; Davies and Tommasini, 2000; Oliver et al., 2000; Brown, 2001). We have better information
on cooling histories because there are a number of mineral thermo-chronometers from which
information can be retrieved along the whole exhumation P–T path.

3.1.1. In situ dating


Age determination of geologic materials conventionally has been accomplished by isotope
dilution thermal ionization mass spectrometry to measure isotope ratios in minerals separated
from crushed rock or by cutting individual grains out of thin sections, particularly zircon, mon-
azite, titanite and rutile. An alternative that has become more widely available in the past decade is
in situ secondary ionization mass spectrometry at individual points within sections through well-
characterized individual mineral grains, particularly zircon and monazite (e.g. Williams et al.,
1996; Frazier et al., 2000; Hermann et al., 2001). The U–Th–Pb dating of monazite by secondary
ion mass spectrometry microprobe analysis is a powerful tool in determining the timing relations
in metamorphic rocks (Spear and Parrish, 1996; Zhu et al., 1997; Ireland and Gibson, 1998),
particularly when used in conjunction with isotope dilution thermal ionization mass spectrometry
(e.g. Crowley et al., 2000; Foster et al., 2000).
The development of in situ mapping of Th, U and Pb using the electron probe micro-analyzer
(Suzuki and Adachi, 1991; Montel et al., 1996; Cocherie et al., 1998) has enabled rapid determination
of monazite ages in metamorphic rocks (Fig. 4). A critical aspect in rocks where multiple generations
of deformation and recrystallization can result in complex P–T–t–d histories, is the fact that electron
probe microanalysis enables dating of individual monazite grains within its micro-structural context.
Thus, potentially meaningful ages that will constrain the relationships between deformation and
mineral growth may be determined by this technique. Monazite is a ubiquitous accessory phase in
many metamorphic rocks, making it potentially useful in polymetamorphic terrains (e.g. Williams
et al., 1999), although the Th–U–Pb system in monazite may be complex and detailed petrography
and geochemical characterization may be necessary to interpret ages properly (Zhu and O’Nions,
1999a,b). An important potential use is on monazite inclusions in porphyroblast phases, especially
those with multiple generations of inclusion fabrics, and where monazite can be tied by micro-
structure to a particular fabric. Chemical dating of monazite using the electron probe micro-analyzer
does not eliminate the need for high-resolution geochronological studies using other methods.
M. Brown / Journal of Geodynamics 32 (2001) 115–164 127

Fig. 4. X-ray maps and age determinations of monazite separated from a metamorphic rock from the Western Gneiss
region of Norway. Note that complex zoning in Th (A) and U (B) composition is not seen in the calculated age map
(C), which shows only two age domains, consistent with the histogram of age determinations shown in (D). Repro-
duced from Williams et al. (1999, Fig. 1) with permission of the Geological Society of America.

3.1.2. Protolith ages


In Phanerozoic orogens, the age of deposition of supracrustal sequences is commonly well con-
strained on the basis of extrapolation by mapping from lower-grade rocks into higher-grade
regions, using normal stratigraphic principles. Also there are many age data available to constrain
the age of deposition using, for example, the U–Pb method on zircon from contemporaneous
horizons of volcanic rock, and the age of crystallization of plutons using, for example, the U–Pb
method on magmatic zircon and monazite. However, even in Phanerozoic orogens, there are pro-
blems in identifying beyond a reasonable doubt the protolith age of basement rocks. This parti-
cular issue has bedeviled studies of the Variscides, although progress in recent years has identified
both Cadomian and Caledonian events in the basement rocks (e.g. see Franke et al., 2000).
128 M. Brown / Journal of Geodynamics 32 (2001) 115–164

3.1.3. Prograde P–T path and age of peak P–T metamorphic conditions
During the past decade, techniques have been developed, principally utilizing garnet, both to
date close-to-peak P–T conditions and to constrain the period of heating by knowing the time
scale for growth of prograde garnet based on separate core and rim samples (e.g. Mezger et al.,
1989, 1993; Vance and O’Nions, 1992; Vance et al., 1998; Vance and Harris, 1999). Advantages to
utilizing garnet as a chronometer include its common occurrence in metapelites and its ubiquity
in thermobarometry. In principle, information on prograde metamorphism also can be gained by
careful use of the U–Pb method on accessory phases. This is possible because of the high closure
temperature for diffusion and refractory nature of minerals such as zircon and monazite. The
utility of the U–Pb method, in combination with metamorphic petrology, has been demonstrated
by Hawkins and Bowring (1999), Simpson et al. (2000) and Essex and Gromet (2000).
The issue that bedevils all chronometry in metamorphic petrology is to be sure of exactly what
event has been dated. To do this, an understanding of the genesis of U–Pb chronometers is nee-
ded in order to read from these minerals the record of processes during orogeny. To understand
the crystallization and recrystallization of accessory minerals requires an approach that combines
multiple techniques, scanning electron microscopy and cathode luminescence imaging, trace ele-
ment analysis by electron probe micro-analyzer, and U–Th–Pb analysis by in situ secondary ion
mass spectrometry. If necessary, conventional analysis by isotope dilution thermal ionization
mass spectrometry can be done on single-grains of a fully characterized accessory phase to obtain
high-precision age data. Using mineral separates requires correct interpretation of whether ages
reflect metamorphic or magmatic events, and whether these ages are inherited or overprinted.
This is a particular problem in high-grade and/or polymetamorphic rocks where melting may
have occurred, and multiple events of growth and/or dissolution of zircon and monazite may be
represented. The problem is exacerbated by a lack of information on how recrystallization or new
growth of accessory phases relates to reactions based on major phases that are used in thermo-
barometry. Further, we are not well informed on the temperature at which monazite and zircon
begin growing during metamorphism, which stage of the thermal history of the rock the U–Pb
system of each mineral is likely to record, and which factors influence accessory phase stability in
a given metamorphic environment.
These issues have recently been addressed in two detailed studies of monazite and zircon in
prograde metamorphic sequences, where greenschist to upper amphibolite or granulite facies
metamorphism is recorded over apparently continuous sections (Rubatto et al., 2001; Williams,
2001). The amount of metamorphic growth in both accessory minerals increases with temperature
and metamorphic grade. However, new zircon growth is influenced by rock composition and
driven by partial melting, factors that appear to have little effect on the formation of meta-
morphic monazite. In the study of Rubatto et al. (2001), Pb inheritance is widely preserved in
detrital zircon cores, but only a trace of inheritance is found in monazite in granulite facies rocks.
Thus, the closure temperature of the U–Pb system in large monazite crystals may exceed 750–
800  C, even for very slow rates of cooling.
In principle, in situ analysis allows correlation of accessory mineral relationships with the
metamorphic history. An alternative to in situ analysis is to cut individual mineral grains to iso-
late core and rim components. The aim of such studies is to determine the timing of accessory
mineral growth with respect to specific metamorphic or deformation events recorded in the rock.
This is an important step in the direct calibration of time along P–T paths. Examples of such
M. Brown / Journal of Geodynamics 32 (2001) 115–164 129

studies include those of Hawkins and Bowring (1997, 1999), Fitzsimmons et al. (1997), Finger et al.
(1998), Zhu and O’Nions (1999b) and Bingen et al. (2001). Recently, Viskupic and Hodges (2001)
have used accessory phase thermometry based on element partitioning between coexisting mon-
azite and xenotime, in combination with U–Th–Pb chronology on well-characterized mineral
grains to determine accurate age and cooling rate information for anatectic rocks from the Everest
Region of the Nepalese Himalaya. Monazite-xenotime thermochronometry appears to be a
powerful tool for reconstructing the thermal histories of high-temperature metamorphic terrains.
Metamorphism in the Ivrea Zone, northern Italy, illustrates well the kind of problem that is
generated in attempting to quantify the age of peak P–T metamorphic conditions. At issue is
whether the age of granulite facies metamorphism is older than or identical to the age of the
associated Mafic Complex, which has been interpreted to be the heat source for the granulite
facies metamorphism (Schmid and Wood, 1976; Sills, 1984; Schnetger, 1994; Sinigoi et al., 1996),
although Barboza et al. (1999) have questioned this interpretation. Recent detailed work using
U–Pb geochronology on zircons by Vavra et al. (1999) confirms an age of 2995 Ma for the
granulite facies metamorphism and anatexis. However, the uncertainty on this age does not allow
resolution of the argument over whether the granulite facies metamorphism and anatexis is older
than or coeval with the Mafic Complex, which yields ages from different units in the range 299 5
to 285+7/5 Ma (Pin, 1986; Vavra et al., 1999). Subsequent alteration and Pb-loss have affected
zircon crystals that grew during granulite facies metamorphism; it is clear that the Ivrea Zone
rocks have suffered a long history during cooling (Vavra et al., 1999). Previous interpretations of
monazite data as cooling ages are debated (Henk et al., 1997; Vavra and Schaltegger, 1999).

3.1.4. Post-peak thermal histories


Post-peak thermal histories are better known than prograde thermal histories, and these data are
commonly used to constrain models for the thermo-mechanical evolution of orogens (e.g. Batt and
Braun, 1997). It has become clear during the past decade that the time-integrated rates of cooling of
orogens vary considerably. Thus, exhumation may vary from very slow, for example 1.5 C Ma1
for at least 150 million years following the last phase of high-grade metamorphism (Mezger et al.,
1991), to extremely rapid, with rates of up to 100  C Ma1 during 1–10 million years (e.g. Baldwin et
al., 1993; Brown and Dallmeyer, 1996). This variation suggests a range between limited vertical dis-
placement and approximate isostatic equilibrium and significant vertical displacement probably
reflecting tectonic exhumation through upper crustal extension or wedge extrusion.
Hermann et al. (2001) present an example of the effectiveness of a combined petrological and
geochronological approach in unraveling the post-peak thermal history of an orogen by relating
zircon geochronology to metamorphic P–T along the post-peak exhumation path. Using trans-
mitted light microscopy, cathode luminescence imaging, trace element analysis and mineral
inclusion data, four different domains are distinguished within individual zircon crystals and
related to specific metamorphic P–T conditions. The metamorphic conditions record exhumation
of deeply subducted continental crust from P>43 kbar and T 950  C to P5 kbar and T
600  C during the time interval 5275–5265 Ma. The mean age of zircons formed during
exhumation and analyzed in that study is 5283 Ma, which indicates that decompression took
place in less than 6 Ma at rates >18 mm Ma1. The P–T data and inferred path suggest two-
stages of exhumation, first melt-assisted, buoyancy-driven return to the depth of the lower crust,
and second during tectonic extension after a period of isobaric cooling. Rates of cooling may
130 M. Brown / Journal of Geodynamics 32 (2001) 115–164

change during the period of exhumation (e.g. Hawkins and Bowring, 1999), which likely reflect
variations in the rate (e.g. Dempster, 1984) and/or mechanism of uplift and exhumation (Her-
mann et al., 2001).
The inference of denudation rates from the temperature-time histories of rocks is not straight-
forward. This is because the geothermal gradient evolves with time, as denudation advects heat
towards the land surface, which invalidates a common assumption used in such studies, that
temperature increases linearly with depth and that this geothermal gradient is constant over time.
As Moore and England (2001) have shown, the curvature in the temperature–time path recorded
by an exhumed rock is strongly dependent upon denudation rate and the temperature of the rock
before denudation began. However, it appears only weakly dependent on the initial geothermal
gradient, the original depth of burial of the rock, or the duration of the denudation. This result
permits the robust estimation of denudation rates from temperature–time data, with minimal
prior knowledge of the geothermal gradient. Moore and England (2001) have developed a process
for inferring denudation rates from temperature-time data and, in applying the process to data
from the Nanga Parbat Massif in the Himalaya, they infer denudation rates lower than pre-
viously suggested. Nonetheless, it is likely that we still have much to understand in converting
temperature–time information to denudation rates!

3.2. New tools at the new millennium

Most new tools relate to microscopy or in situ analysis, although the increase in computer power
and its easy availability have allowed more complex thermo-mechanical models of orogens to be
developed (Koons, 1995; Thompson et al., 1997; Huerta et al., 1998; Waschbusch et al., 1998;
Beaumont et al., 1999; Ellis et al., 1999; Pfiffner et al., 2000). There remains the difficult issue of
relating experimental petrology to the petrology of natural materials (e.g. Brown and Rushmer,
1997), and relating experimental and geological rheology to the rheological assumptions that need
to be made in geodynamic modelling (Handy et al., 2001; Paterson, 2001). Advances in image
analysis have opened up new research directions (e.g. Heilbronner, 2000), and two-dimensional
numerical simulations of the evolution of rock microstructures during deformation and meta-
morphism now are possible (e.g. Jessell et al., 2001).
In garnet, there is a coupling between major and accessory phases during reaction progress and
because trace elements are sensitive to changes in accessory mineral assemblage and/or fluid
composition, commonly having slower diffusivities than divalent cations, except Ca (Pyle and
Spear, 1999), these features can be used to calibrate trace element thermobarometers (e.g. Pyle
and Spear, 2000). Features such as changes or discontinuities in composition of trace elements
identify changes in reacting assemblage (Fig. 5) and reveal information not recorded by divalent
cations (e.g. Chernoff and Carlson, 1999). Mapping major and trace element distributions (Tracy,
1982; Pyle and Spear, 1999) and high-resolution mapping of matrix phases in relation to matrix
microstructure (e.g. Williams et al., 2001) are now routine elements of all modern petrographic
studies. Mapping in situ distributions of elements has allowed advances in understanding pro-
cesses reflected by element distributions. For example, the geometry of growth of porphyroblast
phases can be examined by serial sectioning and 3-D visualization based on information from
optical, backscatter electron and quantitative X-ray composition images (Spear and Daniel, 1998,
2001; Daniel and Spear, 1999; Yang and Rivers, 2001).
M. Brown / Journal of Geodynamics 32 (2001) 115–164 131

Fig. 5. Element distribution maps for garnet: (a) Sc; (b) Cr; (c) Y; (d) P. The zone of high Sc/Cr/Y (white arrows) is
interpreted to have grown after moderate P dehydration melting of muscovite and during dehydration melting of bio-
tite (Pyle and Spear, 1999). Garnet contains xenotime in the core and in the low-Y rim (red arrow), but no xenotime is
observed in the high-Y zone produced during biotite dehydration melting. Note that the melting reaction is not
recorded by any detectable change in P. The image is approximately 8 mm wide. Reproduced from Pyle and Spear
(1999, Fig. 12) with permission of the Mineralogical Society of America.

Recent developments in scanning electron microscope based measurements of crystallographic


orientations are at the heart of many new microstructural studies. For example, electron back-
scatter diffraction (EBSD) in the scanning electron microscope enables crystallographic orienta-
tions of points as small as a micrometer to be measured easily, precisely and quickly. When coupled
to orientation contrast imaging, this technique provides qualitative maps of crystallographic
orientation, which allows objective definition of domain size and domain boundary irregularity
(Prior et al., 1999). Further, the crystallographic misorientation between pairs of orientation
measurements can be determined, and misorientation data sets can be collected easily (Leiss et al.,
2001). Thus, microstructural features in metamorphic tectonites can be quantified using mis-
orientation analysis, which allows hypotheses on the formation of such microstructural features
132 M. Brown / Journal of Geodynamics 32 (2001) 115–164

to be tested (Wheeler et al., 2001). The principal advantage of EBSD over more established
methods of textural and micro-structural analysis is that it can be applied to any crystalline
material, such that monomineralic quartz, calcite or olivine rocks will no longer dominate micro-
structural and textural studies.
High resolution X-ray computed tomography has revolutionized the manner in which we can
examine the 3-D spatial relationships among features in rocks without destroying them, as would
occur in serial sectioning (e.g. Carlson and Denison, 1992; Denison and Carlson, 1997; Denison
et al., 1997). In migmatites, HR X-ray CT illustrates well the 3-D distribution of the melt flow

Fig. 6. High-resolution X-ray computed tomographic image of leucosome structure in migmatite from the contact
aureole of the Onawa pluton, Maine, USA. Granite leucosome is highlighted in red and host material is rendered
transparent. Note crosscutting granite leucosome hosted in extensional shear zone that disrupts layer-parallel granite
leucosomes. Image provided by Nathalie Marchildon, University of Maryland, USA.; created using VoxBlast in the
HRx-ray CT facility at the University of Texas, Austin, USA.
M. Brown / Journal of Geodynamics 32 (2001) 115–164 133

network represented by leucosome (Fig. 6) due to the mass density contrast between leucosome,
mesosome and melanosome (Brown et al., 1999). In anatectic rocks, SEM-CL allows identifica-
tion of textural features related to melting, crystallization, and melt movement that are not
resolvable with the petrographic microscope (Watt et al., 2000). These two novel techniques offer
exciting potential to characterize the nature and distribution of petrologic features, in particular
as related to melt evolution in anatectic rocks (e.g. Marchildon and Brown, 2001). One limitation
of HR X-ray CT is that spatial resolution is not sufficient for many problems of interest, CT
values, which represent contrast in CT images, are not quantitative, and it takes a long time to
obtain 3-D images. The newly developed X-ray CT systems using synchrotron radiation will
allow qualitative evaluation of CT values based on linear absorption coefficients and 3-D images
with high spatial resolutions are easily obtained (Tsuchiyama et al., 2001).
Recently, new advances have been made in laser ablation-inductively coupled plasma mass
spectrometry (LA–ICP–MS), a technique that bridges the gap in sample size between isotope
dilution mass spectrometry and secondary ion mass spectrometry (SIMS). LA–ICP–MS is likely to
revolutionize terrestrial geochemistry, both in elemental and isotopic analyses of suitable materials
(Horn et al., 2000; Prince et al., 2000), but the significant capabilities of SIMS for thermo-chron-
ological investigations remain largely unexploited (e.g. Grove and Harrison, 1999—depth profiling
to recover continuous, high-T thermal histories). These techniques are complementary in their
analytical strengths, and undoubtedly some of the ambitious analytical goals in petrology will be
met by this combination of tools (Ireland, 1999).

