Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Gondwana Research 18 (2010) 106–137

Contents lists available at ScienceDirect

Gondwana Research
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / g r

A new perspective on metamorphism and metamorphic belts☆


Shigenori Maruyama a,⁎, H. Masago b, I. Katayama c, Y. Iwase d, M. Toriumi e, S. Omori a, K., Aoki a
a
Department of Earth and Planetary Sciences, Tokyo Institute of Technology, Japan
b
Center for Deep Earth Exploration, Japan Agency for Marine-Earth Science and Technology, Japan
c
Department of Earth and Planetary Systems Science, Graduate School of Science, Hiroshima University, Japan
d
Department of Earth and Ocean Sciences, National Defense Academy of Japan, Japan
e
Department of Earth and Planetary Science, Graduate School of Frontier Sciences, The University of Tokyo, Japan

a r t i c l e i n f o a b s t r a c t

Article history: The discovery of ultrahigh-pressure rocks from collision-type orogenic belts has revolutionized the classic
Received 8 February 2010 interpretation of (1) progressive and retrogressive metamorphism recorded on surface exposures of regional
Received in revised form 22 March 2010 metamorphic belts, (2) geochronology of the various stages of metamorphism, (3) origin of metamorphic
Accepted 23 March 2010
textures, (4) P–T–t path, (5) metamorphic facies series, (6) exhumation model, and (7) role of fluids during
Available online 30 March 2010
regional metamorphism. Based mainly on our recent studies of the Kokchetav, Dabie Shan, Indonesia,
Franciscan and Sanbagawa belts, we suggest the following revolutionary paradigm shifts.
Keywords:
Metamorphism
The so-called mineral isograds defined on the maps of regional metamorphic belts were a misunderstanding
Facies series of the progressive dehydration reaction during subduction because extensive late-stage hydration has
Isograd mostly obliterated the progressive minerals in pelitic–psammitic and metabasic rocks. Progressively zoned
Barrovian hydration garnet has survived as the sole progressive mineral that was unstable with the majority of matrix-forming
Exhumation minerals. The classic Barrovian isograds should therefore be carefully re-examined. The well-documented
Tectonics SHRIMP chronology of spot-dating zoned zircons with index minerals from low-P in the core, through HP–
UHP in the mantle to low-P on the rim clearly shows that the slow exhumation speed of 23–40 My from
mantle depth to mid-crustal level was followed by mountain building with doming at latest stage. Extensive
hydration of the UHP–HP unit occurred due to fluid infiltration underneath, when the UHP–HP unit intruded
the low-grade to unmetamorphosed unit at a mid-crustal level. Most deformation textures such as mineral
lineations, porphyroblasts, pull-apart, or boudinaged amphiboles, formed during extensive hydration at the
late stage do not constrain the progressive stress regime. The P–T–time path determined by thermo-
barometry using mineral inclusions in garnet and forward modeling of garnet zoning, independent of the
matrix minerals, indicates an anticlockwise trend in the P–T space, and follows an independent P–T change in
the metamorphic facies series. This is consistent with the numerically calculated geotherm along the
Wadati–Benioff plane. Collision-type orogenic belts have long been regarded as being characterized by the
intermediate-pressure type metamorphic facies series. The kyanite–sillimanite is an apparent type facies
series formed by the late-stage extensive hydration. In contrast, the original high-P to ultrahigh-P type facies
series with an anticlockwise kink-point at around 10 kb is a progressive type. A collision-type regional
metamorphic belt crops out as a very thin unit sandwiched between overlying and underlying low-P or
weakly metamorphosed units. The metamorphic belt has an aspect ratio (thickness vs width) of 1:100, and it
extends for several hundreds to a thousand km. It resembles a thin mylonitic intrusion from the mantle
extending from 100 to 200 km depth into the crustal rock unit. The underlying unit is thermally
metamorphosed in the andalusite–sillimanite type facies series. The major reason for the misunderstanding
of the progressive metamorphism in collision-type orogenic belts is the underestimation of the role of fluids
derived from the underlying low-grade metamorphic unit, when juxtaposed at a mid-crustal level. The
circulation of fluids along the plate boundaries is more important than a P–T change.
© 2010 International Association for Gondwana Research. Published by Elsevier B.V. All rights reserved.

1. Introduction
☆ Adapted and modified from the article by Maruyama, S., Masago, H., Katayama, I.,
Iwase, Y., Toriumi, M., 2004. A revolutionary new interpretation of a regional Our understanding of regional metamorphism is pivoted on infor-
metamorphism, its exhumation and consequent mountain building. Journal of mation from experimental petrology combined with thermodynam-
Geography, 113, 727–768 (in Japanese).
⁎ Corresponding author.
ics, structural and petrological studies of regional metamorphic belts
E-mail addresses: smaruyam@geo.titech.ac.jp (S. Maruyama), from microscopic to map scale, and the paradigm established by plate
omori@geo.titech.ac.jp (S. Omori). tectonics (e.g. recent works by Brown 2010-this issue, Omori et al.,

1342-937X/$ – see front matter © 2010 International Association for Gondwana Research. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.gr.2010.03.007
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 107

2009; Santosh and Kusky, 2010, Zhang et al., 2009, among others). The under the surface, at temperature and pressure conditions prevailing
empirical and conceptual developments during the period 1965 to in the wet region. However, exposed regional metamorphic belts such
1980 led to various advances in this field and laid the basis for a as those of Alps, Himalaya and other areas show the presence of high
modern approach (Miyashiro 1998). Thermodynamic database and temperature minerals such as plagioclase and pyroxene on a regional
experimental data on subsolidus phase relations became available in scale. The common presence of anhydrous high-T minerals in region-
the late 70s (Helgeson et al., 1978) which was further improved by ally exposed metamorphic belts suggests disequilibrium under
later workers. Analyses of pressure–temperature–time (P–T–t) paths surface environment (Fig. 1). A possible explanation for this phe-
became possible through the application of electron microprobe in the nomenon is the absence of fluid during the exhumation from deep
study of rock-forming minerals in regional metamorphic belts. These, crust to the surface.
combined with radiogenic isotope geochronology and thermal history Mineral isograds have been traditionally constructed to explain
by numerical simulation, led to the proposal of tectonic models for the the progressive nature of regional metamorphic belts (see example
exhumation of regional metamorphic belts such as the benchmark shown in Fig. 2). From low- to high-grade of metamorphism, the
work of England and Thompson (1984). It was an epoch making change in pelitic mineral assemblages define the isograds, such as for
success which demonstrated the potential of metamorphic petrology example the chlorite, biotite, garnet, staurolite, sillimanite and
in modeling mountain building processes along consuming plate kyanite isograds in the collision-type Scottish Highland (Barrow,
boundaries by combining information from geology and petrology 1893; Fettes, 1979) (Fig. 2), and chlorite, garnet, biotite and oligoclase
(e.g., Spear et al., 1990). Thus, systematization of metamorphic in the Pacific-type Sanbagawa belt, Japan (Higashino, 1990). These
petrology under Earth Science was generally believed to have been isograds can be explained by P and T increase, which promote dehy-
successfully achieved. dration reactions at given chemical composition. Interpretation was
However, it was at this period that ultrahigh-pressure (UHP) that excess fluid was present during the P–T increase leading to the
metamorphic rocks were discovered from some orogenic belts. This completion of chemical equilibria promoting the progressive meta-
discovery was a trigger to reconsider the existing concepts of regional morphism (Fig. 1a,b). However, after the thermal peak (P–T
metamorphism from the narrow realm of metamorphic petrology to a maximum) the recrystallized metamorphic belts return to the surface.
much larger canvas, particularly incorporating processes in consum- If there is no fluid, the mineral assemblages formed at specific P–T
ing plate boundaries as well as the information from geophysics. conditions should remain frozen and will not undergo any change
Although the first finding of coesite-bearing of UHP rocks from Alps (Fig. 1c). The presence of high-grade mineral assemblages in the
(Chopin, 1984) was a major breakthrough, it soon became apparent exposed metamorphic belts has been taken to indicate the absence of
that this is not an isolate case, and that UHP rocks occur in many fluids after the thermal peak, a speculation that has been well
regional metamorphic belts formed by continent–continent collision demonstrated in petrology and supported by empirical considera-
as revealed by reports in the last 20 years (e.g. Zhang et al., 2009 and tions. Therefore, even though the logic is somewhat circular, most
references therein; Masago et al., 2009). Traditionally, and empiri- petrologists endorsed this concept.
cally, intermediate-pressure type metamorphism was considered to
be a characteristic for the collision-type orogeny, and was regarded to 2.2. Meaning of mineral isograd
be a progressive metamorphism. However, this was proved wrong by
subsequent studies as progressive metamorphism ranged from low-T As shown typically in the case of Scottish and Appalachian meta-
high-P type facies series covering blueschist through eclogite to morphic belts, the appearance of anhydrous minerals seems to be
coesite and or diamond facies metamorphism. Recently, similar common with increasing metamorphic grade. Therefore, in a given
results were also obtained from the Pacific-type regional metamor- rock system, such as pelitic or mafic, the mineral assemblages indicate
phic belts (Aoki, 2009). Therefore, metamorphic petrology systema- dehydration reactions at the high-T side to expelling fluids at the
tized through the generally accepted models of England and higher-grade side. This process is envisioned as progressive meta-
Thompson (1984) and Spear et al. (1990) faced new challenges. The morphism and records a P–T increase during regional metamorphism.
new discoveries warranted a restructuring of the basic framework of Therefore, the mineral isograds preserve the P–T structure during the
metamorphic petrology in terms of the geodynamics and tectonics of maximum-P–T conditions.
regional metamorphic belts. A paradigm shift in this field should be However, if we consider the texture of metamorphic rocks, the
able to address: (1) progressive regional metamorphism; (2) margin of amphibole or garnet is often replaced by low-temperature
retrogressive and probably hydration metamorphism; (3) geochro- minerals, which indicate minor hydration and recrystallization
nology; (4) metamorphic textures and structures and the estimation following the peak metamorphic conditions. Thus, hydration recrys-
of stress; (5) P–T–t path; (6) metamorphic facies series; (7) tallization was considered to be minor during the exhumation of
exhumation tectonics and (8) the behavior of metamorphic fluid. regional metamorphic belts. It was thus considered that high-grade
Most of these themes are now facing several new challenges and metamorphic rocks could be transported to the surface without any
drastic revision of previous concepts based on technological and significant hydration–recrystallization.
conceptual advancements. In this paper we synthesize the traditional As will be discussed in detail later, this notion was wrong. The late-
ideas and concepts of metamorphic petrology and evaluate the recent stage retrogression was extensive to obliterate almost all progressive
observations, finally leading to the proposal of some new concepts. mineralogy at high grades, hence we need to reconsider the basic
concept of metamorphism and metamorphic belts.
2. A brief historical review of regional metamorphism
2.3. Exhumation model
In this section we provide a brief overview of the existing major
concepts including salient assumptions on regional metamorphism. Almost all regional metamorphic belts were considered to be
generally stable within the stability field of plagioclase. Plagioclase
2.1. Metamorphic rocks exposed on the surface indicate absence of fluids occupies more than 50% modal content among the mineral assem-
blages in psammitic to pelitic sedimentary rocks and metabasites. The
If excess fluid enriched in water is always available during meta- upper stability limit of plagioclase in both metabasite and metasedi-
morphism, a chemical equilibrium must prevail, and hence high- mentary rocks is up to 12 kbar (Miyashiro, 1973, 1994a,b). Therefore,
grade metamorphic rocks cannot be present on the surface of the transportation depth of regional metamorphic belts was considered to
Earth. Zeolite facies minerals such as clays are the stable assemblages have never exceeded the bottom of continental crust (ca. 35 km).
108 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

Fig. 1. Schematic diagram of regional metamorphism in subduction zones. (a) The formation of protolith. Accretionary complex formed at trench and gradually decreased its water
contents during subduction (shown with change in density of dots in the metamorphic belts in the diagram from a and b). As most prograde metamorphic reactions are dehydration
with excess aqueous fluid, chemical equilibration is achieved at prograde to peak stages. (b) On the contrary, most retrograde reactions during the exhumation stage are hydration
reactions, which cannot proceed without sufficient water being supplied from outside the system. For this reason, it has long been believed that metamorphic rocks exposed on the
surface preserve near peak stage of the metamorphic mineral assemblages (c). However, recent studies have clearly revealed that metamorphic belts have undergone
recrystallization with a large amount of water infiltrating at the mid-crustal level (d).

Based on this assumption, it was generally held that continental crust grade (300 °C) to high grade (700 °C), suggesting that the belt
cannot subduct into mantle and would only underplate leading to exposes the subducted Indian crust from surface to the lower
crustal duplication as in the case of the model proposed for Himalayas structural level. The protoliths of Himalayan metamorphic rocks
(England and Thompson, 1984). Advances in experimental petrology still preserve some of the lithostratigraphy derived from the
led to the compilation of an exhaustive database on the thermody- Gondwana continental assembly, and the paleontological records
namic properties of most of the rock-forming minerals (e.g. Helgeson are also consistent with this inference (Gansser, 1964; Lefort, 1975).
et al., 1978, Holland and Powell, 1998). Using this database, Therefore, the Asian continental crust, which constitutes the upper
metamorphic P–T paths have been traced by the continuous change part of duplicated continents, seems to have been significantly
in the composition of minerals such as garnet that are assumed to be eroded, the debris from which now constitute the voluminous Bengal
in equilibrium with the matrix minerals. If chemical equilibria were fan deposits within the Indian Ocean. Judging from Fig. 3b, the
perfectly maintained, then the radiometric ages derived from key surface regions which recrystallized below 300 °C were not heated
minerals provide valuable insights into the exhumation mechanism of up above by the high-temperature Asian lower crust, and the
regional metamorphic belts. exhumation must have been extremely fast and short that occurred
One of the best examples for widely cited exhumation model is the within a few million years.
Himalayan metamorphic belt proposed by England and Thompson We now consider the P–T architecture of the subducted Indian crust.
(1984). We will now consider the tectonic history of this belt. The The Indian crust was transported to a depth of ca. 45 km within 1 million
estimated horizontal movement is about 70 to 80 km/My from year. The fast rate of subduction means that during the initial stage of
computations based on the northward migration of Indian continent subduction, only the pressure must have increased with temperature
at the rate of 7 to 8 cm/year (England and Thompson, 1984). If the remaining nearly constant (Fig. 3c). At the depth of 45 km, the
subduction angle is 30°, the leading edge of the continent would reach subducted crust stays for a considerable time when P must be constant
Moho depths within 1 million year (Fig. 3a). Therefore, thermal whereas T must gradually increase. This is due to the fact that heat
calculation justifies the instantaneous overlapping of two continental transfer occurs only by conduction which is a very slow process as
crusts (Fig. 3b). In this calculation, it was assumed that the temperature compared to the speed of the plate movement (England and Thompson,
at the surface of the subducted continental crust was zero degree 1984). It should be noted that due to isostatic rebound, metamorphic
centigrade. The surface of the subducted crust is heated up through time belts slowly uplift leading to temperature increase vis-à-vis pressure
by the overlying lower crust of Asia, to an estimated temperature of decrease to exhume the metamorphic belts to the surface by buoyancy.
500 °C. The geotherm on the top surface of the Indian continent is thus Simultaneously, extensive erosion occurs on the surface (Fig. 3c). As
gradually heated up to attain a steady state geothermal gradient which expected in such a model, a clockwise rotation of the P–T path occurs
is a function of time (Fig. 3b). In this case, a time span of only 5 million which matches well with the P–T time path deduced from chemically
years is sufficient to attain such a state. zoned garnets (Spear, 1993). This model has been supported by several
Because the subducted continental crust is expected to have large petrological studies and Ganguly et al. (1998) estimated the average
buoyancy, this crust must uplift right after the duplication of the two exhumation speed of metamorphic belts in the Himalaya based on the
blocks following isostatic requirements (Fig. 3a, right upper panel). diffusion velocity of trace elements from zoned garnet (see also Tirone
The present day surface exposure of the Himalayan regional and Ganguly, 2010-this issue for recent models on garnet diffusion). The
metamorphic belts exhibit a continuous sequence from very low results were also consistent with an extremely rapid exhumation.
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137
Fig. 2. (a) Map view of metamorphic zoning and mineral isograd on the Scottish Highland. (b) The appearance and disappearance of index minerals in pelitic rocks with increasing metamorphic grade (Fettes, 1979). Intermediate-pressure
type metamorphic facies series in Scottish Highland has been called as Barrovian and regarded to be the world standard of progressive metamorphism in the regional metamorphic belts. (c) AFM ternary expression of the progressive
disappearance of staurolite. Staurolite-bearing assemblages change to an assemblage of sillimanite + garnet + biotite (right) with the disappearance of staurolite due to temperature increase (center). (d) The appearance and disappearance of
index minerals in a model pelitic rock with increasing metamorphic grade along prograde P–T path of the Kokchetav UHP metamorphic rock (after Masago et al., 2010-this issue). Compare (b) and (d) to evaluate the late stage of hydration to
mask the stability field of progressive HP minerals at highest grades.

109
110 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

Fig. 3. Classic model for continent collision orogeny of England and Thompson (1984). (a) Schematic cross section of the collisional orogen showing crustal duplication by collision
(left) and subsequent isostatic uplift and erosion (right). (b) Thermal profiles along the depth of the collision zone at various stages after collision. The discontinuous thermal profile
formed by double thickened crust (0 Ma) recovers to a steady state geotherm with time. (c) Pressure–temperature paths for the hanging wall and footwall of the orogen. Both show
clockwise P–T time path through the orogeny.

3. A revolutionary change in the concepts of metamorphism that UHP metamorphic rocks are remarkably exotic in terms of P–T
condition, many people still consider the England and Thompson (1984)
3.1. Discovery of UHP metamorphic rocks model for Himalayan metamorphic belts as an acceptable model.