4. Styles of orogenic metamorphism and relationship to tectonics

4.1. Orogenic P–T–t paths

In the early days of definition of P–T–t paths from rocks, the resulting paths were fairly simple.
Clockwise and counterclockwise (in P–T space) loops were recognized, isobaric cooling and iso-
thermal decompression segments were identified, advection of heat in magma to produce thermal
spikes was accepted, and continued refrigeration during subduction exhumation to preserve high-
P/T type metamorphic rocks was understood (e.g. Brown, 1993). Thermo-mechanical models of
orogens generally predict metamorphism along clockwise P–T–t paths, and although counter-
clockwise P–T–t paths surely occur, in many of the reported paths the connection between the
prograde segment and the isobaric cooling segment is poorly justified, and tight clockwise loops
remain a possible alternative for this connecting segment (Williams and Karlstrom, 1996; Fig. 7).
The shape of the P–T–t path is non-unique in respect of different styles of thickening and
exhumation. Pre-1988, isothermal decompression paths were related to exhumation by erosion,
but post-1988 isothermal decompression paths have been related instead to tectonic exhumation
(Selverstone, 1994). Resolution between such competing models, however, may be achieved by
using field information, such as kinematic data, and consideration of the T–t evolution, and by
modelling the P–T–t evolution to distinguish probable paths, which fit the data, from impossible
paths (e.g. Platt et al., 1998).
The past decade has seen increased complexity in documented P–T–t paths as our ability to
extract quantitative information from rocks improved. Further, metamorphic belts that preserve
134 M. Brown / Journal of Geodynamics 32 (2001) 115–164

Fig. 7. Generalized styles of P–T paths for proterozoic rocks from Nevada, Arizona and New Mexico, south-west
USA (HTLP, low-P–T type metamorphism). Reproduced from Williams and Karlstrom 1996, Fig. 2) with permission
of the Geological Society of America.

evidence of more than one orogeny are better distinguished from those that evidence only a
monocyclic evolution. This has been accomplished by the development of better dating techni-
ques that yield precise ages from micro-domains in single crystals. As a result, some P–T–t paths
constructed in earlier work have been revealed to link data from superimposed metamorphic
events. Well-documented counterclockwise P–T–t paths are rare, although examples include
west-central New Hampshire (Spear et al., 1990, 1995), the Cooma complex (Australia) (Johnson
M. Brown / Journal of Geodynamics 32 (2001) 115–164 135

and Vernon, 1995), and the low-P-high-T metamorphism of the Vredefort Dome (Africa), which
is explained by intraplating of mafic and ultramafic magma in lithosphere over a mantle plume
(Gibson and Stevens, 1998).

4.2. P–T–t paths of subduction zone rocks

Clockwise P–T paths that vary from hairpins to loops according to the degree of refrigeration or
heating during exhumation are characteristic of high-P/T and ultrahigh-P (UHP) metamorphism. As
correctly identified by Miyashiro (1961), high-P/T type metamorphism is related to the subduction of
oceanic lithosphere, and ultrahigh-P metamorphism involves the elimination of an ocean basin and
partial subduction of continental lithosphere during terminal collision. Thus, both high-P/T and
ultrahigh-P metamorphism are diagnostic of the presence of a former ocean basin and of subduction.
P–T–t paths of subduction zone rocks have been grouped into two types (Ernst, 1988; Mar-
uyama et al., 1996). These types are paths that involve rapid, nearly isothermal decompression,
called ‘low-P back-reaction pervasive’ or A-type, including UHP metamorphism, and associated
with collision orogenic belts; and, those paths that suggest exhumation approximately retraced
the prograde P–T path, called ‘low-P back-reaction absent’ or B-type and associated with non-
collision orogenic belts. Recently, it has been suggested that different parts of the same subduc-
tion complex may record different retrograde P–T paths (Wallis and Aoya, 2000; Wallis et al.,
2000), which contradicts the postulate that subduction complexes are of two fundamentally dif-
ferent types (Ernst, 1988). These authors relate the different retrograde P–T paths shown by
subduction complexes to the thickness of the unit being exhumed and its rate of ascent. They
argue that thin sheets rising slowly have time to equilibrate thermally, which leads to similar ret-
rograde and prograde P–T paths, whereas thick sheets rising faster will follow a decompression
segment before the retrograde segment of the P–T path. Thus, if a single subduction complex
with similar rates of exhumation shows both types of path but in different parts of the complex,
this may reflect variation in the thickness of the nappe that is being exhumed.

4.2.1. Seismic consequences of subduction zone metamorphism


Subduction zones are sites of earthquakes. In zones of continental deformation, earthquakes
occur at relatively shallow depths (<30 km), but in subduction zones, earthquakes may occur as
deep as 670 km. At these depths, earthquakes must be triggered by some mechanism other than
brittle failure. It is likely that intermediate-depth earthquakes ( 30–300 km) and deep earth-
quakes ( 300–670 km) are triggered by metamorphic reactions (e.g. Green and Burnley, 1989;
Kirby et al., 1991, 1996a,b; Peacock and Wang, 1999). Intermediate-depth earthquakes most
likely are triggered by dehydration reactions associated with the transformation of amphibolite to
eclogite in the crust of the subducting lithosphere slab (Kirby et al., 1996a; Peacock and Wang,
1999; Rushmer, 1996). Deep earthquakes occur within the interior of the subducting lithosphere
slab, and are most likely triggered by the metastable transformation of olivine to spinel (Green
and Burnley, 1989; Kirby et al., 1991, 1996b).
The depth at which dehydration reactions or metastable transformation of olivine to spinel
occurs is dependent upon the geothermal gradient (Peacock, 1996). For example, in south-wes-
tern Japan, arc volcanism is rare and subduction zone earthquakes do not extend deeper than
65 km. In contrast, in north-eastern Japan, arc volcanism is common, and subduction zone
136 M. Brown / Journal of Geodynamics 32 (2001) 115–164

earthquakes extend to  200 km depth. This difference is attributed to a difference in thermal


structure of the subducting slab (Peacock and Wang, 1999), where the subducting slab beneath
north-eastern Japan has a lower geothermal gradient than the slab beneath south-western Japan.
As a consequence, dehydration reactions and eclogite-forming reactions occur deeper in the
‘cooler’ slab beneath north-eastern Japan than in the ‘warmer’ slab beneath south-western Japan.

4.3. Are different types of P–T–t path of suture zones truly diagnostic of a particular tectonic
process?

During crustal metamorphism uniqueness rather than conformity is the rule (Fig. 8). We still
recognize isothermal decompression and isobaric cooling, although these terms only refer to the
immediate post-peak part of the retrograde path (Harley, 1989, 1998). Isothermal decompression
paths include variants with heating (e.g. Platt et al., 1998; Soto and Platt, 1999) and cooling (e.g.
Albarède, 1976; Brown and Earle, 1983; Jamieson et al., 1998). Also, there are stepped decom-
pression-cooling paths (e.g. Brown and Dallmeyer, 1996; Brown and Raith, 1996; Touret, 2001),
wiggly paths (e.g. Bakker et al., 1989; Pattison et al., 1999) and decompression paths with thermal
spikes (e.g. De Yoreo et al., 1989; Whitney and Dilek, 1998). Looping P–T–t paths (Fig. 7)
represent a variant of the isobaric cooling type, the clockwise loop connecting the prograde seg-
ment to the isobaric cooling segment ending many former counterclockwise terranes (Schenk,
1989; Williams and Karlstrom, 1996).
Isothermal loading – isothermal decompression P–T–t paths (Brown, 1996; Hiroi et al., 1998;
Whitney et al., 1999) are common in rocks exhumed from former arc settings (e.g. Fiordland
Complex, New Zealand; Abukuma Plateau, Japan; Coast Mountains—North Cascades Belt, N.
America). The origin of such P–T paths is controversial, with tectonic loading (Hiroi and Kishi,
1989; Brown, 1998b; Hiroi et al., 1998; Evans and Davidson, 1999; Whitney et al., 1999) and
magma loading (Brown and Walker, 1993; Brown, 1996) as alternative models. The majority view
leans toward thrust loading or pure shear thickening partly driven by folding or transpressive
deformation as the dominant mechanism of thickening.
Although it is clear that suture zone metamorphism is related to plate margin processes, unra-
veling the details of rifting, subduction, collision, slab detachment and lithosphere delamination
has proven difficult (Henk et al., 2000). Furthermore, tectonic processes such as orogenic collapse
(Dewey, 1988) coupled with partial melting and flow of thickened orogens (Vanderhaeghe et al.,
1999; Vanderhaeghe and Tessier, 2001) undoubtedly cause some of the complexities being iden-
tified in P–T–t path studies. Thus, although Barrovian regional metamorphism, the int.-P/T type
metamorphism of Miyashiro (1961), is characteristic of suture zones, commonly this type of
metamorphism is overprinted by low-P/T type metamorphism as part of the same tectonic cycle.
The role of advection of heat in magma during such a low-P/T type metamorphic overprint
remains uncertain, in spite of its popularity as an explanation. Thermal relaxation and collapse of
thickened orogens may lead to low-pressure granulite facies metamorphism superimposed on a
higher-P history as part of one continuous cycle, or obduction of ultrahigh-P metamorphic rocks
may predate a low-P/T type metamorphic overprint due to extensive magmatism after exhumation
to the upper crust. Indeed, a sequence of superimposed metamorphic events of int.-P/T type fol-
lowed by low-P/T type is recorded in the Himalaya (e.g. England et al., 1992; England and
Molnar, 1993; Zeitler et al., 1993; Whittington et al., 1998, 1999; Searle et al., 1999).
M. Brown / Journal of Geodynamics 32 (2001) 115–164 137

Fig. 8. P–T diagram of the anatectic zone, that region in P–T space above the wet solidus for crustal rocks in which
granite melt may be present. Reaction (1) is kyanite ! sillimanite, the Ms (muscovite) granite–wet solidus is from
Huang and Wyllie (1981); reaction (2) muscovite+albite+quartz ! K-feldspar+aluminum silicate+melt is from
Petö (1976) and marks the lower T limit of plagioclase reactions (Thompson and Tracy, 1979); reaction (3) muscov-
ite+plagioclase+quartz ! K-feldspar+sillimanite+biotite+melt is from Patiño Douce and Harris (1998); reaction
(4) biotite+plagioclase+aluminum silicate+quartz ! garnet+K-feldspar+melt is from LeBreton and Thompson
(1988); and reaction (5) biotite+plagioclase+quartz ! orthopyroxene+garnet+K-feldspar+melt is from Vielzeuf
and Montel (1994). The symbol Xm w is used to denote the mole fraction of H2O in the melt and is considered to equal
the activity of H2O in the melt (Burnham, 1979); isopleths of Xm v
w are from Thompson (1996). The symbol aw is used to
denote activity of H2O in the volatile phase, whereas a(H2O) refers to water activity in the environment; isopleths of avw
are from Thompson (1996). Schematic P–T paths: I. Isobaric heating–cooling path characteristic of deep-seated con-
tact metamorphism; II. Stepped clockwise path (based on Brown and Dallmeyer, 1996); and, III. Stepped clockwise
path at UHT conditions (based on Brown and Raith, 1996; Raith et al., 1997).
138 M. Brown / Journal of Geodynamics 32 (2001) 115–164

5. Comments on UHP metamorphism

The identification of metamorphic rocks of increasingly deeper origin is of considerable impor-


tance for understanding the geodynamics of lithosphere interactions. In particular, there are the
issues of preservation of UHP minerals in metamorphosed continental crust during exhumation,
and the exhumation process itself (Ernst and Peacock, 1996). Results of the analog mechanical
modelling by Chemenda et al. (1995, 1996; Fig. 9) have proven popular. In these models, subduc-
tion of continental crust (continental lithosphere) to several hundred kilometers depth is termi-
nated by failure of the crust at depths of several tens of kilometers and rapid buoyancy-driven
uplift of the subducted slice of crust (continental lithosphere). This buoyancy-driven response is
postulated to be sufficiently rapid in nature that UHP minerals survive the transport back to
average crustal depths, from which they are exhumed by rapid erosion or tectonic extension. The
necessity to subduct continental crust (lithosphere) to several hundred kilometers depth is a direct
consequence of the thermobarometric results retrieved from UHP rocks. In 1996, Dobrzhi-
netskaya and colleagues (Dobrzhinetskaya et al., 1996) proposed that the garnet peridotite of
Alpe Arami, Switzerland, was exhumed from astonishing depths of >300 kilometers. The
extreme pressure apparently recorded by these rocks raises the specter of tectonic overpressure.

Fig. 9. Sketch of mechanical model of highly compressional mode of continent subduction and rock exhumation.
Reproduced from Chemenda et al. (1995, Fig. 3) with permission of Elsevier Science.
M. Brown / Journal of Geodynamics 32 (2001) 115–164 139

In the past decade, papers by Mancktelow (1993, 1995), StZwe and Sandiford (1994) and Pet-
rini and Podladchikov (2000) have readdressed the question of whether in theory tectonic over-
pressure can be significant during deformation. Tectonic overpressure refers to pressures greater
than the lithostatic value at any given depth due to effect of tectonic deformation (Rutland, 1965;
Ernst, 1971b). For example, Petrini and Podladchikov (2000) investigate increases in pressure
above the lithostatic gradient that result from the contribution of horizontal stresses and flexural
vertical loads, the latter generated by the deflection of the upper crust and the mantle due to the
presence of topographic relief and a root, respectively. They conclude that in the upper crust and
uppermost mantle, where rocks are thought to deform brittlely, pressure gradients up to two
times the lithostatic gradient can be achieved.
Below the brittle-viscous transition, the pressure in a material (viscous fluid) in motion is different
from that in the material at rest. During motion, the pressure at a point is the mean stress, which, in
general, differs from the lithostatic value. The assumption of an Andesonian stress state may not
apply during deformation in tectonic environments such as transpressive high-strain zones (Robin
and Cruden, 1994) or subduction channels (Mancktelow, 1995). In these environments, it is possible
for all three principal stresses to be greater than the lithostatic value, leading to a high mean stress
and a high tectonic overpressure. However, the differential stress in this case can remain low, as
required by experimental studies (Brace and Kohlstedt, 1980; Kohlstedt et al., 1995).
The recent work by Mancktelow (1993, 1995), StZwe and Sandiford (1994) and Petrini and
Podladchikov (2000) suggests that in theory tectonic overpressures can be generated, in which
case the critical issue is whether these overpressures can be dynamically maintained so as to be
recorded by the rock. One consequence if tectonic overpressures can be maintained is that pressure
will not correlate with depth and the pressure gradient will exceed lithostatic. This has implications
for the origin of UHP metamorphic rocks, which may have formed at depths considerably shal-
lower than currently believed.

5.1. UHP metamorphism today

Coesite- and diamond-eclogite facies metamorphic rocks in ancient suture zones were mostly
derived from rocks of continental margin affinity. Coesite is now recognized from a wide dis-
tribution of orogenic belts, and is found in a variety of protoliths. Diamond occurs as an included
phase in rock of supracrustal origin, such as graphitic biotite schists and dolomitic marbles, in
several orogenic belts. Mineral assemblages and calculated peak P–T conditions require that these
rocks were subducted to depths of >100 km (P >3 GPa) and then returned to the surface, still
retaining evidence of their UHP formation. The geological record, supported by mechanical model-
ling, suggests that such UHP rocks can be formed by subduction of thinned continental margin
lithosphere beneath ophiolites or arcs (Chemenda et al., 1996) and in continent:¼continent collision
zones (Chemenda et al., 1995, 2000).
The type example of continent–continent collision is the Himalaya. Based on the geometry of
the seismically active Hindu Kush continental subduction zone and its restoration, assuming
current plate motions, Searle et al. (2001) present a model to explain how supracrustal rocks
deposited along the northwest Indian plate margin were dragged down to UHP depths. They
argue that the Hindu Kush provides a modern analog for the formation of coesite- and diamond-
eclogite facies rocks in ancient suture zones. In this region, seismic tomography is interpreted to
140 M. Brown / Journal of Geodynamics 32 (2001) 115–164

suggest that a northward dipping slab is present in the upper 600 km of the mantle under the
Hindu Kush, which is still attached to the Indian lithosphere (Van der Voo et al., 1999). A large
volume of black shale is present along the north-western plate margin of India, which would
provide a suitable graphitic protolith to produce diamonds at UHP metamorphic conditions.
Searle et al. (2001) posit that subduction of these supracrustal rocks was enabled because they
were anchored by the eclogitized leading edge of the thinned Indian lithosphere. Further, they
argue that coesite- and diamond-eclogite facies UHP metamorphism is occurring today at depth
along the Hindu Kush continental subduction zone.