The attractive models proposed for the formation and exhumation of 3.2. Mode of occurrence: coherent, allochthonous or autochthonous?
the Himalayan regional metamorphic belts provide a typical case of the
systematization of metamorphic petrology and illustrate the consisten- Following the discovery of UHP rocks, one of the main topics of
cy between theoretically predicted model and the field and petrologic debate was the mode of occurrence of the metamorphic belts. The
data. Hence, metamorphic petrologists and geologists widely applied pressure conditions of UHP rocks range from 30 to 70 kbar, which is
the Himalayan model for other regional metamorphic belts formed by clearly outside the stability field of plagioclase, the stability of which
continent–continent collision employing a similar petrological, tectonic denotes pressures of only less than 12 kbar. This led to a contradiction in
geochronological method. This template study thus became a common the conceptual framework of regional metamorphism. One possibility is
currency over the world and similar models for different regional that UHP rocks were caught up within low-pressure regional meta-
metamorphic belts in space and time were proposed. morphic belts as tectonic blocks. The second view is that it is similar to
However, the discovery of UHP metamorphic rocks from Alps gravel within conglomerates or olistostrome (Wang et al., 1992). The
(Chopin, 1984) marked a historic change in our concepts on regional third possibility is that the regional metamorphic belt itself was
metamorphism and cast the first shadow of doubt on the Himalayan subjected to high-pressure to ultrahigh-pressure metamorphism
models. Chopin's (1984) report of coesite-bearing white schist from the followed by hydration–recrystallization along intermediate-P series
highest grade of the Penninic nappe in the Italian Alps was followed by below 12 kbar (see a summary by Wang et al., 1992). Among these three
further similar findings such as the discovery of coesite from Norwegian interpretations, most people believed the former two cases, possibly
Caledonides (Smith, 1984). Thereafter, ultrahigh-pressure metamor- because UHP rocks were derived from eclogites which were thought to
phic rocks have been reported from a number of collisional orogenic be derived from ophiolites. Another reason for this contention was due
belts over the world (see Zhang et al., 2009 and references therein), and to the fact that the major portion of regional metamorphic belts was
new reports continue even today adding to the number of UHP localities, derived from psammitic to pelitic or orthogneiss which do not exhibit
which already exceed 25 (Fig. 4). Although there is a general consensus the HP–UHP metamorphic conditions.
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 111

Fig. 4. Global distribution of UHP metamorphic rocks (partly modified after Parkinson et al., 2002). Numbers below the localities represent ages in Ma.

However, continued studies on UHP metamorphism revealed that et al., 2010; Santosh and Kusky, 2010), followed by exhumation up to
the possibility (3) mentioned above is the most viable scenario. A mid-crustal level when hydration obliterated part of the progressive
detailed case study was reported by Katayama et al. (2000) who metamorphic history through replacement by intermediate facies
analyzed zircon inclusions separated from pelitic and psammitic rocks series recrystallization.
from diamond-grade Kokchetav UHP regional metamorphic belt using The prograde and retrograde metamorphic history and the time
laser Raman spectrometry. More than 20 minerals were identified gap between these two episodes were clearly identified through spot
within the zircons. Metamorphic mineral zoning was established analysis of zircons from UHP rocks using SHRIMP method pioneered
based on the mineral inclusions in zircons, and the results clearly by Katayama et al. (2001). The history recorded from the core and
demonstrated that the psammitic and pelitic metamorphic units of margin of single zircon grains provided new insights into the
Kokchetav must have undergone regional metamorphism under HP– markedly different P–T conditions; for example, a diamond-bearing
UHP conditions. These metasedimentary rocks comprise more than core against graphite and plagioclase bearing rim. Thus, two
90% of the belt, and therefore the metamorphic belt itself must have independent metamorphic episodes were distinguished with ca. 20
been subjected first to HP–UHP facies series and was later affected by to 30 million years difference between the two events. This has been
intermediate-pressure metamorphism within Moho depth (Katayama confirmed not only in Kokchetav, but also in other UHP terranes
et al., 2001; Masago et al., 2010-this issue). This study, and similar including the Dabie–Sulu and Himalayas (Wang et al., 1992; Tabata
other works were a turning point on the concept of collision-type et al., 1998; Liu et al., 2002; Kaneko et al., 2003; Sachan et al., 2004;
regional metamorphism. Zhang et al., 2009). The time gap reaching 20 to 30 million years
indicate extremely slow exhumation rate.
3.3. Concept of regional metamorphism faces serious challenge The Kokchetav and Dabie–Sulu examples, among other terranes,
finally led to the understanding that HP–UHP metamorphism was not
Following the Kokchetav research, similar results were obtained an exotic phenomenon and represents part of a progressive metamor-
from the Dabie–Sulu Triassic collisional orogenic belts in South China phic cycle association with the deep subduction of continental margins
(Liu et al., 2002; see Zhang et al., 2009 for a recent review). Moreover, followed by late-stage hydration–recrystallization during exhumation.
coesite was also discovered in granitic gneiss from the western part of The information so far available indicates that UHP metamorphic rocks
Himalayan belt, the Pakistan and Indian Himalayas (Kaneko et al., are generally restricted to Phanerozoic orogenic belts, and no typical
2003). This led to a drastic change in concept and it became apparent Archean and Proterozoic examples have been reported, except minor
that most of the collisional metamorphic belts must have suffered first exceptions for latest Proterozoic examples from Brazil and Africa (Caby,
HP–UHP metamorphism during the prograde stage (e.g. Nishimiya 1994; Campos Neto, 2000; Parkinson et al., 2002).
112 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

3.4. Impact on the Pacific-type orogenic belts

The new findings from collision-type regional metamorphic belts


such as those of Kokchetav were soon extended to the Pacific-type
orogenic belts such as the Sambagawa metamorphic belt in Japan (Ota
et al., 2004; Aoki et al., 2007, 2008, 2009). At the highest metamorphic
grade for Sambagawa belt in the Shikoku island of SW Japan, an
amphibolite mass of approximately 10 × 4 km2 is exposed with a relict
mineral assemblage corresponding to eclogite facies, the pressure
conditions of which are markedly higher than the already described
surrounding amphibolite facies rocks (below 12 kbar). However,
these eclogites were considered as tectonic blocks occurring along
fault zones (Takasu 1989). More recently, eclogite facies mineral
assemblages were widely mapped over the entire mass (Ota et al.,
2004). Moreover, small eclogite bodies were also reported from the
surrounding Sambagawa metamorphic rocks (Aoya and Wallis, 1999).
These findings radically revised the traditional concepts of progres-
sive Sambagawa metamorphism along HP intermediate facies series
below 12 kbar, which represents only a later stage hydration–
recrystallization after exhumation and emplacement at mid-crustal
level. Thus, the high-pressure progressive metamorphism might have
occurred at an earlier stage.
Fig. 5. Metamorphic facies series and the representative metamorphic belts (modified
4. New paradigm after Miyashiro, 1965). Three metamorphic facies series (high-P/T, medium-P/T and
low-P/T) and two intermediate series were proposed. Medium- and high-P/T series
were regarded to characterize collision- and Pacific-type metamorphism respectively.
The framework of concepts on the dynamics of regional meta- Zeo: zeolite facies; PA: Pumpellyite–actinolite facies; PrA: Prehnite–actinolite facies;
morphism underwent rapid changes in the recent years and in this GS: Greenschist facies; EA: Epidote–amphibolite facies; AM: Amphibolite facies; BS:
section we summarize some of the crucial aspects. Blueschist facies; EC: Eclogite facies; HGR: High-P granulite facies; LGR: Low-pressure
granulite facies; Px-HR: Pyroxene-hornfels facies.

4.1. Intermediate facies series


Kokchetav and Dabie–Sulu. However, the volume of metabasites is small
One of the aspects to be evaluated is whether or not the kyanite– and generally less than 10% of the total volume of rocks exposed in such
sillimanite type characterizes metamorphism in collisional orogenic belts. The remaining 90% is composed of para- or orthogneiss. The
belts. Miyashiro (1961) defined three independent metamorphic matrix minerals of metasedimentary rocks have recrystallized under
facies series and two intermediate groups between the three. The type mid-crustal level. The rare metabasites, even though occurring as thin
localities listed by Miyashiro (1961) are the Franciscan, Sambagawa, units, locally preserve jadeite + coesite+ garnet, as in the case where
Scottish Highland, New Zealand and the Ryoke belt (Miyashiro, 1973). they are enclosed within metacarbonates (Ogasawara et al., 2002). In
Based on the description of regional metamorphic belts, it was some cases, metamorphosed sandstones surrounded by metacarbo-
pointed out that the intermediate-pressure type metamorphic belts nates preserve similar UHP assemblages. The stable HP and UHP
characterize the collision type. On the contrary, high-P to high- minerals are preserved within zircon crystals and include coesite,
pressure-intermediate variety denotes the Pacific-type subduction, diamond, garnet, jadeite and silica-rich phengite. These minerals and
also low-pressure type paired with HP type. Furthermore, the mineral assemblages are different from the matrix-forming minerals in
metamorphic facies series were employed to define a clockwise P–T psammitic and pelitic rocks. In this case, UHP metamorphic belts were
path (Fig. 5). The finding of UHP metamorphic rocks from several generated by subduction of passive margin sediments down to 100 or
regional metamorphic belts proved that Miyashiro's generalizations even 200 km depth, dehydrated, and dried at maximum depths.
do not hold true. Regardless of whether the orogeny is collision-type Thereafter they exhume along the Benioff plane up to mid-crustal
or Pacific-type, the regional metamorphic belts display progressive levels of 10 to 15 km depth, where extensive hydration and recrystal-
metamorphism under HP to UHP conditions. Also, the P–T–time path lization occurred when most of HP–UHP minerals disappeared except
is not clockwise, but anticlockwise as will be shown in a later section. those occurring as inclusions within robust minerals such as garnet and
zircon (Masago et al., 2010-this issue). This critical observation refuted
4.2. Re-examination of mineral isograd the generally accepted notion that the composition and assemblages of
matrix minerals are the basic starting point (e.g., Spear, 1993) to
We will now examine the question whether it is possible to understand regional metamorphism. Some of the fundamental concepts
reconstruct the progressive metamorphism during subduction. Major thus collapsed warranting a new framework of metamorphism.
minerals of pelitic and psammitic metamorphic rocks in collisional belts
are plagioclase, quartz, biotite, muscovite and amphibole which are 4.3. Tracing prograde metamorphism from zoned garnet
obviously not the products of the prograde high-pressure meta-
morphism during subduction as discussed before. If so, what were the Garnet in pelitic and psammitic metamorphic rocks is the only major
minerals and their composition, the stable mineral assemblages and rock-forming mineral, which may have been stable during prograde
reactions during the progressive stage? Can these be retrieved? metamorphism. Nevertheless garnet was not in equilibrium with matrix
Compared to pelitic and psammitic schists, metabasites are massive minerals. Since zoned garnet has been regarded as the ideal mineral to
and more competent and therefore, the central portion of a metabasite draw pressure–temperature change through time, garnet solid solution
body tend to preserve the progressive stage of metamorphism even if chemistry has been used most frequently to estimate P–T–t path and a
the unit witnessed late-stage hydration–recrystallization. Thus careful large number of papers have been published. However, some of the
field investigations, mapping and detailed sampling is essential to trace underlying concepts were erroneous. We cannot use matrix minerals at
the signature of progressive metamorphism as illustrated in the case of all, because of the late-stage extensive hydration–recrystallization. Is it
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 113

possible to reconstruct prograde change of mineral assemblages from metamorphism, some of the fundamental assumptions in studies,
pelitic and psammitic regional metamorphic rocks? There are two ways which traced clockwise P–T paths, were erroneous. A more realistic
to estimate the prograde P–T path from the zoning of garnet. approach is based on mineral inclusions within garnet and zircon,
The inversion using Gibb's method (e.g., Spear et al., 1990; which show progressive compositional change during progressive
Okamoto and Toriumi, 2001; Inui and Toriumi, 2002; Okamoto and metamorphism. The results indicate a remarkable difference from the
Toriumi, 2004) starts from a provided initial condition and traces solid previous approach. Thus, (1) maximum pressure estimates changed
solution chemistry of garnet and other equilibrium minerals to from less than 12 kbar up to 70 kbar; (2) clockwise path in P–T space
reproduce the prograde P–T change. This method is a feasible turned to anticlockwise; and (3) the P–T time path was identical to the
approach, although if the prograde garnet and other minerals have metamorphic facies series (Masago et al., 2009) as shown in a later
been lost during retrogressive metamorphism this method cannot be section.
used. On the other hand, Gibb's method does not include enthalpy for
the calculation and uncertainties derived are minimal. Another 4.5.2. Low-grade metamorphism
method is a forward modeling of the mineral assemblage by extensive The discovery of UHP–HP metamorphism radically revised the
thermodynamic calculation using an internally consistent thermody- concepts of high-grade metamorphism, although the finding was
namic database of the minerals such as Holland and Powell (1998). restricted to the highest-grade metamorphic rock such as eclogite. A
The calculation predicts equilibrium mineral assemblages in P–T–X similar misunderstanding was prevalent in the case of low-grade
space. Even though the matrix minerals, except garnet, are totally metamorphic rocks. Metabasites under low-grade metamorphic condi-
retrograded, the prograde P–T path can be estimated from mineral tions of pumpellyite–actinolite facies sometimes show the texture of
inclusions in garnet grain and composition of garnet near by the glaucophane enclosed by actinolite (Fig. 7) (Maruyama et al., 1986). This
inclusion. If some geothermobarometer could be applied to the texture indicates early high-pressure conditions followed by tempera-
inclusion and garnet, the estimate will become more quantitative ture increase at constant pressure to stabilize actinolite because the
(Omori and Masago, 2004; Masago et al., 2009). However, such a reaction between glaucophane and actinolite has a gentle, positive
forward calculation needs enthalpies of phases, which is a parameter slope. The idea of a clockwise P–T path was derived from the concept of
with a large uncertainty. Both the methods have advantages and underplating where an accretionary complex is transported to about
disadvantages and therefore if these are used in combination, a 15 km depth by descending oceanic plate and underplated against the
realistic P–T path can be obtained. hanging wall. In this case, the transportation to 15 km depth takes only
0.2 million year if the subduction speed is 10 cm/year. Since this time
4.4. The Sambagawa metamorphism span is geologically very short, the resultant scenario is only pressure
increases at constant temperature. However, once accreted to the
With increasing metamorphic grade, the minerals in pelitic schist hanging wall, then temperature increases at constant depth since the
range from chlorite, garnet and biotite to oligoclase. These minerals hanging wall is always of higher temperature being part of the mantle
have long been believed to be the products of progressive metamor- wedge.
phic reactions. That is, these minerals were formed by the dehydration However, the reaction mentioned above is not dehydration, but a
reactions by plate subduction with resultant pressure–temperature hydration reaction and substantial volume of water is necessary to
increase to reach P–T maximum under which oligoclase was formed at generate actinolite through the consumption of glaucophane. Meta-
maximum-P and T (Fig. 6). The estimated P–T conditions were 12 kbar, morphic belts require the supply of large amounts of water from
600–700 °C. The contradiction then was that oligoclase is not stable outside. Therefore, this texture can be explained by pressure decrease
under these conditions. at constant temperature instead of isobaric temperature increase
If all minerals except garnet were formed during hydration (Fig. 7). Important process here is the infiltration of water-rich fluids,
recrystallization at later stage, the progressive metamorphism in the similar to the case of eclogite at highest grade. The lower grade units
Sambagawa belt needs to be re-examined, particularly using zoned underlying the regional metamorphic belts such as those of the
garnet and its inclusions at different metamorphic grades. Studies on Shimanto belt contain large amounts of hydrous minerals, indicating
these lines by Okamoto and Toriumi (2001, 2004), Inui and Toriumi the infiltration of large volumes of hydrous fluids during the
(2002), and Aoya et al. (2003) led to the identification anticlockwise exhumation from depths to shallow levels. The structurally interme-
P–T path instead of the previously believed clockwise P–T path for the diate units of the Sambagawa belt contain eclogite facies rocks that
Sambagawa belt. show the highest P–T grade (Banno et al., 1978; Banno and Sakai,
Quartz schist and calcareous schist, which are impervious to fluid 1989; Ota et al., 2004) (Fig. 8). These are remarkably hydrated and
infiltration, tend to preserve the history of progressive metamorphic recrystallized to amphibolite. Below the structural intermediate level,
minerals. The identification of talc, phengite, paragonite and jadeite lower pressure and temperature rocks are present which are more
from pelitic schist shows that although these rocks are minor hydrated and recrystallized than those in the highest-grade meta-
components in HP–UHP belts, they provide valuable information on morphic core at the structurally intermediate level (Fig. 8). The source
the progressive nature of metamorphism. The studies also indicate of fluids would have been the underlying low-grade metamorphic
that plagioclase was not stable at the highest pressure and rocks of the Shimanto belt (Fig. 8).
temperature conditions (Aoki, 2009).
4.5.3. Re-examination of metamorphic age
4.5. A re-examination of pressure–temperature path If continent–collision-type regional metamorphic belts were
formed by overlapped continental crust, the clockwise P–T–t
4.5.1. Drastic shift in pressure estimate and P–T–t path exhumation of regional metamorphic belt must have occurred within
P–T path of metamorphic belts have been calculated by zoned an extremely short time span. From the beginning of subduction
garnet in many cases, although in most examples plagioclase is through transport to great depths and final exhumation to the shallow
present as a stable matrix mineral and all the minerals stable below levels must have occurred within a few million years as calculated by
12 kbar are assumed to be in chemical equilibrium. The results always the numerical simulation of overlapping continental crust (see Fig. 3).
show clockwise P–T paths below 12 kbar. On the contrary, the actual On the contrary, zircon U–Pb SHRIMP ages clearly indicate a much
pressures of prograde metamorphism were 3 to 7 times higher (30 to longer time span than that predicted above. For example the SHRIMP
70 kbar) and the nature of P–T path could also have been different. age of zircon, which contains coesite inclusions, shows 48 Ma at 100 km
Because plagioclase is unstable at the highest grade during regional depth in the Pakistan Himalaya (Kaneko et al., 2003). After return from
114 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