5.2. The Alpe Arami enigma

In 1996, Dobrzhinetskaya et al. (1996) made the controversial proposal that the Alpe Arami
garnet peridotite records evidence, in the form of abundant FeTiO3 rods in olivine that implies an
origin at a depth >300 km (T 51600 K, P 510 GPa), which has provoked much subsequent
discussion and a search for corroborating evidence (Green et al., 1997, 2000). From the abun-
dance, morphology, crystallography, and topotaxy of these FeTiO3 rods, Dobrzhinetskaya et al.
(1996) argued that the inferred very high solubility of highly charged cations like Ti and Cr
represents a previously unrecognized mantle environment, and that pressures in excess of 10 GPa
were the most likely explanation of the observations. Using high-pressure experiments, Dobrzhi-
netskaya et al. (1999) have shown that olivine can dissolve very high concentrations of TiO2
(>0.6 wt.%) at P 510 GPa, consistent with their hypothesis. In addition, Bozhilov et al. (1999)
have discovered in the same rocks exsolution lamellae of clinoenstatite in diopside. These lamellae
contain antiphase domains, which indicate that the originally-precipitating phase was a C2/c
pyroxene. Additional geological and crystallographic observations reported by Bozhilov et al.
(1999) strongly suggest that the precipitating phase was high-pressure clinoenstatite, thereby
providing independent evidence of a minimum depth of origin of this massif of 250 km. Recently,
Green et al. (2000) have reported exsolution of SiO2 from omphacite of the Alpe Arami eclogite.
The extreme pressure conditions postulated by Dobrzhinetskaya et al. (1996, 1999) and Bozhi-
lov et al. (1999) have received support from the thermobarometric data of Brenker and Brey
(1997) and Paquin and Altherr (2000). However, Risold et al. (1996, 1997), Hacker et al. (1997),
Ulmer et al. (1998), Trommsdorff et al. (2000) and Nimis and Trommsdorff (2001) have disputed
the arguments in favor of the extreme pressures in spite of this corroborating evidence. The very
high-pressure estimates are ascribed to inaccuracies in thermobarometric evaluations that rely on
aluminum in orthopyroxene barometry and garnet–olivine and garnet–orthopyroxene thermo-
metry. Whatever the real depth of origin, and there is agreement on minimum P of >3 GPa,
these rocks were evidently brought to the surface quickly.
Other examples that imply very great depth of exhumation of mantle rocks and/or subduction
to and return from such depths are known from other continental collision terranes, including the
Sulu terrane of eastern China where Yang et al. (1993) estimate P 56 GPa. The recent discovery
of exsolution of pyroxene from garnet in the peridotites of the Western Gneiss terrane of Norway
(van Roermund and Drury, 1998; van Roermund et al., 2000) provides unambiguous evidence of
a very deep origin for these rocks (>185–200 km). Evidence of coesite– and diamond–eclogite
facies metamorphism in the Erzgebirge of Saxony (Massonne, 1998) and the Kokchetav Massif,
northern Kazakhstan (Ota et al., 2000) adds to the growing list of continental terranes exhumed
M. Brown / Journal of Geodynamics 32 (2001) 115–164 141

from >100 km. These examples of deep subduction and exhumation suggest that this process has
occurred multiple times during the Phanerozoic, which indicates that this phenomenon is a nor-
mal part of the process of continent–continent collision (Maruyama et al., 1996; Maruyama and
Parkinson, 2000). Whether such exhumation is a single- or multiple-step process is an important
question for future research.

5.3. Metamorphism in the Earth’s deep interior

In the subducted lithosphere slab, basaltic crust transforms to amphibolite and then to eclogite
by depths equivalent to 3–4 GPa (Peacock, 1996), and eclogite transforms to garnetite as the
components that form clinopyroxene are progressively dissolved into the garnet structure by 15
GPa. A fast-growing area of petrological studies concerns understanding the mineralogy and
mineral transformations of the Earth’s mantle (e.g. Agee, 1993; Harte and Harris, 1994; Hemley
and Mao, 1998; Kesson et al., 1998; Stoffler, 1997). Main areas of interest include the upper and
lower boundaries of the transition zone, marked by the 400 and 670 km discontinuities, and
the deep mantle beneath (Mao and Hemley, 1998). The boundaries of the transition zone corre-
spond to the transformation of olivine (b-phase) to wadsleyite (g-phase) and ringwoodite (g-
phase) plus majorite to Mg-perovskite plus wZstite, respectively. An interesting petrological issue
concerns the amount of water that is present in the mantle and how it might be stored. It appears
that water may be stored in small amounts in nominally anhydrous minerals (Bell and Rossman,
1992), and in larger amounts in hydrous phases such as amphibole, antigorite (and other serpentine
minerals), phlogopite, and hydrous-phase A (Thompson, 1992).
As with all areas of metamorphic petrology, it is advances in technology that are enabling
exciting new research about the mineralogy and mineral transformations of the deep mantle
(Hemley and Mao, 1998). These advances in technology have enabled high-pressure experiments
using diamond-anvil cell and multi-anvil high-pressure, high-temperature environments. Addi-
tionally, the development of in situ and ex situ characterization methods compatible with the
microscopic size of high-pressure samples has been a key element in progress, for example the
application of synchrotron radiation to high-pressure studies (Mao and Hemley, 1998).

6. High-T and UHT metamorphism

6.1. High-T metamorphism

Low-P/T type metamorphism requires a tectonic setting that allows enhanced heat flux, transient
advection of heat due to magma ascent, usually large internally generated heat, or some combination
of these features. One mechanism to accomplish high T at shallow depth is subduction of a spreading
ridge system, which process creates a slab window to allow transfer of sub-slab asthenospheric
mantle across the slab to create an enhanced heat flux beneath the overriding plate. This process is
important along subduction zones that develop subduction—accretion complexes, and drives
low-P/T type metamorphism and anatexis. Heat transport by rising magma is suggested by near-
trench igneous activity with characteristic contaminated MORB geochemistry. Features of the
Cretaceous Ryoke and Abukuma low-P/T type metamorphic belts in Japan are consistent with
142 M. Brown / Journal of Geodynamics 32 (2001) 115–164

the expectations of ridge subduction and slab window formation (Brown, 1998a,b). It has been
argued that diachroneity of the orogenic event along the metamorphic belt is one of the char-
acteristic features of triple junction interactions (Pavlis and Sisson, 1995). However, other tectonic
scenarios may generate a migrating high heat flow event, such as the progressive detachment of a
subducting slab along a complex collisional boundary, for example the Carpathian (Nemcok et al.,
1998; Wenzel et al., 1998) or Taiwan (Teng et al., 2000).
One widely accepted paradigm for regional granulite facies metamorphism is that it is caused by
intrusion of mafic magma beneath or into the crust (Bohlen, 1987). The amphibolite-to-granulite
facies rocks of the Ivrea Zone, northern Italy, are commonly cited as a classic example of such a
genetic relationship. Detailed examination of the evidence, however, suggests that final emplace-
ment of the mafic rocks occurred subsequent to the regional thermal maximum, with only a
spatially restricted contact melting event associated with emplacement (Barboza et al., 1999). The
important question in this case is: what triggered the mantle-melting event?

6.2. UHT metamorphism

Coexistence of sapphirine and quartz (Fig. 10), or high aluminum orthopyroxene coexisting
with garnet, sillimanite or sapphirine, indicates UHT metamorphic conditions (Hensen and
Harley, 1990; Harley, 1998). Recently, Harley and Motoyoshi (2000) have shown that peak tem-
peratures of at least 1120  C are recorded by the Al2O3 content in orthopyroxene in a sapphirine-
orthopyroxene-quartz granulite from the Tula Mountains of Enderby Land, East Antarctica. Suc-
cessively overprinted corona and symplectite reaction textures in Mg–Al granulites (Figs. 10 and 11)
have allowed them to be used to deduce reaction histories, and from these to infer P–T–t paths of
ultrahigh-T terranes (Harley, 1989, 1998; Brown and Raith, 1996; Moraes et al., 2001). However,
misreading the textural record due to back reaction with residual melt must be avoided (Kriegs-
man and Hensen, 1998) since crust under UHT metamorphic conditions generally is melt-bearing
(Hensen and Harley, 1990; Raith et al., 1997). Both post-peak isobaric cooling and isothermal
decompression are documented in different UHT granulites, but in the absence of well-con-
strained prograde P–T histories ambiguity remains about clockwise versus counterclockwise P–
T–t paths. Composite isothermal decompression–isobaric cooling–isothermal decompression ret-
rograde P–T paths may record periods of exhumation separated by a period of isostatic stability
(Brown and Raith, 1996; Raith et al., 1997).

7. Comments about heat

Modellers have had difficulty in getting enough heat to achieve peak temperatures demanded by
the rocks in Barrovian and high (and ultra-high) temperature metamorphism (e.g. discussions in
England and Thompson, 1986; Chamberlain and Sonder, 1990; Thompson and Connolly, 1995).
This problem led Jamieson et al. (1998) (see also Huerta et al., 1996, 1999; Gerdes et al., 2000) to
ask the question: where’s the heat? Although it is intuitively obvious that regional metamorphism
is controlled by interactions between mechanical and thermal processes, it is by no means obvious
what combination of heat production from crust and mantle sources, refrigeration by subduction
and cooling from the surface controls the temperature field, or how this may be modified by
M. Brown / Journal of Geodynamics 32 (2001) 115–164 143

Fig. 10. Photomicrograph of a reaction microstructure in a UHT granulite derived from a metapelite protolith (from
the Neoproterozoic Braziliano Belt of Brazil; Moraes et al., 2001). The reactants are quartz (low relief colorless phase
outside the mineral aggregate in the center of the field of view), sapphirine (blue–green phase inside quartz at bottom
left-hand side of the field of view, and inside sillimanite in the mineral aggregate in the left-center of field of view) and
garnet (high relief, colorless mineral forming the interior part of the right-hand half of the mineral aggregate in the center
of the field of view). The products of the reaction are high-Al orthopyroxene (pale green/pink mineral surrounding garnet
in the mineral aggregate in the center of the field of view) and sillimanite (high relief colorless mineral with a single clea-
vage inside orthopyroxene and surrounding sapphirine in the mineral aggregate in the left-center of the field of view). The
long dimension of the field of view is 6 mm. This reaction has a steep slope in P–T space, and most likely was crossed
during a cooling stage of the metamorphic evolution, but at UHT conditions around T of  1325 K.

advection during orogenesis (Jamieson et al., in press). Models that reproduce appropriate peak
metamorphic conditions have several common features: (1) the model must incorporate high heat
production or redistribute the high heat-producing material during the evolution or incorporate a
higher than average heat flux from the mantle; (2) maximum temperatures are achieved after
maximum pressures, during the decompression (exhumation) segment of the P–T evolution; (3)
P–T paths are clockwise; and, (4) a substantial part of the P–T evolution is above the wet solidus
for crustal rocks.

7.1. Thermal structure of orogens

The presence at the surface of metamorphic rocks that crystallized at pressures equivalent to
depths of 10, 20 or even 50 km, excluding UHP rocks that record extreme conditions equivalent
144 M. Brown / Journal of Geodynamics 32 (2001) 115–164

Fig. 11. Photomicrograph of a reaction microstructure in a UHT granulite derived from a metapelite protolith (from
the Neoproterozoic Braziliano Belt of Brazil; Moraes et al., 2001). The reactants are sillimanite (aggregate of high relief,
colorless mineral with a single cleavage in the bottom right-hand side of field of view), garnet (high relief, colorless
mineral in the upper and left-hand parts of the field of view) and sapphirine1 (blue–green phase in the center of field of
view, with opaque/dark green hercynitic spinel1) and the products of the reaction are cordierite (low relief colorless
mineral intergrown with fine-grained sapphirine and hercynitic spinel in the center of the field of view, and forming a
narrow moat adjacent to garnet), sapphirine2 (symplectite with cordierite) and hercynitic spinel2 (symplectite with cor-
dierite). A coarser symplectite of cordierite and hercynitic spinel develops preferentially at the inner side of the cordierite
moat (e.g. at the left-hand side of the symplectite). The long dimension of field of view is  4 mm. This reaction has a
shallow slope in P–T space, and most likely was crossed during a decompression stage of the metamorphic evolution.

to 100–150 km depth, implies considerable exhumation subsequent to crystallization. Further,


these metamorphic rocks are underlain by an average thickness of continental crust, which implies
an initial tectonic anomaly that thickened existing crust or created pristine over-thickened crust;
the relaxation of this anomaly was followed by erosion. The importance of erosion on the thermal
structure of orogens was first pointed out by England and Richardson (1977), who investigated
thermal evolution in three geodynamic environments: over-thrusting of continental crust; subduc-
tion of a sediment wedge; and, crustal thickening by magma addition. These authors conclude ‘‘. . .
convection of heat by uplift and erosion is as potent a means of energy supply to the upper litho-
sphere in this situation as are radiogenic heating or transfer of heat from beneath the lithosphere.
Consequently, rocks in eroded orogenic regions will generally reflect temperatures higher than
those predicted from the latter effects alone’’ (see also Zeitler et al., 2001). A final point empha-
sized by England and Richardson is the three dimensional reality of orogens, which are often
discussed and modelled in one dimension only (see also Richardson, 1970).
M. Brown / Journal of Geodynamics 32 (2001) 115–164 145

Metamorphism in orogens typically is syntectonic; it varies from int.-P/T (kyanite–sillimanite)


to low-P/T (andalusite–sillimanite) type. Clockwise P–T paths are characteristic (e.g. England
and Richardson, 1977; England and Thompson, 1986; Thompson et al., 1997; Jamieson et al.,
1998; Huerta et al., 1999). These paths result from the combined effects of: (1) material transfer
within the orogen that advects hot crust to shallow depths and heat conduction to the surface
during syntectonic erosion of the thickening orogen; (2) increased radioactive heat generation due
to thickening and redistribution of high heat-producing materials; (3) dissipation of mechanical
energy generated during deformation; and, (4) at temperatures higher than the solidus tempera-
ture, increased strain rates due to weakening in the presence of melt. The initial distribution of
radioactive heat production with depth, including the location of any anomalous heat-producing
layers, is an important variable (e.g. Jaupart and Provost, 1985; Chamberlain and Sonder, 1990;
Mildren and Sandiford, 1995; Sandiford et al., 1995). Recent modelling suggests shear heating
during thrusting may be important (Molnar and England, 1990; England and Molnar, 1993;
Harrison et al., 1997, 1998; Leloup et al., 1999; Nabelek and Liu, 1999).
Variables include the rate and amount of thickening, the amount and initial spatial distribution
of radiogenic heat producing materials, and the rate and amount of accretion or removal of mate-
rial by under-plating, extension and erosion. These processes strongly perturb the geothermal
gradient by displacing isotherms toward the surface to create a thermal antiform. This forms a
near-isothermal corridor along the core of the orogen (e.g. Beaumont et al., 1992; Royden, 1993a;
Huerta et al., 1996, 1998, 1999; Thompson et al., 1997; Jamieson et al., 1998; Batt and Braun, 1999;
Zeitler et al., 2001; Jamieson et al., in press). Because of this coupled mechanical and thermal
evolution of orogens, temperatures in the middle crust exceed those at which anatexis is initiated
in many common rock types (Brown and Solar, 1999). The effect of this feature is considered in
the following section.

7.2. Thermal buffering and pluton-driven metamorphism

Melting proceeds in a stepwise fashion, beginning at the wet solidus (Fig. 8), which is crossed at
middle to lower crustal depths during the prograde (increasing T) evolution. Melt production at
the wet solidus is limited by the amount of water-rich volatile phase present in the limited por-
osity remaining under amphibolite facies conditions. Muscovite dehydration is likely to be the
major melt-producing step in the evolution of many orogens (Patiño Douce, 1999), which is
generally crossed during the prograde (increasing T) to decompression (decreasing P) segments of
the P–T evolution. Muscovite dehydration melting is particularly important in the metapelitic
component of the supracrustal succession in orogens, and the amount of melt produced is generally
limited by the amount of muscovite in the protolith (Patiño Douce and Harris, 1998). High (and
ultra-high) temperature metamorphism may involve biotite dehydration, which is particularly
important for the greywacke component of supracrustal successions. Thus, temperature is buf-
fered sequentially by the latent heat of melting required at the wet solidus, and during muscovite
dehydration melting and biotite dehydration melting (Vielzeuf et al., 1990).
In contrast, temperatures at shallower structural levels in rocks below the solidus may be
buffered by crystallization of melt that has escaped from the melting zone at lower structural
levels. As a consequence, a popular paradigm for low-P/T metamorphism has been advection
of heat with migrating magma, the magma having been generated by anatexis in regions of
146 M. Brown / Journal of Geodynamics 32 (2001) 115–164

thickened continental crust (De Yoreo et al., 1989). The term ‘pluton-driven metamorphism’ can
be used to describe the metamorphic product of such a process; it refers to upper-crustal meta-
morphism in which the higher-grade metamorphic zones appear to mimic pluton contacts, the
implication being that the metamorphism is pluton-related (De Yoreo et al., 1991). Such a term is
more appropriate than ‘regional contact metamorphism’ (Kays, 1970) since magmatic under-
plating, which is a common paradigm for granulite facies metamorphism, also can be thought of
as regional contact metamorphism. There is no doubt that plutons can be an important con-
tributor to the driving force for low-P/T type metamorphism, particularly in the upper crust
where commonly there is a spatial relationship between plutons and isograds. The maximum
thermal effect of pluton-driven metamorphism in the upper crust occurs when intrusions are
contemporaneous, but significant heating will occur even when intrusions are temporally sepa-
rate (Hanson and Barton, 1989). If the granites are derived from crust, however, the first order
problem is what causes melting, and we should avoid following casually an implied cause and
effect.

7.3. An example: the Northern Appalachians

For illustrative purpose, I take as an example the Devonian Acadian metamorphism of the
Northern Appalachians, USA (see Brown and Solar, 1998a, b; Solar and Brown, 1999, 2000,
2001a,b). The Acadian metamorphic belt is characterized by elevated modern-day heat flow ( 65
m Wm2) and high heat production (3.5106 Wm3). Metamorphic field gradients suggest
high-to-moderate rates of temperature change during metamorphism, but reveal only small var-
iations in pressure. The stratigraphic sequence includes formations with high heat production,
which was a consequence of high U and Th contents fixed in strongly reduced sediments of the
precursor anoxic basin (Chamberlain and Sonder, 1990). Oblique translation during contrac-
tional deformation thickened the stratigraphic sequence and displaced isotherms toward the sur-
face, to create the thermal structure imaged by the ‘migmatite front’, essentially an isothermal
surface given the high-P/T slope of the beginning of melting in most crustal protoliths (Fig. 8).
Melt transport to progressively shallower levels by differential stress-induced processes and
buoyancy helps propagate the thermal corridor upward by advecting of heat into the upper crust
(Brown and Solar, 1999).
Once melt is trapped and crystallized in a pluton, its source can be traced using isotopic fin-
gerprinting, so that the volume of melt derived from different sources in the crust can be assessed.
This information can be used to determine if the thermal perturbation associated with high-tem-
perature metamorphism extended to the Moho, reflected in voluminous granite magmatism from
lower crust or mixed (crust and mantle) sources, or was damped in the lower crust. In western
Maine—eastern New Hampshire the absence of a significant volume of granite with a geochem-
ical signature indicating derivation from basement inferred to underlie the Central Maine belt
(e.g. Lathrop et al., 1996; Pressley and Brown, 1999; Brown and Pressley, 1999) is consistent with
calculated intermediate-to-low reduced heat flow from the lower crust and mantle (Chamberlain
and Sonder, 1990). This implies low thermal gradients in the lower crust under assumed granulite
facies conditions. Magma was sourced from the thickening supracrustal sequence. Thus, Acadian
orogenesis involved redistribution of energy and mass within the crust, rather than addition of
energy and mass by mantle processes.
M. Brown / Journal of Geodynamics 32 (2001) 115–164 147

8. Exhumation

The processes by which rocks become buried to depths of tens of kilometers and transformed
by metamorphism are understood to a first order, but the processes that return these rocks to the
surface are much less clear. The two principal mechanisms for exhumation of metamorphic rocks
to the Earth’s surface are erosion and tectonics (subduction, buoyancy and extension), but the
relative contribution of each, the rates at which these processes operate, and the depth ranges
over which they are effective remain uncertain. Erosion is fastest in areas of active tectonism (e.g.
Koons, 1995; Zeitler et al., 2001), and collision processes are more effective at exhuming rocks
than rifting processes. In an example from the Himalaya studied by Ganguly et al. (2000), meta-
morphic mineral equilibria and compositional zoning of garnet were used to derive rates of
exhumation by modelling. Initial fast rates of 15 mm Ma1 were followed by slower rates of
2 mm Ma1, where the change in rate is interpreted to reflect a change of process from tectonic
thinning to erosion and/or horizontal flow at shallow depth.
Erosion can exert an influence as important as thickening and extensional collapse in orogens,
and may play a critical role in the exhumation of UHP rocks (Chemenda et al., 1995, 1996).
Rubatto and Hermann (2001) have used an integrated petrological and geochronological study to
estimate rates of exhumation of 36 mm Ma1 and 16 mm Ma1 for the Dora Maira UHP meta-
morphic rocks of the Western Alps. They postulate exhumation to be due to buoyancy and nor-
mal faulting, rather than fast erosion, because there is no evidence of debris from the latter
process present in the geologic record until the final stages of exhumation. Hermann et al. (2001)
determined a similar rate of 18 mm Ma1 for exhumation of the deeply subducted continental
crust of the Kokchetav Massif in Kazakhstan. These rates are within the range for plate move-
ments. In another example, Bosse et al. (2000) have been able to date petrologically significant
minerals in eclogite-facies rocks of the Champtoceaux complex, in the Variscides of the Armor-
ican Massif. By integrating geochronology with petrology, these authors have demonstrated that
exhumation occurred early in orogenesis while plate convergence remained active. Important
questions in relation to exhumation include (Ring et al., 2000): is there a relation between exhu-
mation rate and exhumation process? Do UHP rocks preserve a reliable record of exhumation
history? Is there a relationship between exhumation of UHP metamorphic rocks and magmatism?
Although the problem posed by the Alpe Arami enigma and the truth of the P–T conditions
that are recorded by the garnet peridotite are important to resolve, these issues also pose a chal-
lenge concerning whether we are too limited by conventional thinking. In this context, it is worth
taking a moment to consider the following concluding remarks by Green et al. (2000) ‘‘Is it more
difficult to conceive (or model) exhumation from 150 km than 100 km? Is it more difficult to
imagine a mechanism that could be responsible for exhumation from 200 km? Or 300 km? To our
understanding, only buoyancy forces can achieve exhumation from such depths where rocks are
hot and ductile. Thus, arrival at the surface of a peridotite from great depths requires that a lar-
ger body of lower density was subducted to depths in excess of the last resting place of the peri-
dotite and rafted it to crustal levels.’’
In some orogens, there is clear evidence for melting during decompression (Whittington et al.,
1998, 1999), and melt probably has a role in enabling both uplift and exhumation of deep crust
(England and Molnar, 1993; Hollister, 1993; Brown and Dallmeyer, 1996) and UHP rocks
(Hermann et al., 2001). Furthermore, there is a clear spatial and temporal relationship between
148 M. Brown / Journal of Geodynamics 32 (2001) 115–164

plutonic magmatism and the development of metamorphic core complexes (e.g. Lister and Bald-
win, 1993; Hill et al., 1995), although the exact role that magma may play in the process remains
to be explained. It is an open question whether melt-producing perturbations arise from ‘tectonics
as usual’ or from special mantle events.
It is likely that extension plays an important role in exhumation, and this may produce a
characteristic imprint on the P–T–t path recorded by the rocks (e.g. Baldwin, 1996). High-grade
metamorphic terranes characterized by near-isothermal decompression during a single meta-
morphic cycle, commonly have experienced rapid post-metamorphic cooling, which has been
interpreted to imply rapid exhumation during extension (e.g. Hill et al., 1992; Brown and Dall-
meyer, 1996; Zeck, 1996; Searle et al., 1999; Chemenda et al., 2000). The driving force of a
regional extension may result from: (1) gravitational and thermal instability of thickened crust
(Sonder et al., 1987; Gaudemer et al., 1988), which leads to extensional collapse (Dewey, 1988;
Burg et al., 1994; Barboza et al., 1999; Handy et al., 1999); and, (2) removal of part of the litho-
sphere by delamination (Bird, 1978; Nelson, 1992), convective removal of the thermal boundary
layer (Houseman et al., 1981; Platt and England, 1994) or slab breakoff (Davis and von Blanck-
enburg, 1995; von Blanckenburg and Davis, 1995).
The imprint of any one of these processes will be recorded in rocks at the orogen, perhaps in
reaction textures (Brown, 1993) and closely spaced closure ages for different thermo-chron-
ometers (Baldwin et al., 1993). For example, Soto and Platt (1999) report evidence for widespread
disequilibrium and overstepping of metamorphic reactions in support of decompression in the
Alboran Sea Basin in the western Mediterranean. This is inferred to have been a rapid event,
consistent with metamorphism driven by an external heat source beneath an extensional basin,
and possibly the result of complete removal of the mantle lithosphere.

9. Concluding remarks

Twenty-five years ago the P–T field of metamorphic petrology stretched to  1200 K and  1
GPa ( 2 GPa?), now we routinely think in terms of 1500 K and 3 GPa ( 10 GPa?). Addi-
tionally, advances in technology have enabled access to P–T of the mantle transition zone and the
lower mantle. Thus, the field of metamorphic petrology now is extended to studies of mantle
minerals and mineral transformations.
Are we limited by the conservative element inherent in conventional thinking? Yes, it is more
difficult to imagine a mechanism for something beyond conventional limits. Nonetheless, our
advance during these 25 years in understanding the dynamic nature of the P–T–t–d evolution of
orogens is a significant step. Petrology has benefited from dramatic advances in both analytical
capabilities and quantitative methods to determine the depth-time evolution of rocks in orogens.
Well-characterized P–T–t-d paths represent an important link between petrology and tectonics,
and between small-scale interactions and orogeny.
Improvements in P–T information will come largely in the form of relative P and not absolute
P–T. Continuous advances in instrumentation and the wider availability of these new tools will
provide more data and more complete datasets about orogens. The single biggest improvement
will be the wider availability of temporal information and better quantification of rates of pro-
cesses. Indeed, the regional scale availability of P–T–t–d information across and along an orogen,
M. Brown / Journal of Geodynamics 32 (2001) 115–164 149

coupled with geophysical imaging and petrogenetic isotope geochemistry, will enable us to better
constrain the 4-D evolution of orogens, and thus discriminate among different possible tectonic
scenarios using more sophisticated models.
Scientific endeavor is linked to philosophy. In the reductionist approach that has served science
so well in the last century, complex systems are broken down into constituent simpler parts.
Properties of these simpler parts are used to explain the behavior of the whole system, and
interactions are assumed to be fully predictable from the properties of these parts. However,
reductionism is the wrong approach in pursuit of an understanding of complex systems! This
requires a deterministic approach, in which different parts of the system are linked and affect one
another in a synergistic manner. Thus, these effects involve feedback relations between parts of
the system, and they are important in complex self-organized systems such as orogens.
Given perfect knowledge of the initial condition, a deterministic dynamical system is perfectly
predictable; it is always predictable in the short term. But dynamical systems are continually
changing. They are neither completely deterministic nor completely stochastic, but exhibit both
characteristics. A deterministic dynamical system that is sensitive to initial conditions exhibits
effectively unpredictable long-term behavior. These systems are chaotic. Chaos is the study of
how simple systems can generate complicated behavior.
In contrast, complex systems such as orogens are spatially and temporally extended nonlinear
systems characterized by collective properties associated with the system as a whole, which
properties are different from the characteristic behaviors of the constituent parts. Complexity is
the study of how complicated systems can generate simple behavior, for example by self-organi-
zation. However, complexity is a function of combinations and interactions of the constituent
parts of the system; emergent properties must be explained at the level of their appearance,
because they cannot be predicted from properties of the separate parts of the system. Thus, in
complex self-organized systems such as orogens, the influence of unique histories must be
respected.
Within the context that orogens are complex self-organized systems with feedback relations,
metamorphic petrology is an essential tool available to the geodynamicist. But it is only a valu-
able tool if it forms part of an integrated study linked with structural geology and geochronology
on the one hand, and geochemistry and geophysics on the other. In the future, tectonic models
will use data assimilation methodology. Predictions from modelling will be tested and the results
will inform future models. This will lead to more refined predictions that will require more precise
data to discriminate among alternatives.
However, we should remember the advice of Box (1976) ‘‘Since all models are wrong, the sci-
entist must be alert to what is importantly wrong. It is inappropriate to be concerned about mice
when there are tigers abroad.’’ Quantification and modelling provide the guiding light, but the
essential data come from the field, the hand sample and the thin section, especially using various
types of microscopy, and bulk and in situ chemical analysis. Additionally, these data are the facts
against which our models must be evaluated.
For the future we must work within a context of the ‘big picture’ and understand that we are
dealing with complex systems. This is modern metamorphic petrology, and it is an essential part
of geodynamics.

‘‘Plus ça change, plus c’est la même chose!’’


150 M. Brown / Journal of Geodynamics 32 (2001) 115–164

Acknowledgements

Discussion with many people, most recently Paddy O’Brien and Gary Solar; my ideas are inevi-
tably entangled with their ideas, but any mistakes most likely are mine. Reviewers Tracy Rushmer
and Peter Treloar are thanked for constructive comments, Peter Treloar for advice on restructuring
the sequence of topics in the paper, and Wolf Jacoby for his editorial advice and patience.

References

Agee, C.B., 1993. Petrology of the mantle transition zone. Annual Reviews of Earth and Planetary Science 21, 19–41.
Albarède, F., 1976. Thermal models of post-tectonic decompression as exemplified by the Haut-Allier granulites,
Massif Central, France. Bulletin de la Société Géologique de France 18, 1023–1032.
Alsdorf, D., Nelson, D., 1999. Tibetan satellite magnetic low: evidence for widespread melt in the Tibetan crust?
Geology 27, 943–946.
Alsdorf, D., Brown, L., Nelson, K.D., Makovsky, Y., Klemperer, S., Zhou, W, 1998. Crust deformation of the Lhasa
Terrane, Tibet Plateau from Project INDEPTH Deep Seismic Reflection Profiles. Tectonics 17, 501–519.
Axen, G.J., Selverstone, J., Byrne, T., Fletcher, J.M., 1998. If the strong crust leads, will the weak crust follow? GSA
Today 8, 1–8.
Bakker, H.E., DeJong, K., Helmers, H., Biermann, C., 1989. The geodynamic evolution of the Internal Zone of the
Betic Cordilleras, south-east Spain: a model based on structural analysis and geothermobarometry. Journal of
Metamorphic Geology 7, 359–381.
Baldwin, S.L., 1996. Contrasting P–T–t histories for blueschists from the western Baja terrane and the Aegean; effects
of synsubduction exhumation and backarc extension. In: Bebout, G.E., School, D.W., Kirby, S.H., Platt, J.P. (Eds.),
Subduction top to bottom. Geophysical Monograph, Vol. 96, pp. 135–141.
Baldwin, S.L., Lister, G.S., Hill, E.J., Foster, D.A., McDougal, I., 1993. Thermochronologic constraints on the tec-
tonic evolution of active metamorphic core complexes, D’Entrecasteaux Islands, Papua New Guinea. Tectonics 12,
611–628.
Barboza, S.A., Bergantz, G.W., 2000. Metamorphism and anatexis in the Mafic Complex contact aureole, Ivrea Zone,
northern Italy. Journal of Petrology 41, 1307–1327.
Barboza, S.A., Bergantz, G.W., Brown, M., 1999. Regional granulite facies metamorphism in the Ivrea Zone: is the
Mafic Complex the smoking gun or a red herring? Geology 27, 447–459.
Barrow, G., 1893. On an intrusion of muscovite-biotite gneiss in the south-eastern Highlands of Scotland and its
accompanying metamorphism. Quarterly Journal of the Geological Society of London 49, 330–358.
Barrow, G., 1912. On the geology of lower Dee-side Southern Highland Border. Proceedings of the Geologists Asso-
ciation 23, 274–290.
Batt, G.E., Braun, J., 1997. On the thermomechanical evolution of compressional orogens. Geophysical Journal
International 128, 364–382.
Batt, G.E., Braun, J., 1999. The tectonic evolution of the Southern Alps, New Zealand: insights from fully thermally
coupled dynamical modelling. Geophysical Journal International 136, 403–420.
Beaumont, C., Fullsack, P., Hamilton, J., 1992. Erosional control of active compressional orogens. In: McClay, K.R.
(Ed.), Thrust Tectonics. Chapman and Hall, New York, pp. 1–18.
Beaumont, C., Ellis, S., Hamilton, J., Fullsack, P., 1996. Mechanical model for subduction-collision tectonics of
Alpine-type compressional orogens. Geology 24, 675–678.
Beaumont, C., Ellis, S., Pfeiffer, A., 1999. Dynamics of sediment subduction-accretion at convergent margins: short-
term modes, long-term deformation, and tectonic implications. Journal of Geophysical Research 104, 17573–17601.
Bell, D.R., Rossman, G.R., 1992. Water in Earth’s mantle: the role of nominally anhydrous minerals. Science 255,
1391–1396.
Bell, T.H., Johnson, S.E., Davies, B., Forde, A., Hayward, N., Wilkins, C., 1992. Porphyroblast inclusion-trail orien-
tation data: eppure non son girate. Journal of Metamorphic Geology 10, 295–307.
M. Brown / Journal of Geodynamics 32 (2001) 115–164 151