Fig. 6. (a) Metamorphic zoning on the map view. (b) and (c) Mineral appearance/disappearance diagram for Sambagawa pelitic schist in central Shikoku (Higashino, 1975, 1990;
Enami, 1982), (d) Interpretation for progressive mineral change during subduction. Higher-grade index minerals indicate higher P and T, hence oligoclase must be highest P and T
minerals during subduction. (e) Progressive change of mineral assemblage in pelitic system where oligoclase is not stable at highest grade (Aoki, 2009). Late-stage extensive
hydration caused the birth of Barrovian minerals such as biotite and oligoclase during the return pass to the surface.

this depth to the mid-crustal level (15 to 10 km depth), the extensive in the crustal duplication model. Obviously, the field observations
hydration recrystallization occurred at 25 Ma, as plagioclase occurs as refute the model of continent duplication as discussed earlier in Fig. 3.
inclusions the zone developed in the zircons at this stage. These results In the case of Himalayan metamorphic rocks, most of the radio-
show that it took about 23 million years for the return journey from UHP metric ages of K–Ar and Ar–Ar concentrate at about 25 Ma derived
depths to the Barrovian recrystallization depths. Moreover, the drill core from phengite or muscovite (Searle, and Freyer, 1986; Hubbard and
data of Bengal fan deposit in the Indian Ocean (ODP Leg 116) show that Harrison, 1989; Copeland et al., 1991; Guillot et al., 1994). According
the mountain building in the Himalayas began since 8 Ma (Derry and to the previous speculation, 25 Ma is the timing when the high
France-Lanord, 1996). This is grossly inconsistent with the exhumation temperature regional metamorphic belts pass through the depth
rate of regional metamorphic belts from mantle depths to mid-crustal equivalent to 400 °C by the exhumation of deep-seated regional
level. Therefore, mountain building of the Himalayas seems to be metamorphic belt to the mid-crustal depth. Therefore, peak meta-
unrelated to the event of regional metamorphism. morphism must be much older than 25 Ma. Recent observations
Similarly, the time span for transport from mantle depths of 100 to suggest that the concept of closure temperature might be wrong, and
200 km to the mid-crustal level took about 30 million years that the K–Ar dates may represent recrystallization ages at the above
(Katayama et al., 2001) (Fig. 9). The time span for the subduction depths by extensive fluid infiltration underneath. One way to test the
and exhumation is one order of magnitude higher than that envisaged closure temperature is the spot analysis of white mica formed during
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 115

Fig. 7. P–T path estimation from the growth texture of amphibole (Maruyama et al., 1986).
Attached figure shows crossite mantled by actinolite. The texture can be explained by
either isobaric heating (a) or isothermal decompression (b). In either case extensive
hydration is necessary. Crs — crossite; Act — actinolite; Chl — chlorite; Ab — albite.

the progressive stage and which now occurs as mineral inclusion


within garnet or zircon and compare with the spot analysis of white
mica from matrix in the same rocks. In the case of Himalayas, if matrix
white mica yields 25 Ma ages from eclogite facies rocks, then
the white mica inclusions within zircon and garnet must be 48 Ma.
If such is the result, then the popular interpretations based on closure
temperature would be wrong. This idea needs to be tested in future
studies. Fig. 9. P–T-time path for the Kokchetav massif (Katayama et al., 2001). Inset figure
shows a zoned zircon with a core which is as old as 1.1 to 1.4 Ga mantled by
In the case of Sambagawa metamorphic belt, a similar problem metamorphic overgrowth with coesite, diamond and jadeite. The outermost margin
exists as that in the Himalayas. The age of prograde eclogite facies includes quartz, graphite and albite. These three zones were dated by SHRIMP yielding
metamorphism is SHRIMP date of 120 Ma (Okamoto et al., 2004). On the ages shown in the figure. Extremely slow average exhumation speed is estimated by
the contrary, the ages obtained from white mica are 60 to 80 Ma by K– this method. (d) Shown on the bottom of the figure are the P–T conditions when high
temperature solid intrusion of HP–UHP belt occurred generating a contact metamor-
Ar and Ar–Ar ages (Itaya and Takasugi, 1988; Takasu and Dallmeyer,
phic aureole (D) at 501 ± 10 Ma with temperatures of 500–650 °C and pressures of 3 to
1990). Whether the 60 to 80 Ma age represents the time of hydration 4 kbar.
recrystallization at mid-crustal level or the much older white mica
under UHP conditions at 120 Ma was reset at 400 °C due to closure
temperature effects need to be examined by spot analyses of mica facies rocks in regional metamorphic belts (Miyashiro 1961, 1965)
inclusions in garnet. (Fig. 10a).
In general terms, metamorphism begins with increasing pressure
and temperature from the surface condition to P–T maximum at depth
5. Re-examination of metamorphic facies series and thereafter the return to surface. Thus, the rocks in regional
metamorphic belts could have been subjected to several metamorphic
It has been previously believed that one metamorphic rock records grades (Fig. 10b). de Roever and Nijhuis (1963) proposed the pioneering
one metamorphic facies. The concept of metamorphic facies series concept of plurifacial metamorphism based on the observation that one
was formulated by integrating the distribution of low to high-grade metamorphic rock preserves several facies under the microscope such

Fig. 8. Hydration effect on the Sambagawa belt from the underlying Shimanto belt. The cross section along the N–S in central Shikoku is schematically shown. Note the underlying
Shimanto belt is very low grade under PA (Pumpellyite–actinolite facies) to PrA (Prehnite–actinolite facies). The structurally intermediate unit of Sambagawa is composed of eclogite
facies rocks formed under 20 kbar and 800 °C (Ota et al., 2004). Peak mineral assemblages and texture are only preserved in the core of large mafic bodies such as the Iratsu eclogite
amphibolite due to extensive fluid infiltration from underneath (Ota et al., 2004).
116 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

curve changes at about 13 kbar and 700 °C to a much straight path


with no significant increase in temperature up to 40 kbar. This trend
appears to be common not only in continent collision zones but also in
Pacific-type orogenic belts (Inui and Toriumi, 2002, Aoya et al., 2003;
Okamoto and Toriumi, 2004; Masago et al., 2009).