Berman, R.G., 1988. Internally consistent thermodynamic data for minerals in the system Na2O–K2O–CaO–MgO–
FeO-Fe2O3–Al2O3–SiO2–TiO2–H2O–CO2. Journal of Petrology 29, 445–522.
Bingen, B., Austrheim, H., Whitehouse, M., 2001. Ilmenite as a source of zirconium during high-grade metamorphism?
Textural evidence from the Caledonides of western Norway and implications for zircon geochronology. Journal of
Petrology 42, 355–376.
Bird, P., 1978. Initiation of intracontinental subduction in Himalaya. Journal of Geophysical Research 83, 4975–4987.
Bleeker, W., Ketchum, J., Jackson, V., Villeneuve, M., 1999a. The Central Slave Basement Complex, Part I: its struc-
tural topology and autochthonous cover. Canadian Journal of Earth Sciences 36 (7), 110–1083.
Bleeker, W., Ketchum, J., Davis, W., 1996b. The Central Slave Basement complex, Part II: age and tectonic sig-
nificance of high-strain zones along the basement-cover contact. Canadian Journal of Earth Sciences 36 (7), 1111–
1130.
Bohlen, S.R., 1987. Pressure–temperature–time paths and a tectonic model for the evolution of granulites. Journal of
Geology 95, 617–632.
Bosse, V., Ferrud, G., Ruffet, G., Ballèvre, M., Peucat, J.-J., De Jong, K., 2000. Late Devonian subduction and early-
orogenic exhumation of eclogite-facies rocks from the Champtoceaux complex (Variscan belt, France). Geological
Journal 35, 297–325.
Box, G.E.P., 1976. Science and statistics. Journal of the American Statistics Association 71, 791–799.
Bozhilov, K., Green II, H.W., Dobrzhinetskaya, L., 1999. Clinoenstatite in Alpe Arami peridotite: additional evidence
of very high pressure. Science 284, 128–132.
Brace, W.F., Kohlstedt, D.L., 1980. Limits on lithospheric stress imposed by laboratory experiments. Journal of
Geophysical Research 85, 6248–6252.
Brenker, F.E., Brey, G.P., 1997. Reconstruction of the exhumation path of the Alpe Arami garnet-peridotite body
from depths exceeding 160 km. Journal of Metamorphic Geology 15, 581–592.
Brouwer, F.M., 2000. Thermal evolution of high-pressure metamorphic rocks in the Alps. Geologica Ultraiectina 199,
221.
Brown, E.H., 1996. High-pressure metamorphism caused by magma loading in Fiordland, New Zealand. Journal of
Metamorphic Geology 14, 441–452.
Brown, E.H., Walker, N.W., 1993. A magma-loading model for the Barrovian metamorphism in the southeast Coast
Plutonic Complex, British Columbia and Washington. Geological Society of America Bulletin 105, 479–500.
Brown, M., 1983. The petrogenesis of some migmatites from the Presqu’ile de Rhuys, southern Brittany, France. In:
Atherton, M.P., Gribble, C.D. (Eds.), Migmatites, Melting and Metamorphism. Shiva Publishing Limited, Nant-
wich, pp. 174–200.
Brown, M., 1993. P–T–t evolution of orogenic belts and the causes of regional metamorphism. Journal of the Geolo-
gical Society 150, 227–241.
Brown, M., 1998a. Unpairing metamorphic belts: P–T paths and a tectonic model for the Ryoke Belt, southwest
Japan. Journal of Metamorphic Geology 16, 3–22.
Brown, M., 1998b. Ridge-trench interactions and high-T–low-P metamorphism, with particular reference to the Cre-
taceous evolution of the Japanese Islands. In: Treloar, P.J., O’Brien, P.J. (Eds.), What drives metamorphism and
metamorphic reactions. Geological Society of America Special Publication, Vol. 138, pp. 131–163.
Brown, M., 2001. Crustal melting and granite magmatism: key issues. Physics and Chemistry of the Earth 26, 201–212.
Brown, M., Dallmeyer, R.D., 1996. Rapid Variscan exhumation and role of magma in core complex formation:
Southern Brittany metamorphic belt, France. Journal of Metamorphic Geology 14, 361–379.
Brown, M., Earle, M.M., 1983. Cordierite—bearing schists and gneisses from Timor, eastern Indonesia: P–T condi-
tions of metamorphism and tectonic implications. Journal of Metamorphic Geology 1, 183–203.
Brown, M., O’Brien, P.J., 1997. Evolution of metamorphic belts: a changing view. In: Xianglin, Q., Zhendong, Y., Hall,
H.C. (Eds.), Precambrian Geology and Metamorphic Petrology. Proceedings of the 30th International Geological
Congress, VSP, Zeist, The Netherlands, Vol. 17, pp. 217–231.
Brown, M., Pressley, R.A., 1999. Crustal melting in nature: prosecuting source processes. Physics and Chemistry of the
Earth (A) 24 (2), 305–316.
Brown, M., Raith, M., 1996. First evidence of ultrahigh-temperature decompression from the granulite province of
Southern India. Journal of the Geological Society 153, 819–822.
152 M. Brown / Journal of Geodynamics 32 (2001) 115–164

Brown, M., Rushmer, T., 1997. The role of deformation in the movement of granite melt: views from the laboratory
and the field. In: Holness, M.B. (Ed.), Deformation-enhanced Fluid Transport in the Earth’s Crust and Mantle. The
Mineralogical Society Series: 8. Chapman and Hall, London, pp. 111–144.
Brown, M., Solar, G.S., 1998a. Shear zone systems and melts: feedback relations and self-organization in orogenic
belts. Journal of Structural Geology 20, 211–227.
Brown, M., Solar, G.S., 1998b. Granite ascent and emplacement during contractional deformation in convergent
orogens. Journal of Structural Geology 20, 1365–1393.
Brown, M., Solar, G.S., 1999. The mechanism of ascent and emplacement of granite magma during transpression: a
syntectonic granite paradigm. Tectonophysics 312, 1–33.
Brown, M., Solar, G.S., 2000. Feedback relations between deformation and melt, the evolution from weakening to
hardening in transpressive orogens. Journal of the Czechoslovakian Geological Society 45, 215–216.
Brown, M.A., Brown, M., Carlson, W.D., Denison, C., 1999. Topology of syntectonic melt-flow networks in the deep
crust. American Mineralogist 84, 1793–1818.
Burg, J.P., Van den Driessche, J., Brun, J.P., 1994. Syn-thickening to post-thickening extension—Mode and con-
sequences. Compte Redus de l’Academie de Sciences, II, 319, 1019–1032..
Burman, R.G., 1991. Thermobarometry using multi-equilibrium calculations: a new technique with petrological
applications. Canadian Mineralogist 29, 833–855.
Burnham, C.W., 1979. Magmas and hydrothermal fluids. In: Barnes, H.L. (Ed.), Geochemistry of Hydrothermal Ore
Deposits, 2nd Edition. Wylie-Interscience, New York, pp. 71–136.
Carlson, W.D., Denison, C., 1992. Mechanisms of porphyroblast crystallization: results from high-resolution com-
puted X-ray tomography. Science 257, 1236–1239.
Chamberlain, C.P., Sonder, L.J., 1990. Heat-producing elements and the thermal and baric patterns of metamorphic
belts. Science 250, 763–769.
Chemenda, A.I., Burg, J.-P., Mattauer, M., 2000. Evolutionary model of the Himalaya–Tibet system: geopoem based
on new modelling, geological and geophysical data. Earth and Planetary Science Letters 174, 3–4.
Chemenda, A.I., Mattauer, M., Bokun, A.N., 1996. Continental subduction and a mechanism for exhumation of high-
pressure metamorphic rocks: new modelling and field data from Oman. Earth and Planetary Science Letters 143,
173–182.
Chemenda, A.I., Mattauer, M., Malavieille, J., Bokun, A.N., 1995. A mechanism for syn-collisional rock exhumation
and associated normal faulting: Results from physical modelling. Earth and Planetary Science Letters 132, 225–232.
Chernoff, C.B., Carlson, W.D., 1999. Trace element zoning as a record of chemical disequilibrium during garnet
growth. Geology 27, 555–558.
Clark, M.K., Royden, L.H., 2000. Topographic ooze: building the eastern margin of Tibet by lower crustal flow.
Geology 28, 703–706.
Cocherie, A., Legendre, O., Peucat, J.J., Kouamelan, A., 1998. Geochronology of polygenic monazites constrained by
in situ electron microprobe Th–U-total Pb determination: implication for Pb behavior in monazite. Geochimica et
Cosmochimica Acta 62, 2475–2497.
Crowley, J.L., Ghent, E.D., Carr, S.D., Simony, P.S., Hamilton, M.A., 2000. Multiple thermotectonic events in a
continuous sequence, Mica Creek area, southeastern Canadian Cordillera. Geological Materials Research 2, 21–45.
Cruden, A.R., 1998. On the emplacement of tabular granites. Journal of the Geological Society 155, 853–862.
Daniel, C.G., Spear, F.S., 1999. The clustered nucleation and growth process of garnet in regional metamorphic rocks
from northwest Connecticut, USA. Journal of Metamorphic Geology 17, 503–520.
Davies, G.F., 1999. Dynamic Earth: Plates, Plums and Mantle Convection. Cambridge University Press, Cambridge.
p. 458.
Davies, G.R., Tommasini, S., 2000. Isotopic disequilibrium during rapid crustal anatexis: implications for petrogenetic
studies of magmatic processes. Chemical Geology 162, 169–191.
Davies, J.H., von Blanckenburg, F., 1995. Slab breakoff: a model of lithosphere detachment and its test in the mag-
matism and deformation of collisional orogens. Earth and Planetary Science Letters 129, 85–102.
Dempster, T.J., 1984. Localized uplift in the Scottish Dalradian. Nature 307, 156–159.
Denison, C., Carlson, W.D., 1997. Three-dimensional quantitative textural analysis of metamorphic rocks using high-
M. Brown / Journal of Geodynamics 32 (2001) 115–164 153

resolution computed X-ray tomography: Part II: application to natural samples. Journal of Metamorphic Geology
15, 45–57.
Denison, C., Carlson, W.D., Ketchum, R.A., 1997. Three-dimensional quantitative textural analysis of metamorphic
rocks using high-resolution computed X-ray tomography. Part I: methods and techniques. Journal of Metamorphic
Geology 15, 29–44.
Dewey, J.F., 1975. Finite plate implications: some implications for the evolution of rock masses at plate margins.
American Journal of Science 275A, 260–284.
Dewey, J.F., 1980. Episodicity, sequence, and style at convergent plate boundaries. In: Strangway, D.W. (Ed.), The
Continental Crust and Its Mineral Deposits. Geological Association of Canada Special Paper 20, pp. 553–573.
Dewey, J.F., 1988. Extensional collapse of orogens. Tectonics 7, 1123–1139.
Dewey, J.F., Bird, J.M., 1970. Mountain belts and the new global tectonics. Journal of Geophysical Research 75,
2625–2647.
De Yoreo, J.J., Lux, D.R., Guidotti, C.V., 1989. The role of crustal anatexis and magma migration in regions of
thickened continental crust. In: Daly, J.S., Cliff R.A., Yardley B.W.D. (Eds.), Evolution of Metamorphic Belts.
Geological Society of America Special Publication 43, pp. 187–202.
De Yoreo, J.J., Lux, D.R., Guidotti, C.V., 1991. Thermal modeling in low-pressure/high temperature metamorphic
belts. Tectonophysics 188, 209–238.
Dobrzhinetskaya, L., Bozhilov, K.N., Green II, H.W., 1999. The solubility of TiO2 in olivine: implications for the
mantle wedge environment. Chemical Geology 160, 357–370.
Dobrzhinetskaya, L.F., Green II, H.W., Wang, S., 1996. Alpe Arami: a peridotite massif from depths of more than 300
kilometers. Science 271, 1841–1845.
Ellis, S., Beaumont, C., 1999. Models of convergent boundary tectonics: implications for the interpretation of
Lithoprobe data. Canadian Journal of Earth Sciences 36, 174–1711.
Ellis, S., Beaumont, C., Jamieson, R.A., Quinlan, G., 1998. Continental collision including a weak zone: the vise model
and its application to the Newfoundland Appalachians. Canadian Journal of Earth Sciences 35, 1323–1346.
Ellis, S., Beaumont, C., Pfiffner, O.A., 1999. Geodynamic models of crustal-scale episodic tectonic accretion and
underplating in subduction zones. Journal of Geophysical Research 104, 15169–15190.
Ellis, S., Wissing, S., Pfiffner, A., 2000. Strain localization as a key to reconciling experimentally derived flow-law data
with dynamic models of continental collision. International Journal of Earth Science. DOI 10.1007/s005310000151.
England, P., Houseman, G., 1989. Extension during continental convergence, with application to the Tibetan Plateau.
Journal of Geophysical Research 94, 17561–17579.
England, P., Molnar, P., 1993. Cause and effect among thrust and normal faulting anatectic melting and exhumation in
the Himalaya. In: Treloar, P.J., Searle, M.P. (Eds.), Himalayan Tectonics. Geological Society of America Special
Publication 74, pp. 401–411.
England, P.C., Richardson, S.W., 1977. The influence of erosion upon the mineral facies of rocks from different
metamorphic environments. Journal of the Geological Society 134, 201–213.
England, P.C., Thompson, A., 1986. Some thermal and tectonic models for crustal melting in continental collision
orogens. In: Coward, M.P., Ries, A.C. (Eds.), Collision Tectonics. Geological Society of America Special Publication
19, pp. 83–94.
England, P., Le Fort, P., Molnar, P., Pêcher, A., 1992. Heat sources for tertiary metamorphism and anatexis in the
Annapurna-Manaslu region, central Nepal. Journal of Geophysical Research 97, 2107–2128.
Ernst, W.G., 1971a. Metamorphic zonations on presumably subducted lithospheric plates from Japan, California and
the Alps. Contributions to Mineralogy and Petrology 34, 43–59.
Ernst, W.G., 1971b. Do mineral parageneses reflect unusually high pressure conditions of Franciscan metamorphism?
American Journal of Science 270, 81–108.
Ernst, W.G., 1988. Tectonic history of subduction zones inferred from retrograde blueschist P–T paths. Geology 16,
1081–1084.
Ernst, W.G., Peacock, S.M., 1996. A thermotectonic model for preservation of ultrahigh-pressure phases in meta-
morphosed continental crust. In: Bebout, G.E., Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction Top to
Bottom. Geophysical Monograph 96, pp. 171–178.
154 M. Brown / Journal of Geodynamics 32 (2001) 115–164

Eskola, P., 1915. On the relations between the chemical and mineralogical composition in the metamorphic rocks of
the Orijärvi region. Bull. Comm. Geol. Finlande No. 44, English summary, pp. 109–145.
Eskola, P., 1920. The mineral facies of rocks. Norsk Geologisk Tidsskrift 6, 143–194.
Essex, R.M., Gromet, L.P., 2000. U–Pb dating of prograde and retrograde titanite growth during the Scandian oro-
geny. Geology 28, 419–422.
Evans, B.W., Davidson, G.S., 1999. Kinetic control of metamorphic imprint during synplutonic loading of batholiths:
an example from Mt. Stuart, Washington. Geology 27, 415–418.
Finger, F., Broska, I., Roberts, M.P., Schermaier, A., 1998. Replacement of primary monazite by apatite–allanite–
epidote coronas in an amphibolite facies granite gneiss from the eastern Alps. American Mineralogist 83, 248–258.
Fitzsimmons, I.C.W., Kinny, P.D., Harley, S.L., 1997. Two-stages of zircon and monazite growth in anatectic leu-
cogneiss: SHRIMP constraints on the duration and intensity of Pan-African metamorphism in Prydz Bay, east
Antarctica. TERRA Nova 9, 47–51.
Flesch, L.M., Holt, W.E., Haines, A.J., Shen-Tu, B., 2000. Dynamics of the Pacific–North American Plate Boundary
in the western United States. Science 287, 834–836.
Flesch, L.M., Haines, A.J., Holt, W.E. 2001. Dynamics of the India–Eurasia collision zone. Journal of Geophysical
Research. 106, 16435–16460.
Foster, G., Kinny, P., Vance, D., Prince, C., Harris, N., 2000. The significance of monazite U–Th–Pb age data in
metamorphic assemblages; a combined study of monazite and garnet chronometry. Earth and Planetary Science
Letters 181, 327–340.
Franke, W., Haak, V., Oncken, O., Tanner, D., 2000. Orogenic processes: quantification and modeling in the Variscan
belt. Geological Society Special Publication 179.
Frazier, G., McDougall, I., Ellis, D.J., Williams, I.S., 2000. Timing and rate of isothermal decompression in Pan-
African granulites from Rundvagshetta, east Antarctica. Journal of Metamorphic Geology 18, 441–454.
Friend, C.R.L., Nutman, A.P., Baadsgaard, H., Kinny, P.D., McGregor, V.R., 1996. Timing of late Archaean terrane
assembly, crustal thickening and granite emplacement in the Nuuk region, southern West Greenland. Earth and
Planetary Science Letters 142 (3-4), 353–365.
Ganguly, J., Dasgupta, S., Cheng, W., Neogi, S., 2000. Exhumation history of a section of the Sikkim Himalayas,
India: records in the metamorphic mineral equilibria and compositional zoning of garnet. Earth and Planetary
Science Letters 183, 471–486.
Gaudemer, Y., Jaupart, C., Tapponnier, P., 1988. Thermal control on post-orogenic extension in collision belts. Earth
and Planetary Science Letters 89, 48–62.
Gerdes, A., Wörner, G., Henk, A., 2000. Post-collisional granite generation and HT–LP metamorphism by radiogenic
heating: the Variscan South Bohemian Batholith. Journal of the Geological Society 157, 577–587.
Gibson, R.L, Stevens, G., 1998. Regional metamorphism due to anorogenic intracratonic magmatism. In: Treloar, P.J,
O’Brien, P.J. (Eds.), What drives metamorphism and metamorphic reactions? Geological Society of America Special
Publication 138. , pp. 121–135.
Goldschmidt, V.M., 1911. Die Kontaktmetamorphose im Kristianiagebiet. Oslo Vidensk. Skr., I, Math.-Nat. Kl. #11.
Gordon, T.M., 1992. Generalized thermobarometry: solution of the inverse chemical equilibrium problem using data
for individual species. Geochimica et Cosmochimica Acta 56, 1793–1800.
Green II, H.W., Burnley, P.C., 1989. A low self-organizing mechanism for deep-focus earthquakes. Nature 341, 733–
737.
Green II, H.W., Dobrzhinetskaya, L., Bozhilov, K., 1997. Determining the origin the ultra-high pressure lherzolites
(Response). Science 278, 704–707.
Green II, H.W., Dobrzhinetskaya, L., Bozhilov, K.N., 2000. Mineralogical and experimental evidence for very deep
exhumation from subduction zones. Journal of Geodynamics 30, 61–76.
Grove, M., Harrison, T.M., 1999. Monazite Th–Pb age depth profiling. Geology 27, 487–490.
Grubenmann, U., 1906. Die Kristanninen Schiefer.
Hacker, B., Sharp, T., Zhang, R.Y., Liou, J.G., Hervig, R.L., 1997. Determining the origin of ultrahigh-pressure
lherzolites (Discussion). Science 278, 702–704.
Hamilton, W.B., 1998. Archean magmatism and deformation were not products of plate tectonics. Precambrian
Research 91, 143–179.
M. Brown / Journal of Geodynamics 32 (2001) 115–164 155

Handy, M.R., 4 others, 1999. Multistage accretion and exhumation of the continental crust (Ivrea crustal section, Italy
and Switzerland). Tectonics 18, 1154–1177.
Handy, M.R., 9 others, 2001. Rheology and geodynamic modelling: the next step forward. International Journal of
Earth Science. 90, 149–156.
Hanson, R.B., Barton, M.D., 1989. Thermal development of low-pressure metamorphic belts: Results from two-
dimensional numerical models. Journal of Geophysical Research 94, 10363–10377.
Harker, A., 1919. Anniversary Address to the Geological Society. Quarterly Journal of the Geological Society of
London 24, 63–75.
Harley, S.L., 1989. The origins of granulites: a metamorphic perspective. Geological Magazine 126, 215–247.
Harley, S.L., 1998. On the occurrence and characterization of ultrahigh-temperature crustal metamorphism. In: Tre-
loar, P.J., O’Brien, P.J. (Eds.), What Drives Metamorphism and Metamorphic Reactions? Geological Society of
America Special Publication 138, pp. 81–107.
Harley, S.L., Motoyoshi, Y., 2000. Al zoning in orthopyroxene in a sapphirine quartzite: evidence for >1120 C Uht
metamorphism in the Napier Complex, Antarctica, and implications for the entropy of sapphirine. Contributions to
Mineralogy and Petrology 138, 293–307.
Harris, N., Vance, D., Ayres, M., 2000. From sediment to granite: timescales of anatexis in the upper crust. Chemical
Geology 162, 155–167.
Harrison, T.M., Grove, M., Lovera, O.M., Katlos, E.J., 1998. A model for the origin of Himalayan anatexis and
inverted metamorphism. Journal of Geophysical Research- Solid Earth 103 (B11), 27017–27032.
Harrison, T.M., Lovera, O.M., Grove, M., 1997. New insights into the origin of two contrasting Himalayan belts.
Geology 25, 899–902.
Harte, B., Harris, J.W., 1994. Lower mantle mineral associations preserved in diamonds. Mineralogical Magazine 58,
384–385.
Haughton, S., 1856. On slatey cleavage and the distortion of fossils. Philosophical Magazine 12, 409–421.
Hawkins, D.P., Bowing, S.A., 1997. New-Pb systematics of monazite and xenotime: case studies from the Paleopro-
terozoic of the Grand Canyon, Arizona. Contributions to Mineralogy and Petrology 127, 87–103.
Hawkins, D.P., Bowring, S.A., 1999. U–Pb monazite, xenotime and titanite geochronological constraints on the pro-
grade to post-peak metamorphic thermal history of Paleoproterozoic migmatites from the Grand Canyon, Arizona.
Contributions to Mineralogy and Petrology 134, 150–169.
Heilbronner, R., 2000. Automatic grain boundary detection and grain size analysis using polarization micrographs or
orientation images. Journal of Structural Geology 22, 969–981.
Helgeson, H.C., Delaney, J.M., Nesbitt, H.W., Bird, D.K., 1978. Summary and critique of the thermodynamic prop-
erties of rock-forming minerals. American Journal of Science 278A, 1–299.
Hemley, R.J., Mao, H.-K., 1998. Ultra high-pressure mineralogy. Reviews in Mineralogy, 37.
Henk, A., 1995. Late Variscan exhumation histories of the southern Rhenohercynian zone and western mid-German
crystalline rise—results from thermal modeling. Geologische Rundschau 84, 578–590.
Henk, A., 1997. Gravitational orogenic collapse vs. plate boundary stresses: a numerical modeling approach to the
Permo-Carboniferous evolution of central Europe. Geologische Rundschau 84, 39–55.
Henk, A., 1999. Did the Variscides collapse or were they torn apart? A quantitative evaluation of the driving forces for
postconvergent extension in central Europe. Tectonics 18, 774–792.
Henk, A., 2000. Foreland-directed lower-crustal flow and its implications for the exhumation of high-pressure—high-
temperature rocks. In: Franke, W., Haak, V., Oncken, O., Tanner, D. (Eds.), Geological Society Special Publication
179.
Henk, A., Franz, L., Teufel, S., Oncken, O., 1997. Magmatic underplating, extension, and crustal equilibration: insights
from a cross section through the Ivrea Zone and Strona Ceneri Zone, northern Italy. Journal of Geology 105, 367–377.
Henk, A., von Blanckenburg, F., Finger, F., Schaltegger, U., Zulauf, G., 2000. Syn-convergent high-temperature
metamorphism and magmatism in the Variscides—a discussion of potential heat sources. In: Franke, W., Altherr,
R., Haak, V., Oncken, O., Tanner, D. (Eds.), Orogenic Processes—Quantification and Modeling in The Variscan
Belt Of Central Europe. Geological Society Special Publication 179, pp. 387–399.
Hensen, B.J., Harley, S.L., 1990. Graphical analysis of P–T–X relations in granulite facies metapelites. In: Ashworth, J.R.,
Brown, M. (Eds.), High Temperature Metamorphism and Crustal Anatexis. Unwin Hyman, London, pp. 19–56.
156 M. Brown / Journal of Geodynamics 32 (2001) 115–164

Hermann, J., Rubatto, D., Korsakov, A., Shatsky, V.S., 2001. Multiple zircon growth during fast exhumation of dia-
mondiferous, deeply subducted continental crust (Kokchetav Massif, Kazakhstan). Contributions to Mineralogy and
Petrology. 141, 66–82.
Hill, E.J., Baldwin, S.L., Lister, G.S., 1992. Unroofing of active metamorphic core complexes in the D’Entrecasteaux
Islands, Papua New Guinea. Geology 20, 907–910..
Hill, E.J., Baldwin, S.L., 1995. Magmatism as an essential driving force for formation of active metamorphic core
complexes in Guinea. Journal of Geophysical Research, B. Solid Earth and Planets 100 (6), 10441–10451.
Hiroi, Y., Kishi, S., 1989. P–T evolution of Abukuma metamorphic rocks in north-east Japan: metamorphic evidence
for oceanic crust obduction. In: Daly, J.S., Cliff, R.A., Yardley, B.W.D. (Eds.), Evolution of Metamorphic Belts.
Geological Society of America Special Publication 43, pp. 481–486.
Hiroi, Y., Kishi, S., Nohara, T., Sato, K., Gotto, J., 1998. Cretaceous high-temperature rapid loading and unloading in
the Abukuma metamorphic terrane, Japan. Journal of Metamorphic Geology 16, 67–81.
Hodges, K.V., 1998. The thermodynamics of Himalayan orogenesis. In: Treloar, P.J., O’Brien, P.J. (Eds.), What
Drives Metamorphism and Metamorphic Reactions? Geological Society of America Special Publication 138, pp. 7–
22.
Hodges, K.V., 2000. Tectonics of the Himalayas and southern Tibet from two perspectives. Geological Society of
America Bulletin 112, 324–350.
Holland, T.J.B., Powell, R., 1985. An internally consistent thermodynamic data set with uncertainties and correlations:
2: data and results. Journal of Metamorphic Geology 3, 343–370.
Holland, T.J.B., Powell, R., 1990. An enlarged and updated internally consistent thermodynamic data set with uncer-
tainties and correlations: the system K2O–Na2O–CaO–MgO–MnO–FeO–Fe2O3–Al2O3–TiO2–SiO2–C–H2–O2. Jour-
nal of Metamorphic Geology 8, 89–124.
Holland, T.J.B., Powell, R., 1998. An internally consistent thermodynamic data set for phases of petrological interest.
Journal of Metamorphic Geology 16, 309–343.
Hollister, L.S., 1993. The role of melt in the uplift and exhumation of orogenic belts. Chemical Geology 108, 31–48.
Horn, I., Rudnick, R.L., McDonough, W.F., 2000. Precise elemental and isotope ratio determination by simultaneous
solution nebulization and laser ablation-ICP–MS: application to U–Pb geochronology. Chemical Geology 167, 405–
425.
Houseman, G.A., McKenzie, D.P., Molnar, P., 1981. Convective instability of a thickened boundary layer and its
relevance for the thermal evolution of continental convergent belts. Journal of Geophysical Research 86, 6116–6132.
Huang, W.L., Wyllie, P.J., 1981. Phase relationship of S-type granite with H2O to 35 kbar: Muscovite granite from
Harney Peak, South Dakota. Journal of Geophysical Research 86, 1015–1029.
Huerta, A.D., Royden, L.H., Hodges, K.V., 1996. The interdependence of deformational and thermal processes in
mountain belts. Science 273, 637–639.
Huerta, A.D., Royden, L.H., Hodges, K.V., 1998. The thermal structure of collisional orogens as a response to
accretion, erosion, and radiogenic heat. Journal of Geophysical Research 103, 15287–15302.
Huerta, A.D., Royden, L.H., Hodges, K.V., 1999. The effects of accretion, erosion, and radiogenic heat on the meta-
morphic evolution of collisional orogens. Journal of Metamorphic Geology 17, 349–366.
Ireland, T., 1999. New tools for isotopic analysis. Science 286, 2289–2290.
Ireland, T., Gibson, G., 1998. SHRIMP monazite and zircon geochronology of high-grade metamorphism in New
Zealand. Journal of Metamorphic Geology 16, 149–168.
Irwin, W.P., 1972. Terranes of the western Paleozoic and Triassic belt in the southern Klamath Mountains, California.
Geological Survey Professional Paper 800-C1972, 103–111.
Jamieson, R.A., Beaumont, C., Fullsack, P., Lee, B., 1998. Barrovian regional metamorphism: where’s the heat? In:
Treloar, P.J., O’Brien, P.J. (Eds.), What drives Metamorphism and Metamorphic Reactions? Geological Society of
America Special Publication 138, pp. 23–51.
Jamieson, R.A., Beaumont, C., Nguyen, M.H., Lee, B., 2002. Interaction of metamorphism, deformation, and exhu-
mation in large convergent orogens. Journal of Metamorphic Geology 20, in press.
Jaupart, C., Provost, A., 1985. Heat focusing, granite genesis and inverted metamorphic gradients in continental col-
lision zones. Earth and Planetary Science Letters 73, 385–397.
M. Brown / Journal of Geodynamics 32 (2001) 115–164 157

Jessell, M., Bons, P., Evans, L., Barr, T., Stüwe, K., 2001. Elle: the numerical simulation of metamorphic and defor-
mation microstructures. Computer and Geosciences 27, 17–30.
Johnson, S.E., 1999. Porphyroblast microstructure: a review of current and future trends. American Mineralogist 84,
1711–1726.
Johnson, S.E., Vernon, R.H., 1995. Stepping stones and pitfalls in the determination of an anticlockwise P–T–t
deformation path: The low-P, high-T Cooma complex, Australia. Journal of Metamorphic Geology 13, 165–183.
Jones, K.A., 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, 551–578.
Kays, M.A., 1970. Mesozoic metamorphism, May Creek Schist Belt, Klamath Mountains, Oregon. Geological Society
of America Bulletin 81, 2743–2758.
Kesson, S.E., FitzGerald, J.D., Shelly, J.M., 1998. Mineralogy and dynamics of a pyrolite lower mantle. Nature 393,
252–255.
Kirby, S.H., Durham, W.B., Stern, L.A., 1991. Mantle phase changes and deep earthquake faulting in subducting
lithosphere. Science 252, 216–225.
Kirby, S.H. Engdahl, E.R., Denlinger, R.P., 1996a. Intermediate-depth intraslab earthquakes and arc volcanism as
physical expression of crustal and uppermost mantle metamorphism in subducting slab. In: Bebout, G.E., Scholl,
D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction Top to Bottom. Geophysical Monograph 96, pp. 195–214.
Kirby, S.H., Stein, S., Okal, E.A., Rubie, D.C., 1996b. Metastable mantle phase transformations and deep earthquakes
in subducting oceanic lithosphere. Reviews of Geophysics 34, 261–306.
Kohlstedt, D.L., Evans, B., Mackwell, S.J., 1995. Strength of the lithosphere: constraints imposed by laboratory
experiments. Journal of Geophysical Research 100, 17587–17602.
Komiya, T., Maruyama, S., Masuda, T., Nohda, S., Hayashi, M., Okamoto, K., 1999. Plate Tectonics at 3. 8-3.7 Ga:
field Evidence from the Isua Accretionary Complex, Southern West Greenland. Journal of Geology 107, 515–554.
Koons, P.O., 1995. Modeling the topographic evolution of collisional belts. Annual Review of Earth and Planetary
Sciences 23, 375–408.
Kriegsman, L.M., Hensen, B.J., 1998. Back reaction between restite and melt: Implications for geothermobarometry
and pressure-temperature paths. Geology 26, 1111–1114.
Lathrop, A.S., Blum, J.D., Chamberlain, C.P., 1996. Nd, Sr and O isotopic study of the petrogenesis of two syntec-
tonic members of the New Hampshire Plutonic Series. Contributions to Mineralogy and Petrology 124, 126–138.
LeBreton, N., Thompson, A.B., 1988. Fluid-absent (dehydration) melting of biotite in metapelites in the early stage of
crustal anatexis. Contributions to Mineralogy and Petrology 99, 226–237.
Leiss, B., Ullemeyer, K., Weber, K. et al., 2000. Recent developments and goals in texture research of geological
materials. Journal of Structural Geology 22, 1531–1540.
Leith, C.K., 1907. The metamorphic cycle. Journal of Geology 1907, 303–313.
Leloup, P.H., Ricard, Y., Battaglia, J., Lacassin, R., 1999. Shear heating in continental strike-slip shear zones: model
and field examples. Geophysical Journal International 136, 19–40.
Lister, G.S., Baldwin, S.L., 1993. Plutonism and the origin of metamorphic core complexes. Geology (Boulder) 21,
7607–7610.
Lyell, C., 1860. Manual of Elementary Geology. Appleton, New York.
Mancktelow, N.S., 1993. Tectonic overpressure in competent mafic layers and the development of isolated eclogites.
Journal of Metamorphic Geology 11, 801–802.
Mancktelow, N.S., 1995. Non-lithostatic pressure during sediment subduction and the development and exhumation of
high pressure metamorphic rocks. Journal of Geophysical Research 100, 571–583.
Mao, H.-K., Hemley, R.J., 1998. New windows on the Earth’s deep interior. Reviews in Mineralogy 37, 1–32.
Marchildon, N., Brown, M., 2001. Melt segregation in late-tectonic migmatites: an example from the Onawa contract
aureole, Maine, U.S.A.. Physics and Chemistry of the Earth 26/4- 5, 225–229.
Maruyama, S., Liou, J.G., Terabayashi, M., 1996. Blueschists and eclogites of the world and their exhumation. Inter-
national Geological Review 38, 485–594.
Maruyama, S., Parkinson, C.D., 2000. Overview of the geology, petrology and tectonic framework of the high-pres-
sure–ultrahigh-pressure metamorphic belt of the Kokchetav Massif, Kazakhstan. The Island Arc 9, 439–455.
Massonne, H.J., 1998. A new occurrence of microdiamonds in quartzofeldspathic rocks of the Saxonian Erzgebirge,
158 M. Brown / Journal of Geodynamics 32 (2001) 115–164