6. Re-examination of textures

6.1. Metamorphic texture

Regional metamorphic rocks contain a variety of textures related to


prograde metamorphism, recrystallization and deformation. For
example, albite porphyroblasts often include a variety of minerals
defining internal deformation fabric. Elongate minerals such as
amphiboles define mineral lineations. Boudin or pull-apart textures
are also common which indicate extension and breakdown of the
elongated minerals into several fragments, all of which have long been
considered to have been formed during subduction, i.e., progressive
textures. Also common are snowball garnets which entrail the rotation
and growth of inclusion minerals, as well as pressure shadows. Other
microstructures include sheath folds, deformation of radiolarians and
rounded cobbles and crenulation fold within mica-rich domains. These
textures offer information on the direction of stress and its relative
magnitude under the assumption of boundary conditions and initial
state.
In the Sambagawa belt, mineral lineation and crenulation fold axis
of mica, boudin or pull-apart texture of mineral grain or elongated
Fig. 10. Definition of the metamorphic facies series by Miyashiro (1965). (a) Miyashiro
minerals, and radiolarian deformation, among other features indicate
defined the metamorphic facies series as the curve that connects the P–T conditions E–W extension, which is parallel to the orogenic belt (e.g., Toriumi,
shown as thick solid lines. (b) An individual rock may have experienced several 1982). This elongation coincides with the direction of metamorphic
metamorphic facies in a single metamorphism. The highest-grade rock in the figure fluid flow in which the elongated minerals have grown. These
experienced 5 metamorphic facies as shown as circled numbers 1 to 5. This was earlier
structures have long been considered to form during progressive
defined as plurifacial metamorphism (de Roever and Nijhuis, 1963). Spear et al. (1990)
redefined the metamorphic facies series as a curve that connects the thermal peak metamorphism through subduction during the Cretaceous time and
conditions of the P–T path of individual rock (thin solid line) within a metamorphic belt. relative plate motion. However, all of these mineral textures and
See text for discussion. deformation were not formed during progressive recrystallization,
but were generated at a later stage after the emplacement at mid-
as zoned Ca–Na amphibole and Ca–Na pyroxene. The clockwise P–T time crustal level by the addition of fluids underneath (Toriumi, 1982, 1985
path connecting several metamorphic facies has become a fashion since and Toriumi and Noda, 1986).
1980 from the analysis of regional metamorphic belts over the world. Albite porphyroblast appears at the highest grade in the Sambagawa
Since one rock specimen preserves several metamorphic facies in cases belt and the zone in which abundant albite is present has been termed as
where high-grade metamorphism has operated, Miyashiro's proposal of the spotted zone. Empirically, this zone corresponds to the biotite zone
metamorphic facies series needed redefinition. Thus to avoid confusion, at the highest grade. The timing of formation of the porphyroblast in
the highest T at given P condition for each rock was connected to draw relation to the progressive or retrogressive metamorphism is important
the P–T path and was defined as the metamorphic facies series (Spear to discuss the origin of the metamorphic texture. This texture appears
et al., 1990) (Fig. 10b). restricted only to the highest-grade portion because its development
The metamorphic facies series thus defined is again clockwise in P– has been traditionally linked to progressive metamorphism at the
T space; with increasing depths the temperature gradient tends to highest, and therefore, the deepest conditions. However, the albite
increase. If we consider the facies series characterized by clockwise P–T porphyroblasts include even retrogressive minerals developed at lower
path at depths over 10 kb along consuming plate boundary, the pressure conditions during hydration (Toriumi, 1975). Therefore, the
temperature must increase beyond the solidus of basalt or peridotites, timing of the formation of these porphyroblasts must be after the
and slab must be partially melted. Thus, the subducted materials above emplacement of the Sambagawa belt into the mid-crustal level.
oceanic crust together with the sediments on the top will be melted Presumably, large amounts of fluids infiltrated into the Sambagawa
almost completely. However, observation from natural examples belt, which promoted extensive scale of recrystallization at 80 to 70 Ma.
along the modern subduction zones does not support this, because arc The texture and deformation of the rocks in this belt also indicate lack of
magmatism behind the trench is not in conformity with the idea of preservation of the progressive stage. An E–W extension strain is
slab-melting in the case of old slab N25 Ma (Defant and Drummond, derived from the deformation pattern of gravels in metaconglomerate
1990; Drummond and Defant, 1990). Seismological observations also schist and the recrystallized age of white mica by K–Ar method. The
do not indicate any slab-melting at great depths. conglomerate schist in Oboke area has been considered to be a part of
The P–T change along the descending slab has been calculated the Sambagawa metamorphic belt from deformational patterns.
numerically (for example, Peacock 1996) (Fig. 11a). The results clearly However, LA-ICPMS U–Pb detrital zircon age separated from quartz
indicate an anticlockwise trend, instead of a clockwise path. Such a porphyry cobble in the conglomerate indicates 92 ± 4 Ma (Aoki et al.,
trend is commonly recognized by geophysicists who consider plate 2007). This age represents the timing of consolidation of the acidic
boundary as a thermal boundary layer and it is well recognized that magma for the protoliths of gravels. Therefore, the sedimentary age of
geothermal gradient is not constant with increasing depth. For conglomerate must be younger than 92 Ma. On the other hand, the
example, the Kokchetav HP–UHP units show an anticlockwise path Sambagawa progressive metamorphism is dated as 120 Ma (Okamoto
(e.g. Masago et al., 2009) (Fig. 11b) and the slope of the estimated P–T et al., 2004) clearly indicating that Oboke conglomerate is not a part of
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 117

Fig. 11. (a) Numerically simulated temperature profile along the Benioff plane of a subducting slab at 10 cm/yr (Peacock, 1996). Geotherm is gentler at shallow level usually less than
50 km and turns to steep at deeper with anticlockwise sense of P–T path. In the case of a young slab, it would be gentler than the older, and where the slab is old, it is steeper. Dotted
line C shows the generally estimated metamorphic P–T path with a clockwise sense. If the clockwise path C is correct, subducted slab must have been melted. In the bottom figure,
line A shows the geothermal gradient within plate shown below. Line B shows the geothermal gradient along Benioff plane, corresponding to those paths on the top figure. Also
shown are wet solidus of MORB and anhydrous solidus of MORB. Numbers of 2, 5, 10, and 50 denote the age of subducted slab. (b) P–T estimate of the Kokchetav massif covering
epidote amphibolite facies grade through amphibolite to high pressure granulite, amphiobole– to zoisite–eclogite and dry eclogite (see Masago et al., 2010-this issue). Metamorphic
facies series shows an anticlockwise path, which is different from the traditional clockwise path. Progressive P–T path follows an isothermal decompression during exhumation.

the Sambagawa metamorphic belt. Hence, the so-called Sambagawa subduction. The deformation fabric during progressive metamorphism
belt must be divided into two different metamorphic belts. When the must be restricted to mineral inclusions within garnet and zircon. The
Sambagawa metamorphic rocks were emplaced at mid-crustal level, snowball garnets may provide one of the tools to reconstruct the three-
above the Shimanto belt, both Sambagawa and Shimanto shared the dimensional movement of the progressive deformation.
same history of recrystallization and deformation with E–W extension.
The microfabrics of the Sambagawa schist are remarkably similar to the 7. Discovery of top and bottom boundary of regional
deformation patterns associated with low-grade metamorphism in the metamorphic belts
Shimanto belt. This shows that the metamorphic texture of the
Sambagawa schist does not correspond to the progressive stage, but is One of the fundamental requirements is to define the top and
related to retrogressive hydration stage. The major texture of the bottom boundaries of the regional metamorphic belt, although this
Sambagawa schist was formed during the retrogressive hydration stage has not been resolved in most cases. Metamorphic petrologists have
at 80 to 70 Ma when Sambagawa was juxtaposed to the Shimanto belt. often neglected the presence of discontinuous boundary by simply
The metamorphism of basic and pelitic schist in the Shimanto belt assigning a continuous gradation into the lower grade rocks gradually
reached up to blueschist facies. The zircon U–Pb chronology yielded without considering a discontinuous gap above and below the
84 Ma for the protolith age, and the abundant grained white mica metamorphic units, although structural geologists had predicted the
yielded K–Ar ages around 60–65 Ma for the metamorphism (Aoki et al., presence of a clear boundary. On outcrop scale, a sharp boundary
2008). It is difficult unequivocally define the textures or fabric formed cannot be identified between the two units because of mode of
during progressive stage of metamorphism, and to differentiate the deformation and metamorphic grade do not show any marked
earlier stage of exhumation from mantle depths to the mid-crustal level difference. The maximum thickness of regional metamorphic units
with the presently available information except in specific cases. One such as the Sambagawa reaches up to 2 km. The pressure gradient
such example is a relict eclogite formed at the highest grade of the within the belt remained unresolved. Among the models proposed for
Sambagawa metamorphism. The elongation axis of the omphacite the Sambagawa, a structural break is commonly noted along the
and related textures show orogen-vertical N–S slide. The disposition of boundary of the metamorphic zone such as in the Kanto Mountains
the Sambagawa belt shows a structurally intermediate position, vertical (Hashimoto et al., 1992).
to the strike of the orogenic belt, and pre-dates the development of UHP–HP belts often exhibit remarkable pressure difference within
the E–W lineation (Toriumi and Kohsaka, 1995; Yamamoto et al., the belts. Detailed studies in the Kokchetav unit covering 250 km×
2004). There are minor rocks in outcrop scale or microscale, which show 20 km helped in identifying nearly horizontal top and bottom boundary
the texture or deformation during the exhumation from the mantle (Kaneko et al., 2000). Locally, tectonic windows are present revealing the
depths. However, all of those do not preserve the progressive stage of underlying unmetamorphosed units because the bottom boundary is
118 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