German, and their metamorphic evolution. In: Proceedings of the 7th International Kimberlite Conference, Cape-
town, South Africa, pp. 552–554, extended abstracts.
Mezger, K., Essene, E.J., Halliday, A.N., 1993. Closure temperatures of the Sm–Nd system in metamorphic garnets.
Earth and Planetary Science Letters 113, 397.
Mezger, K., Hanson, G.N., Bohlen, S.R., 1989. U–Pb systematics of garnet: dating the growth of garnet in the Late
Archean Pikwitonei granulite domain at Cauchon and Natawahuna Lakes, Manitoba, Canada. Contributions to
Mineralogy and Petrology 101, 136–148.
Mezger, K., Rawnsley, C.M., Bohlen, S.R., Hanson, G.N., 1991. U–Pb garnet, sphene, monazite, and rutile ages:
implications for the duration of high-grade metamorphism and cooling histories, Adirondack Mts, N.Y. Journal of
Geology 99, 415–428.
Mildren, S.D., Sandiford, M., 1995. Heat refraction and low-pressure metamorphism in the northern Flinders Ran-
gers, South Australia. Australia Journal of Earth Sciences 42, 241–247.
Miyashiro, A., 1961. Evolution of metamorphic belts. Journal of Petrology 2, 277–311.
Miyashiro, A., 1973. Metamorphism and Metamorphic Belts. Geo. Allen and Unwin, London. pp. 492.
Molnar, P., England, P.C., 1990. Temperatures, heat flux, and frictional stress near major thrust faults. Journal of
Geophysical Research 95, 4833–4856.
Montel, J., Foret, S., Veschambre, M., Nicollet, C., Provost, A., 1996. Electron microprobe dating of monazite. Che-
mical Geology 131, 37–53.
Moore, M.A., England, P.C., 2001. On the inference of denudation rates from cooling ages of minerals. Earth and
Planetary Science Letters 185, 256–284.
Moraes, R., Brown, M., Fuck, R.A., 2001. P–T path of sapphirine-quartz bearing granulites, Anápolis-Itauçu com-
plex, Central Brazil. Eleventh Annual V.N. Goldschmidt Conference, Abstracts CD.
Mortimer, N., 2000. Metamorphic discontinuities in orogenic belts: example of the garnet-biotite-albite zone in the
Otago Schist, New Zealand. International Journal of Earth Science 89, 295–306.
Nabelek, P.I., Liu, M., 1999. Leucogranites in the Black Hills of South Dakota: the consequence of shear heating
during continental collision. Geology 27, 523–526.
Nelson, K.D., 1992. Are crustal thickness variations in old mountain belts like the Appalachians a consequence of
lithospheric delamination? Geology 20, 498–502.
Nelson, K.D., Zhao, W., Brown, L.D., Kuo, J., Che, J., Liu, X., Klemperer, S.L., Makovsky, Y., Meissner, R., Mechie,
J., Kind, R., Wenzel, F., Ni, J., Nabelek, J., Leshou, C., Tan, H., Wei, W., Jones, A.G., Booker, J., Unsworth, M.,
Kidd, W.S.F., Houck, M., Alsdorf, D., Ross, A., Cogan, M., Wu, C., Sandvol, E., Edwards, M., 1996. Partially molten
middle crust beneath southern Tibet: synthesis of project INDEPTH results. Science 174, 1684–1688.
Nemcok, M., Pospisil, L., Lexa, J., Donelick, R.A., 1998. Tertiary subduction and slab break-off model of the Car-
pathian-Pannonian region. Tectonophysics 295, 307–340.
Nimis, P., Trommsdorff, V., 2001. Revised thermobarometry of Alpe Arami and other garnet peridotites from the
central Alps. Journal of Petrology 42, 103–115.
Nutman, A.P., McGregor, V.R., Friend, C.R.L., Bennett, V.C., Kinny, P.D., 1996. The Itsaq Gneiss Complex of
southern west Greenland; the world’s most extensive record of early crustal evolution (3900–3600 Ma). Precambrian
Research 78, 1–3.
Oliver, G.J.H., Chen, F., Buchwaldt, R., Hegner, E., 2000. Fast tectonometamorphism and exhumation in the type
area of the Barrovian and Buchan zones. Geology 28, 459–462.
Ota, T., Terabayashi, M., Parkinson, C.D., Masago, H., 2000. Thermobaric structure of the Kokchetav ultrahigh-
pressure-high-pressure massif deduced from a north-south transect in the Kulet and Saldat-Kol regions, northern
Kazakhstan. The Island Arc 9, 328–357.
Oxburgh, E.R., Turcotte, D.L., 1971. Origin of paired metamorphic belts and crustal dilation in island arc regions.
Journal of Geophysical Research 76, 1315–1327.
Oxburgh, E.R., Turcotte, D.L., 1974. Thermal gradients and regional metamorphism in overthrust terrains with special
reference to the eastern Alps. Schweizerische Mineralogische und Petrographische Mitteilungen 54, 641–662.
Paquin, J., Altherr, R., 2000. Modeling the exhumation path of the UHP-garnet peridotite from Alpe Arami. Eos,
Transactions of the American Geophysical Union 81, F1104.
Passchier, C.W., Trouw, R.A.J., 1996. Microtectonics. Springer-Verlag, Berlin Heidelberg.
M. Brown / Journal of Geodynamics 32 (2001) 115–164 159

Passchier, C.W., Trouw, R.A.J., Zwart, H.J., Vissers, R.L.M., 1992. Porphyroblast rotation: eppur si muove? Journal
of Metamorphic Geology 10, 283–294.
Paterson, M.S., 2001. Relating experimental and geological rheology. International Journal of Earth Science 90, 157–
167.
Patiño Douce, A.E., 1999. What do experiments tell us about the relative contributions of crust and mantle to the
origin of granitic magmas? In: Castro, A., Fernández, C., Vigneresse, J. (Eds.), Understanding Granites: Integrating
New and Classical Techniques. Geological Society of America Special Publication 168, pp. 55–75.
Patiño Douce, A.E., Harris, N., 1998. Experimental constraints on Himalayan anatexis. Journal of Petrology 39, 689–
710.
Pattison, D.R.M., Spear, F.S., Cheney, J.T., 1999. Polymetamorphic origin of muscovite+staurolite+cordier-
ite+biotite assemblages: implications for the metapelitic petrogenetic grid and for P–T paths. Journal of Meta-
morphic Geology 17, 685–703.
Pavlis, T.L., Sisson, V.B., 1995. Structural history of the Chugach metamorphic complex in the Tana River region,
eastern Alaska: a record of Eocene ridge subduction. Geological Society of America Bulletin 107, 1333–1355.
Peacock, S.M., 1996. Thermal and petrologic structure of subduction zones. In: Bebout, G.E., School, D.W., Kirby,
S.H., Platt, J.P. (Eds.), Subduction Top to Bottom. Geophysical Monograph 96, pp. 119–133.
Peacock, S.M., Rushmer, T., Thompson, A.B., 1994. Partial melting of subducting oceanic crust. Earth and Planetary
Science Letters 121, 227–244.
Peacock, S.M., Wang, K., 1999. Seismic consequences of warm versus cool subduction metamorphism; examples from
Southwest and Northeast Japan. Science 286, 937–939.
Petö, P., 1976. An experimental investigation of melting relations involving muscovite and paragonite in the silica-
saturated portion of the system K2O–Na2O–SiO2–H2O to 15 kbar total pressure. Progress in Experimental Petrol-
ogy, NERC, London 3, 41–45.
Petrini, K., Podladchikov, Yu., 2000. Lithospheric pressure-depth relationship in compressive regions of thickened
crust. Journal of Metamorphic Geology 18, 67–77.
Pfiffner, O.A., Ellis, S., Beaumont, C., 2000. Collision tectonics in the Swiss Alps: insight from geodynamic modeling.
Tectonics 19, 1065–1094.
Phillips, J., 1857. Report on cleavage and foliation in rocks and on the theoretical explanation of these phenomena.
British Association for the Advancement of Science 269, 60–61.
Pin, C., 1986. Datation U/Pb sur zircons 285 Ma du complexe gabbro-dioritique du Val Sesia—Val Mastallone et age
tardihercynien du metamorphisme granulitique de lo zone Ivrea-Verbano (Italie). Comptes Rendus de L’Academies
des Sciences 303, 827–830.
Pinet, N., Cobbold, P.R., 1992. Experimental insights into the partitioning of motion within zones of oblique subduc-
tion. Tectonophysics 206, 371–388.
Platt, J., England, P.C., 1994. Convective removal of lithosphere beneath mountain belts; thermal and mechanical
consequences. American Journal of Science 294, 307–336.
Platt, J.P., Soto, J.-I., Whitehouse, M.J., Hurford, A.J., Kelley, S.P., 1998. Thermal evolution, rate of exhumation, and
tectonic significance of metamorphic rocks from the floor of the Alboran extensional basin, western Mediterranean.
Tectonics 17, 671–689.
Powell, R., 1985. Geothermometry and geobarometry: a discussion. Journal of the Geological Society 142, 29–38.
Powell, R., Holland, T.J.B., 1988. An internally consistent thermodynamic data set with uncertainties and correlations:
3. Applications to geobarometry, worked examples and a computer program. Journal of Metamorphic Geology 6,
173–204.
Powell, R., Holland, T.J.B., 1994. Optimal geothermometry and geobarometry. American Mineralogist 79, 120–133.
Powell, R., Holland, T.J.B., Worley, B., 1998. Calculating phase diagrams involving solid solutions via non-linear
equations with examples using THERMOCALC. Journal of Metamorphic Geology 16, 575–586.
Pressley, R.A., 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, 335–366.
Prince, C.I., Kosler, J., Vance, D., Günther, D., 2000. Comparison of laser ablation ICP-MS and isotope dilution REE
analyses—implications for Sm–Nd garnet geochronology. Chemical Geology 168, 255–274.
160 M. Brown / Journal of Geodynamics 32 (2001) 115–164

Prior, D.J., Boyle, A.P., Brenker, F. et al., 1999. The application of electron backscatter diffraction and orientation
contrast imaging in the SEM to textural problems in rocks. American Mineralogist 84, 1741–1759.
Pubellier, M., Cobbold, P.R., 1996. Analog models for the transpressional docking of volcanic arcs in the western
Pacific. Tectonophysics 253, 33–52.
Pyle, J.M., Spear, F.S., 1999. Yttrium zoning in garnet: coupling of major and accessory phases during metamorphic
reactions. Geological Materials Research 1, 49–61.
Pyle, J.M., Spear, F.S., 2000. An empirical garnet (YAG)-xenotime thermometer. Contributions to Mineralogy and
Petrology 138, 51–58.
Raith, M., Karmakar, S., Brown, M., 1997. Ultrahigh-temperature metamorphism and multi-stage decompressional
evolution of sapphirine granulites from the Palni Hill Ranges, Southern India. Journal of Metamorphic Geology 15,
379–399.
Ramberg, H., 1952. Origin of igneous and metasomatic rocks. University of Chicago Press, Chicago.
Richardson, S.W., 1970. The relation between a petrogenetic grid, facies series and the geothermal gradient in meta-
morphism. Fortschritte der Mineralogie 47, 65–76.
Richardson, S.W., England, P.C., 1979. Metamorphic consequences of crustal eclogite production in overthrust oro-
genic zones. Earth and Planetary Science Letters 42, 183–190.
Ring, U., Brandon, M.T., Lister, G.S., Willett, S.D., 2000. Exhumation Processes: Normal Faulting, Ductile Flow and
Erosion. Geological Society of America Special Publication 154, pp. 2000.
Risold, A.C., Trommsdorff, V., Reusser, E., Ulmer, P., 1996. Alpe Arami and Cima di Gagnone garnet peridotites:
Observations contradicting the hypothesis of ultra deep origin. Eos, Transactions of the American Geophysical
Union 877, F761.1996.
Risold, A.C., Trommsdorff, V., Ulmer, P., Reusser, E., 1997. Genesis of FeTiO3 inclusions in garnet peridotites in the
central Alps. TERRA abstracts 9, 28.
Robin, P.-Y.F., Cruden, A.R., 1994. Strain and vorticity patterns in ideally ductile transpression zones. Journal of
Structural Geology 16, 447–467.
Roedder, E., 1984. Fluid inclusions. Reviews in Mineralogy 12, 644.
Royden, L.H., 1993a. The steady-state thermal structure of eroding orogenic belts and accretionary prisms. Journal of
Geophysical Research 98, 4487–4507.
Royden, L.H., 1993b. The tectonic expression of slab pull at continental convergent boundaries. Tectonics 12, 303–325.
Royden, L.H., 1993c. Evolution of retreating subduction boundaries formed during continental collision. Tectonics 12,
629–638.
Royden, L.H., Burchfiel, B.C., King, R.W., Wang, E., Chin, Z.L., Shen, F., Liu, Y.P., 1997. Surface deformation and
lower crustal flow in eastern Tibet. Science 276, 788–790.
Rubatto, D., Hermann, J., 2001. Exhumation as fast as subduction? Geology 29, 3–6.
Rubatto, D., Williams, I.S., Buick, I.S., 2001. Zircon and monazite response to prograde metamorphism in the Rey-
nolds Range, central Australia. Contributions to Mineralogy and Petrology 140, 458–468.
Rushmer, T., 1996. The influence of dehydration and partial melting reactions on the seismicity and deformation in
warm subducting crust. In: Bebout, G.E., School, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction Top to Bottom.
Geophysical Monograph 96, pp. 299–306.
Rutland, R.W.R., 1965. Tectonic overpressures. In: Pitcher, W.S., Flynn, G.W. (Eds.), Controls of Metamorphism.
Oliver and Boyd, pp. 119–139.
Sandiford, M., Frazier, G., Arnold, J., Foden, J., Farrow, T., 1995. Some causes and consequences of high-tempera-
ture low-pressure metamorphism in the eastern Mt. Lofty ranges, south Australia. Australian Journal of Earth
Science 42, 233–240.
Schenk, V., 1989. P–T–t path of the lower crust in the Hercynian fold belt of southern Calabria. In: Daly, J.S., Cliff,
R.A., Yardley, B.W.D. (Eds.), Evolution of mMtamorphic Belts. Geological Society of America Special Publication
43, pp. 443–342.
Schmid, R., Wood, B.J., 1976. Phase relationships in granulitic metapelites from the Ivrea–Verbano Zone (northern
Italy). Contributions to Mineralogy and Petrology 54, 255–279.
Schnetger, B., 1994. Partial melting during the evolution of the amphibolite-to-granulite-facies gneisses of the Ivrea
Zone, northern Italy. Chemical Geology 113, 71–101.
M. Brown / Journal of Geodynamics 32 (2001) 115–164 161

Searle, M., Hacker, B.R., Bilham, R., 2001. The Hindu Kush seismic zone as a paradigm for the creation of ultrahigh-
pressure diamond- and coesite-bearing continental rocks. Journal of Geology 109, 143–153.
Searle, M.P., Waters, D.J., Dransfield, M.W., Stephenson, B.J., Walker, C.B., Walker, J.D., Rex, D.C., 1999. Thermal and
mechanical models for the structural and metamorphic evolution of the Zanskar High Himalaya. In: MacNiocaill, C.,
Ryan, P.D. (Eds.) Continental Tectonics. Geological Society of America Special Publication 164, pp. 139–156.
Selverstone, J., 1994. Integrated petrologic and kinematic constraints on orogenic-scale interplay between collision,
extension, and metamorphism. Controls of metamorphism, University of Liverpool, Program and Abstracts, p. 43.
Sharpe, D., 1847. On slatey cleavage. Quarterly Journal of the Geological Society of London 3, 74–104.
Sharpe, D., 1849. On slatey cleavage. Quarterly Journal of the Geological Society of London 5, 111–115.
Shelley, D., Bossière, G., 2000. A new model for the Hercynian Orogen of Gondwanan France and Iberia. Journal of
Structural Geology 22, 757–776.
Sills, J.D., 1984. Granulite facies metamorphism in the Ivrea Zone, NW Italy. Schweizerische Mineralogische und
Petrographische Mitteilungen 64, 169–191.
Simpson, R.L., Parrish, R.R., Searle, M.P., Waters, D.J., 2000. Two episodes of monazite crystallization during
metamorphism and crustal melting in the Everest region of the Nepalese Himalaya. Geology 28, 403–406.
Sinigoi, S., Quick, J.E., Mayer, A., Budahn, J., 1996. Inference of stretching and density contrasts on the chemical
evolution of continental magmas: an example from the Ivrea–Vervano Zone. Contributions to Mineralogy and Pet-
rology 123, 238–250.
Snoeyenbos, D.R., Williams, M.L., Hanmer, S., 1995. Archean high-pressure metamorphism in the western Canadian
shield. European Journal of Mineralogy 7, 1251–1272.
Solar, G.S., Brown, M., 1999. The classic high-T–low-P metamorphism of west-central Maine, USA: is it post-tectonic
or syn-tectonic? Evidence from porphyroblast-matrix relations. Canadian Mineralogist 37, 289–311.
Solar, G.S., Brown, M., 2000. The classic high-T–low-P metamorphism of west-central Maine, USA: Is it post-tectonic
or syn-tectonic? Evidence from porphyroblast-matrix relations. Reply. Canadian Mineralogist 38.
Solar, G.S., Brown, M., 2001a. Petrogenesis of Migmatites in Maine, USA: possible source of peraluminous leuco-
granite in plutons. Journal of Petrology 42, 789–823.
Solar, G.S., Brown, M., 2001b. Deformation partitioning during transpression in response to Early Devonian oblique
convergence, Northern Appalachian orogen, USA. Journal of Structural Geology 22, 1043–1065.
Solar, G.S., Pressley, R.A., Brown, M., Tucker, R.D., 1998. Granite ascent in convergent orogenic belts: testing a
model. Geology 26, 711–714.
Sonder, L.J., England, P.C., Wernicke, B.P., Christiansen, R.L., 1987. A physical model for Cenozoic extension of
western North America. In: Coward, M.P., Dewey, J.F., Hancock, P.L. (Eds), Continental extensional tectonics.
Feological Society Special Publications 28, 187–201.
Sorby, H.C., 1853. On the origin of slatey cleavage. New Philosophy Journal (Edinburgh) 55, 137–148.
Sorby, H.C., 1856. On the theory of the origin of slatey cleavage. Philosophical Magazine 12, 127–129,135.
Sorby, H.C., 1858. On the microscopic structure of crystals, indicating the origin of minerals and rocks. Quarterly
Journal of the Geological Society of London 14, 453–500.
Soto, J.L., Platt, J.P., 1999. Petrological and structural evolution of high-grade metamorphic rocks from the floor of
the Alboran Sea Basin, Western Mediterranean. Journal of Petrology 40, 21–60.
Spear, F.S., 1989. Relative thermobarometry and metamorphic P–T paths. In: Daly, J.S., Cliff, R.A., Yardley, B.W.D.
(Eds.), Evolution of Metamorphic Belts. Geological Society of America Special Publication 43, pp. 63–81.
Spear, F., 1993. Metamorphic phase equilibria and pressure–temperature–time paths. Mineralogical Society of Amer-
ica1993, 799.
Spear, F.S., 1999. Real-time AFM diagrams on your MacIntosh. Geological Materials Research 1 (3), 1–18.
Spear, F.S., Daniel, C.G., 1998. Three-dimensional imaging of garnet porphyroblast sizes and chemical zoning:
nucleation and growth history in the garnet zone. Geological Materials Research 1 (1), 1–44.
Spear, F.S., Daniel, C.G., 2001. Diffusion control of garnet growth, Harpswell Neck, Maine, USA. Journal of Meta-
morphic Geology 19, 179–195.
Spear, F.S., Parrish, R.R., 1996. Petrology and cooling rates of the Valhalla Complex, British-Columbia, Canada.
Journal of Petrology 37, 733–765.
162 M. Brown / Journal of Geodynamics 32 (2001) 115–164