nearly flat (Kaneko et al., 2000). And more interestingly, a contact structurally intermediate unit, the Penninic nappe, was tectonically
metamorphic aureole in the unit underlying the Kokchetav UHP–HP unit inserted into the overlying and underlying units because this structural
was also identified. This contact aureole is divided into four zones and the intermediate unit is the only unit that was subjected to the Eocene–
metamorphic grade progressively increases towards the boundary. Miocene HP–UHP conditions of metamorphism. In the prevailing
Metamorphic facies series belong to andalusite–sillimanite type and concept, the reconstructed three units are assumed to have possessed
contact aureole reaches up to 600 °C at 2 kbar (Terabayashi et al., 2002). a horizontal relationship as shown in Fig. 13. However, a prominent gap
Metamorphic zircon separated from contact zone yielded SHRIMP U–Pb is evident between units H and P in Fig. 13, a marked discontinuity in
age of 501± 10 Ma which coincides with the age of the extensive terms of lithology, stratigraphy and geochronology. These three are not
hydration and recrystallization stage within the HP–UHP units continuous each other and lithostratigraphically different before nape-
(Katayama et al., 2001). This observation indicates that the contact stacking (Fig. 13).
metamorphism was a result of the emplacement of the HP–UHP units Underneath the Kokchetav HP–UHP unit, a well-developed contact
into upper crustal level at 6 km depth. Thus, a new concept of the metamorphic aureole has been defined with andalusite–sillimanite
exhumation of diamond-grade metamorphic rocks and their high type facies series (Terabayashi et al., 2002), clearly demonstrating
temperature solid intrusion emerged, similar to magma intrusion. This that the exhumation of regional metamorphic belts was not buoyant
idea clearly demonstrated that a radical change from the traditional uplift as proposed by England and Thompson (1984) but a high
concept of buoyancy-driven upward exhumation and erosion of regional temperature intrusion into shallow crustal level as discussed in the
metamorphic belts is undoubtedly wrong. previous section. Moreover, the emplacement depths were shallower
A suite of 7000 samples were collected from the Kokchetav from than 12 km (4 kbar) and the highest temperature recorded from the
which 350 were analyzed by electron microprobe and zircons margin of metamorphic belts is 600–650 °C. This indicates that the
separation performed from 250 samples to evaluate the detailed P–T innermost domain of the metamorphic belts must have witnessed
structure. The study yields maximum pressures of 60–70 kbar at 900– much higher temperatures, albeit lower than 1000 °C, when these
1000 °C (see a summary by Maruyama et al., 2002). Metamorphic rocks were at nearly 200 km depth. Due to this reason, when the
units within the 2 km thick belt were sub-divided into five mineral metamorphic belt was emplaced, large volume of fluids must have
zones with the structurally intermediate zone showing the highest infiltrated with the local development of migmatite.
grade. The metamorphic grade decreases both upward and downward From the structural analysis of the interior of the Kokchetav
with a pronounced pressure gradient with only 5–6 kbar at the metamorphic belt, the root zone was estimated to the south, with a
boundaries. However, the boundaries of mineral zones are continuous northward movement as a sub-horizontal tectonic slice during exhu-
without any sharp structural breaks, indicating ductile deformation to mation. The average exhumation speed was extremely slow, 5 mm/year,
emplace those mineral zones upwards after the thermal peak. Overall as estimated by Katayama et al. (2001). There was no mountain building
thermobaric structures show symmetric pattern with an axis at the on the surface until the metamorphic unit was emplaced at depths
structural intermediate. However, the thickness of each mineral zone shallower than 12 km. The major period of formation of sedimentary
is not symmetric above and below. The deformation structure basin was after the doming stage, and developed by high angle normal
indicates that the structurally intermediate zone moved from south faults after extensive hydration, equivalent to the post-orogenic
to north (Kaneko et al., 2000; Fig. 12c). mountain building stage. Post-collision orogenic granites intruded into
In the case of Pacific-type metamorphic belts such as the the metamorphic belt (Katayama et al., 2001). Traditionally, the main
Sambagawa, the method to define top and bottom boundaries had stage of orogeny was regarded as the exhumation or uplifting of regional
already been proposed by Isozaki and Itaya (1990). The precise ages metamorphic belt, which was also regarded to be the major mountain
were measured from radiolarian microfossils and K–Ar dating of fine- building stage. However, we consider that this concept is incorrect, as
grained white mica separated from low-grade metamorphic rocks exhumation of regional metamorphic belt to the mid-crustal level occurs
along the outcrop scale. Later, Nishimura (1990) used the crystalli- much earlier and without mountain building. The mountain building
zation index of carbonaceous material in the outcrop scale to precisely stage is a post metamorphic event through doming as first proposed
fix the boundary in meter scale. The importance of defining top and Maruyama (1990) and Maruyama et al. (1996) based on wedge
bottom boundaries had been pointed out in earlier studies with the extrusion model. A similar scenario was also recently proposed by
boundary defined either top or bottom, or both for forty examples of Santosh et al. (2009a) to explain the metamorphic belts in the Cambrian
regional metamorphic belts from different regions of the world collisional orogen of Gondwana in southern India.
(Maruyama et al., 1996).
9. A new model of exhumation
8. Chaotic state arising from the concept of nappe tectonics
9.1. “Intrusion” of regional metamorphic belts
The Penninic nappe of the Alpine metamorphic belt is the best
example where the top and bottom boundaries are well defined with The model proposed by England and Thompson (1984) marks the
the intermediate zone enclosed within overlying and underlying pioneering tectonic synthesis to explain the formation of continent
unmetamorphosed units (Fig. 12a). However, the origin of UHP–HP collisional type orogenic belts and their exhumation employing
unit has not been well understood largely due to the confusion arising metamorphic petrology as a powerful tool. If the double crustal model
from the concept of nappe tectonics (Fig. 13). In the following section they envisaged is valid, the structural bottom of the regional meta-
we briefly address this problem. morphic belts must continue to higher-grade metamorphic rocks up to
In this metamorphic belt, there are three tectonic slices demarcated the Moho depths, and underlain by mantle peridotites. On the contrary,
by horizontal faults (Coward and Dietrich, 1989; Maruyama et al., 1996). the units underlying the regional metamorphic belts are composed of
The structurally lowest unit, the Helvetic nappe which constitutes the weakly metamorphosed or unmetamorphosed rocks. Moreover, the
basement of Europe has been classified as an allochthon that has never thermobaric structure of the interior of the regional metamorphic belts
moved after its formation. The overlying unit (Penninic nappe UHP–HP does not support the model, which predicts unidirectional continuous
unit), which was metamorphosed in the Eocene to Miocene, moved pressure temperature increase to the bottom. As mentioned in previous
from south to north and was thrust over the allochthon. Thereafter, the sections, the well-examined metamorphic belts such as Kokchetav
topmost unit, the Austro-Alpine unit or nappe which is considered as a always show structural intermediate position for the highest-grade
basement of Africa, was again thrust over on the top. Although this has units as well as the Himalaya and Alps. The three-dimensional structure
been the prevailing view, it is more reasonable to consider that the indicates marked thermal and baric gradients within the thin (2–3 km)
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 119

Fig. 12. Cross section of representative orogenic belts. (a) Alps (Maruyama et al., 1996); (b) Himalayas (Kaneko, 1997); (c) Kokchetav (Kaneko et al., 2000); (d) Sanbagawa (Masago
et al., 2004). Note the presence of sandwiched structure of all regional metamorphic belts enveloped by low-grade or weakly metamorphosed units above and below. Also note the
large vertical exaggeration to show the internal structure of HP–UHP belts. If there is not vertical exaggeration, the UHP–HP belt will appear as an extremely thin line with aspect
ratio of 1:100.

high-grade unit without any fault boundaries inside, suggesting that morphosed or unmetamorphosed geological units. The average exhu-
ductile deformation prevailed after the peak metamorphic conditions mation speed was extremely slow from mantle depths of 200 to 100 km
and the high-grade metamorphic unit ‘intruded’ into the weakly meta- to the mid-crustal levels of 10 to 15 km, at about 4 mm/year (Katayama
120 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

25 Ma but the mountain building stage has not yet begun. At 8 Ma, the
Himalayan metamorphic belt uplifted by doming and extensive
development of secondary normal faults together with corresponding
formation of Indus–Bengal submarine fan deposit (Fig. 14). Thus, the
formation of large folded mountain belts occurs as the final stage of
exhumation and is unrelated to the main stage of exhumation of the
metamorphic belts from mantle depths to the mid-crustal level. The
Fig. 13. Schematic diagram to show nappe tectonics in the Alpine orogeny. The Himalayan mountain belt was reactivated since the last 1 million year
basement allochthon by H (Helvetic nappe) was overthrust by P (Penninic nappe) possibly in response to the southward jump of plate boundary and
which in turn is overthrust by A (Austro-Alpine unit).
buoyant continental crust was underplated beneath the higher
Himalayas to push up the Himalayan mountain chain (Sakai, 2002).
et al., 2001). The present day ongoing uplift rate measured in Himalayas,
Taiwan and outer Indonesian non-volcanic arc shows only a few cm/
year. The exhumation rate of HP–UHP units up to the mid-crustal depth 9.3. Dynamics of extrusion
is one order of magnitude slower.
The aspect ratio of regional metamorphic belts shows ca. 1:100 in the
case of the Himalayan metamorphic belt, the thickness in only a few km
9.2. Mountain building whereas the belt continues over 1000s of km along the strike with 200–
300 km width. Thus, regional metamorphic belts are thin and platy
The mountain building stage is the second step triggered by which are cut by paired faults: normal fault on the top and reverse fault
doming which occurred after the emplacement of regional metamor- on the bottom, and the belt is extruded above the Benioff thrust on the
phic belts into the mid-crustal level and the development of high consuming plate boundary. The exhumation velocity is one order of
angle normal faults surrounding the dome generating large sedimen- magnitude smaller than the plate movement, which has been proposed
tary basins. The best-documented example is the Himalayan moun- by Maruyama (1990) and Maruyama et al. (1994, 1996). These features
tain building caused by Indian collision against Eurasia (Fig. 14). The can be attributed to the following three processes. (1) Subduction angle
initiation of collision of India against Eurasia started at 50 Ma and UHP changes from deep to shallow through time. The subducted wedge-
metamorphism occurred at 48 Ma (Kaneko et al., 2003). The UHP–HP shaped continental material at depths was extruded by shallowing of
unit was emplaced from mantle depth to the mid-crustal level at the subduction angle. (2) The rheology shows marked contrast in the