Spear, F.S., Selverstone, J., 1983. Quantitative P–T paths from zoned minerals—theory and tectonic applications.
Contributions to Mineralogy and Petrology 83, 348–357.
Spear, F.S., Selverstone, J., Hickmott, D., Crowley, P., Hodges, K.V., 1984. P–T paths from garnet zoning—a new
technique for deciphering tectonic processes in crystalline terrains. Geology 12, 87–90.
Spear, F.S., Hickmott, D.D., Selverstone, J., 1990. Metamorphic consequences of thrust emplacement, Fall Mountain,
New Hampshire. Geological Society of America Bulletin 102, 1344–1360.
Spear, F.S., Kohn, M.J., Paetzold, S., 1995. Petrology of the regional sillimanite zone, west-central New Hampshire,
USA, with implications for the development of inverted isograd. American Mineralogist 80, 361–376.
Spry, A., 1969. Metamorphic Textures. Pergamon Press, Oxford.
Stoffler, D., 1997. Minerals in the deep earth: a message from the Asteroid Belt. Science 278, 1576–1577.
Stüwe, K., Sandiford, M., 1994. Contribution of deviatoric stress to metamorphic P–T paths: an example appropriate
to low-P, high-T metamorphism. Journal of Metamorphic Geology 12, 445–454.
Suzuki, K., Adachi, M., 1991. Precambrian provenance and Silurian metamorphism of the Tsunosawa paragneiss in
the South Kitakami terrane, northeast Japan, revealed by the chemical Th–U-total Pb isochron ages of monazite,
zircon and xenotime. Journal of Geochemistry 25, 257–376.
Tenczer, V., Stüwe, K., Barr, T.D., 2001. Pressure anomalies around cylindrical objects in simple shear. Journal of
Structural Geology 23, 777–788.
Teng, L.S., Lee, C.T., Tsai, Y.B., Hsiao, L.-Y., 2000. Slab breakoff as a mechanism for flipping of subduction polarity
in Taiwan. Geology 28 (2), 155–158.
Thompson, A.B., 1981. The pressure-temperature (P,T) plane viewed by geophysicists and petrologists. TERRA Cog-
nita 1, 11–20.
Thompson, A.B., 1992. Water in the Earth’s mantle. Nature 358, 295–302.
Thompson, A.B., 1996. Fertility of crustal rocks during anatexis. Transactions of the Royal Society of Edinburgh:
Earth Science 87, 1–10.
Thompson, A.B., 1999. Some time-space relationships for crustal melting and granitic intrusion at various depths. In:
Castro, A., Fernåndez, C., Vigneresse, J.L. (Eds.), Understanding Granites: Integrating New and Classical Techni-
ques. Geological Society of America Special Publication 168, pp. 7–25.
Thompson, A.B., Connolly, J.A.D., 1995. Melting of the continental crust: some thermal and petrological constraints
on anatexis in continental collision zones and other tectonic settings. Journal of Geophysical Research 100, 15565–
15579.
Thompson, A.B., England, P.C., 1984. Pressure–temperature–time paths of regional metamorphism; II, their inference
and interpretation using mineral assemblages and metamorphic rocks. Journal of Petrology 25, 929–955.
Thompson, A.B., Tracy, R.J., 1979. Model systems for anatexis of pelitic rocks. Contributions to Mineralogy and
Petrology 70, 429–438.
Thompson, A.B., Schulmann, K., Jezek, J., 1997. Thermal evolution and exhumation in obliquely convergent (trans-
pressive). orogens. Tectonophysics 280, 171–184.
Tilley, C.E., 1924a. A preliminary survey of metamorphic zones in the Southern Highlands of Scotland. Quarterly
Journal of the Geological Society of London 81, 100–112.
Tilley, C.E., 1924b. The facies classification of metamorphic rocks. Geological Magazine 61, 167–171.
Tilley, C.E., 1925. Metamorphic zones in the southern Highlands of Scotland. Quarterly Journal of the Geological
Society of London 81, 100–112.
Touret, J.L.R., 2001. Fluids in metamorphic rocks. Lithos 55, 1–25.
Tracy, R.J., 1982. Compositional zoning and inclusions in metamorphic minerals. In: Ferry, J.M. (Ed.) Characteriza-
tion of Metamorphism Through Mineral Equilibria. Reviews in Mineralogy 10, pp. 355–397.
Trommsdorff, V., Hermann, J., Müntener, O., Pfiffner, M., Risold, A.-C., 2000. Geodynamic cycles of subcontinental
lithosphere in the Central Alps and the Arami enigma. Journal of Geodynamics 30, 77–92.
Tsuchiyama, A., Uesugi, K., Nakano, T., et al., 2001. Development of micro SR X-ray CT system at SPRING-8 for
Earth and planetary samples. Eleventh Annual V.N. Goldschmidt Conference, Abstracts CD.
Turner, F.J., 1968. Metamorphic Petrology. Mineralogical and Field Aspects. McGraw-Hill Book Co. 403 pp.
Ulmer, P., Risold, A.C., Trommsdorff, V., 1998. TiO2 solubility and mantle olivine as a function of pressure, tem-
perature, a(SiO2) and f(H2O). Eos, Transactions of the American Geophysical Union 79, S164.1998.
M. Brown / Journal of Geodynamics 32 (2001) 115–164 163

Uyeda, S., Miyashiro, A., 1974. Plate tectonics and the Japanese Islands: a synthesis. Geological Society of America
Bulletin 85, 1159–1170.
Vance, D., 1995. Rate and time controls on metamorphic processes. Geological Journal 30, 241–259.
Vance, D., Harris, N., 1999. Timing of prograde metamorphism in the Zanskar Himalayas. Geology 27, 395–398.
Vance, D., O’Nions, R.K., 1992. Prograde and retrograde thermal histories from the Central Swiss Alps. Earth and
Planetary Science Letters 114, 113–129.
Vance, D., Strachan, R.A., Jones, K.A., 1998. Extensional versus compressional settings for metamorphism: garnet
chronometry and pressure–temperature–time histories in the Moine Supergroup, northwest Scotland. Geology 26,
927–930.
Vanderhaeghe, O., Teyssier, C., in press. Partial melting and flow of orogens. Tectonophysics.
Vanderhaeghe, O., Burg, J.-P., Teyssier, C., 1999. Exhumation of migmatites in two collapsed orogens: Canadian Cordil-
lera and French Variscides. In: Ring, U., Brandon, M.T., Lister, G.S., Willet, S.D. (Eds.), Exhumation Processes: Nor-
mal Faulting, Ductile Flow and Erosion. Geological Society of America Special Publication 154, pp. 181–204.
Van der Voo, R., Spakman, W., Bijwaard, H., 1999. Tethyan subducted slabs under India. Earth and Planetary Science
Letters 171, 7–20.
Van Hise, 1904. A Treatise on Metamorphism. Monograph No. 47, US Geological Survey.
van Roermund, H.L.M., Drury, M.R., 1998. Ultra-high pressure (P>6 GPa) garnet peridotites in western Norway:
Exhumation of mantle rocks from >185 km depth. TERRA Nova 10, 295–301.
van Roermund, H.L.M., Drury, M.R., Barnhoorn, A., DeRonde, A.A., 2000. Super-silicic garnet microstructures from
an orogenic garnet peridotite, evidence for an ultra-deep (>6 GPa) origin. Journal of Metamorphic Geology 18,
135–147.
Vauchez, A., Tommasi, A., Barruol, G., 1998. Rheological heterogeneity, mechanical anisotropy and deformation of
the continental lithosphere. Tectonophysics 296, 61–86.
Vavra, G., Schaltegger, U., 1999. Post-granulite facies monazite growth and rejuvenation during Permian to lower
Jurassic thermal and fluid events in the Ivrea Zone, Southern Alps. Contributions to Mineralogy and Petrology 134,
405–414.
Vavra, G., Schmid, R., Gebauer, D., 1999. Internal morphology, habit and U–Th–Pb microanalysis of amphibolite-to-
granulite facies zircons: Geochronology of the Ivrea Zone, Southern Alps. Contributions to Mineralogy and Petrol-
ogy 134, 380–404.
Vernon, R.H., 1976. Metamorphic Process. George Allen & Unwin Limited, London.
Vielzeuf, D., Montel, J.M., 1994. Partial melting of metagreywackes. Part I. Fluid-absent experiments and phase rela-
tionships. Contributions to Mineralogy and Petrology 117, 375–393.
Vielzeuf, D., Clemens, J.D., Pin, C., Moinet, E., 1990. Granites, granulites and crustal differentiation. In: Vielzeuf, D.,
Vidal, Ph. (Eds.), Granulites and Crustal Evolution. Klewer Academic Publishers, pp. 59–85.
Viskupic, K., Hodges, K.V., 2001. Monazite–xenotime thermochronometry: methodology and an example from the
Nepalese Himalaya. Contributions to Mineralogy and Petrology 141, 233–247.
Vissers, R.L.M., Platt, J.P., Vanderwal, D., 1995. Late orogenic extension of the Betic Cordillera and the Aalboran
domain—a lithospheric view. Tectonics 14, 786–803.
von Blanckenburg, F., Davies, J.H., 1995. Slab breakoff: a model for syncollisional magmatism and tectonics in the
Alps. Tectonics 14, 120–131.
Wallis, S.R., Aoya, M., 2000. A re-evaluation of eclogite facies metamorphism in SW Japan: proposal for an eclogite
nappe. Journal of Metamorphic Geology 18, 653–664.
Wallis, S.R., Aoya, M., Uehara, S., 2000. Numerical modelling of clockwise P–T paths in subduction-related meta-
morphic belts and implications for exhumation rates. Eos, Trans. Am. Geophys. Union 81, F1104.
Waschbusch, P., Batt, G., Beaumont, C., 1998. Subduction zone retreat and recent tectonics of the South Island of
New Zealand. Tectonics 17, 267–284.
Watt, G.R., Oliver, N.H.S., Griffin, B.J., 2000. Evidence for reaction-induced microfracturing in granulite facies mig-
matites. Geology 28, 327–330.
Wenzel, F., Achauer, U., Enescu, D., Kissling, E., Russo, R., Mocanu, V., Musacchio, G., 1998. Detailed look at final
stage of plate break-off is target of study in Romania. Eos, Transactions of the American Geophysical Union 79,
589–594.
164 M. Brown / Journal of Geodynamics 32 (2001) 115–164

Wheeler, J., Prior, D.J., Jiang, Z., Spiess, R., Trimby, P.W., 2001. The petrological significance of misorientations
between grains. Contributions to Mineralogy and Petrology 141, 109–124.
Whitney, D.L., Dilek, Y., 1998. Characterization and interpretation of P–T paths with multiple thermal peaks. In:
Treloar, P.J., O’Brien, P.J. (Eds.), What Drives Metamorphism and Metamorphic Reactions? Geological Society of
America Special Publication 138, pp. 53–60.
Whitney, D.L., Miller, R.B., Paterson, S.R., 1999. P–T–t evidence for mechanisms of vertical tectonic motion in a
contractional orogen: north-western U. S., Canadian Cordillera. Journal of Metamorphic Geology 17, 75–90.
Whittington, A., Harris, N., Baker, J., 1998. Low-pressure crustal anatexis: the significance of spinel and cordierite
from metapelitic assemblages at Nanga Parbat, northern Pakistan. In: Treloar, P.J., O’Brien, P.J. (Eds.), What
Drives Metamorphism and Metamorphic Reactions. Geological Society of America Special Publication 138,–
198.1998, 183.
Whittington, A., Harris, N., Butler, R., 1999. Contrast in anatectic styles at Nanga Parbat, northern Pakistan. Geo-
logical Society of America Special Paper 328, 129–144.
Williams, I.S., 2001. Response of detrital zircon and monazite, and their U–Pb isotopic systems, to regional meta-
morphism and host-rock partial melting, Cooma Complex, southeastern Australia. Australian Journal of Earth
Science 48, 557–558.
Williams, I.S., Buick, I.S., Cartwright, I., 1996. An extended episode of early Mesoproterozoic metamorphic fluid in
the Reynolds Range, central Australia. Journal of Metamorphic Geology 14, 29–47.
Williams, M.L., 1994. Sigmoidal inclusion trains, punctuated fabric development, and interactions between meta-
morphism and deformation. Journal of Metamorphic Geology 12, 1–21.
Williams, M.L., Karlstrom, K.E., 1996. Looping P–T paths and high-T, low-P middle crustal metamorphism: proter-
ozoic evolution of the southwestern United States. Geology 24, 1119–1122.
Williams, M.L., Jercinovic, M.J., Terry, M.P., 1999. Age mapping and dating of monazite on the electron microprobe:
deconvoluting multistage tectonic histories. Geology 27, 11–1023–1026.
Williams, M.L., Scheltema, K.E., Jercinovic, M.J., 2001. High-resolution compositional mapping of matrix phases:
implications for mass transfer during crenulation cleavage development in the Moretown Formation, western Mas-
sachusetts. Journal of Structural Geology 23, 000.
Worley, B., Powell, R., 2000. High-precision relative thermobarometry: theory and a worked example. Journal of
Metamorphic Geology 18, 91–101.
Yang, P., Rivers, T., 2001. Chromium and manganese zoning in pelitic garnet and kyanite: Spiral, overprint and
oscillatory(?) zoning patterns and the role of growth rate. Journal of Metamorphic Geology 19, 455–474.
Yang, J., Godard, G., Kienast, J.-R., Lu, Y., Sun, J., 1993. Ultrahigh-pressure (60 kbar) magnesite-bearing garnet
peridotites from northeastern Jingsu. China. Journal of Geology 101, 541–554.
Zeck, H.P., 1996. Betic-Rif orogeny: subduction of Mesozoic Tethys under E-ward drifting Iberia, slab detachment
shortly before 22 Ma, and subsequent uplift and extensional tectonic. Tectonophysics 254, 1–16.
Zeitler, P.K., Chamberlain, C.P., Smith, H.A., 1993. Synchronous anatexis, metamorphism, and rapid denudation at
Nanga Parbat, Pakistan Himalaya. Geology 21, 347–350.
Zeitler, P.K., Meltzer, A.S., Koons, P.O., Crew, D., Hallet, B., Chamberlain, C.P., Kidd, W.S.F., Park, S.K., Seeber,
L., Bishop, M., Shroder, J., 2001. Erosion, Himalayan geodynamics, and the geomorphology of metamorphism.
GSA Today 11, 4–9.
Zhu, X.K., O’Nions, R.K., 1999a. Zonation of monazite in metamorphic rocks and its implications for high-tem-
perature thermochronology: a case study from the Lewisian terrane. Earth and Planetary Science Letters 171, 202–
220.
Zhu, X.K., O’Nions, R.K., 1999b. Monazite chemical composition: some implications from monazite geochronology.
Contributions to Mineralogy and Petrology 137, 351–363.
Zhu, X.K., O’Nions, R.K., Bellshaw, N.S., Gibb, A.J., 1997. A Lewisian crustal history from in situ SIMS mineral
chronometry and related metamorphic textures. Chemical Geology 136, 205–218.
Zwart, H.J., 1960. The chronological succession of folding and metamorphism in the central Pyrenees. Geologische
Rundschau 50, 203–218.
Zwart, H.J., 1962. On the determination of polymetamorphic mineral associations, and its application to the Bosost
area (central Pyrenees). Geologische Rundschau 52, 38–65.

You might also like