Fig. 14. (a) Map view of Indian collision and formation of Indus and Bengal fans, which were derived from the mountain building associated with Himalayan orogeny. Total volume of
both fans, which have maximum thickness about 5000 m, is similar to the mountain belts in the Himalayas above sea level. (b) Modal change of clay minerals and Sr-isotope values
through time since 20 Ma for the drill core sediments at ODP site 717–719 shown on (a) (Derry and France-Lanord, 1996). The figure clearly shows that 8 Ma is the turning point for
the rapid increase of detrital material in the fan, indicating uplift of the Himalayan mountain belt. (c) Elevation of Himalayan mountain belt as a function of time back to 50 Ma. HP to
UHP metamorphism started ca. 50 Ma followed by retrogressive hydration around 25 Ma. Extensive mountain building started around 8 Ma and gradually dropped with recent
reactivation probably 1 Ma ago.
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 121

units above and below the regional metamorphic belts. The regional exhumation speed would depend on the deformation speed in the
metamorphic belt is composed of quartzofeldspathic material which wedge region. Deformation speed within the wedge is assumed to be
turns ductile at temperatures above 350 °C, whereas the overlying constant. Exhumation speed must be faster at the deeper regions and
wedge mantle is composed dominantly of olivine and the underlying tends to be slower near the surface. In the above example, the
unit is eclogitic oceanic crust and peridotites which behave as brittle exhumation speed at 120 km depth reduces to one third near the
below 1000 °C. (3) Rheological change during exhumation of quartzo- surface and the material during the exhumation is stretched five times
feldspathic regional metamorphic belts. The regional metamorphic belt (free slip) to 20 times (frictional slip) along the exhumation direction. In
moves along the Benioff thrust up to mid-crustal level and becomes the laboratory experimental studies performed by Chemenda et al.
stagnant at depths of 10 to 15 km because the quartzofeldspathic (2000) to understand the exhumation mechanism of Himalayan
materials transform from ductile to brittle by the temperature decrease regional metamorphic belts, paired faults were clearly reproduced and
at this stage, becoming harder. The 350 °C temperature divide corre- the model of wedge extrusion well illustrated.
sponds to the brittle–ductile transition for the quartzofeldspathic
material at mid-crustal depths of 10 to 15 km. Thus, regional meta- 9.6. Second stage uplift and mountain building
morphic belts, which behave ductile fluid, are inserted between the
overlying rigid mantle wedge and the underlying rigid slab. When the Wedge extrusion mechanism can exhume the regional metamorphic
subduction angle turns to shallower, then the stress field changes to belts along Benioff thrust from mantle depths to the mid-crustal level.
trigger the wedge to extrude into mid-crustal level. This is the most However, another mechanism is necessary from there to move the
essential dynamic process controlling the exhumation of regional meta- material to the surface. This is the second stage process of doming and
morphic belts. mountain building by domal uplift of the unit, which is sandwiched as
the core of the regional metamorphic belt. The typical example is the
9.4. Regional metamorphic belts brought up by buoyancy uplift? Himalayan mountain building, which started since 8 Ma triggered by
the southward jump of plate boundary. Along the new plate boundary,
The density of quartzofeldspathic rocks is about 2.8 g/cc which is continental crust began to subduct underneath the Himalayas and
lower than that of peridotites (3.2 to 3.3 g/cc for depths up to 50 to jacked up part of higher Himalayas (Takada and Matsuura, 2002)
60 km at which stage amphibolite transforms to eclogite with a (Fig. 16).
density 3.3 to 3.4 g/cc) (Komiya et al., 2002; 2004). These values
continue to a depth of 250 km with a maximum of 5% variation. Below 9.7. Modern analogue of regional metamorphic belts and exhumation
250 km depth, K-hollandite (density 4.7 g/cc) and stishovite (density mechanism
4.3 g/cc) become stable. Above 90 kbar, the density of a quartzofelds-
pathic rock is larger than the mantle material (Irifune et al., 1994) and It is interesting to examine why regional metamorphic belts were
therefore buoyancy does not aid in the exhumation of regional generated episodically and exhumed to the surface to through time.
metamorphic belts from depth. On the other hand, it turns to become During the last 500 million years along the circum Pacific region such
a strong slab-pull force. as Japan and California, glaucophane-bearing regional metamorphic
Between 0 and 70 kbar, a quartzofeldspathic rock is buoyant than belts were formed and exhumed to the surface every 100 My on an
mantle peridotites or eclogites. However, the small density difference average (Maruyama et al., 1996). New orogenic belts were formed
would not help the uplift of regional metamorphic belts as a plume underneath and/or oceanward through time episodically. In the case
within overlying mantle wedge (Fig. 15). More difficulties for the of continent collision-type regional metamorphic belts, although the
buoyancy uplift mechanism include the size of regional metamorphic occurrence and timing vary among different cases, our synthesis
belts and three-dimensional shape which is extremely different from shows that about 30 collision-type orogenic belts were formed on the
the mushroom shape predicted in buoyancy-driven models. Moreover, Phanerozoic earth by continent collision (Fig. 4).
as discussed in a previous section, the aspect ratio of the metamorphic In the case of continent collision, the formation and exhumation of
belt is around 1:100, and is quite different from the plume shape, and regional metamorphic belts appear to be related to the size of
platy. If size is too small, even though density difference is large enough, continent. For example, the size of continents to form UHP units
buoyancy uplift is not effective. Thickness of about 1 to 3 km is shows a wide range from India, North China, and Western Europe to
extremely small compared to the whole crust in space. Simple small continents such as Madagascar and Java Island in Indonesia
calculation never supports that it can penetrate into mantle wedge (Liou et al., 2000, 2002; Parkinson et al., 2002). In the case of the
below 1000 °C. Mode of occurrence of regional metamorphic belts in the continents smaller than the size of Madagascar or arc, we do not
field clearly refutes the idea of plume because of the shape. normally see any example of UHP metamorphic belts. We examine
this aspect in more detail below.
9.5. The wedge extrusion model Whether the subducted continental crust is accreted to the
hanging wall without subduction into deep mantle, or returned
The wedge extrusion model proposes a wedge-shaped viscous back to the surface depends on a number of factors including size
fluid (regional metamorphic belt), sandwiched by the underlying slab (buoyancy) of the continent and the degree of coupling between the
and overlying mantle wedge on both sides, that extrudes through the subducting slab and the overlying continental crust. The descending
shallowing of the subduction angle (Maruyama et al., 1996). The Pacific plate immediately before the trench outer wall of the Japan
wedge region A or B on the corner of the triangle (Fig. 15) is assumed Trench in NE Japan is bent to generate horst and graben structures.
to be filled by a non-compressive fluid. The rigid wall OA moves The grabens have a maximum depth of 500 m and are a few km wide,
around the center of point O and extrude the fluid AOB. With regard to which are then filled by quartzofeldspathic trench turbidites. These
boundary conditions, OA and OB behave as either as a free slip or sediments are being subducted into the deep mantle together with
frictional boundary surface. In a typical case of angle AOB shown in Pacific MORB crust (Miura et al., 2003). Buoyancy derived from graben
Fig. 15, the initial conditions are alpha equals 30°, beta equals 5°, and filled sediments would not be effective because of the relatively small
depth of the wedge is 200 km. If the angle of the wedge beta turns to size or volume. However, in the case of intraoceanic arc subduction,
0.5°, it would lead to extrusion of material from 60 km depth to the such as the Kyushu–Palau paleo-arc and Amami plateau (paleo-arc)
surface either along a free slip boundary or frictional boundary. with thickness less than 20 km, subduction proceeds into deep mantle
To move material from a depth of 120 km to the surface, around 80% without any accretion (Kodaira et al., 2000; Yamamoto et al., 2009). In
deformation is required at the wedge triangular region. The average the case of Izu Mariana arc with a thickness of continental crust of
122 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

Fig. 15. Dynamic process of exhumation of regional metamorphic belt. (a) Buoyancy-driven plume model. It seems to be unrealistic that plume-like metamorphic unit penetrates the
rigid mantle wedge and reach finally to the surface. (b) Wedge extrusion model driven by narrowing of the wedge angle between the subducting slab and mantle wedge. (c) Wedge
extrusion model proposed by Maruyama (1990) and Maruyama et al. (1996) that a ductile metamorphic unit intrudes into shallow-level brittle units through two-steps.

about 30 km, at least some of the intraoceanic arc tends to be accreted trench, but parallel to the trench, even 20 km thick arc crust can be
which indicates adequate buoyancy to overcome the coupling force accreted to the hanging wall, a process that was presumably common
exerted by descending slab. Note that buoyancy depends on the in the Archean.
volume, hence if the collision of arc is not vertical or tangential to the We now consider the stress system that promotes the exhumation
of UHP–HP metamorphic belts using modern analogues. Regions of
our interest are shown in Fig. 17 which is a world map to show the
distribution of plate boundary. The ongoing exhumation area of
regional metamorphic belt must be along the active consuming plate
boundary. In the case of collision-type regional metamorphic belts,
the best modern analogue is the non-volcanic outer arc in Timor–
Tanimbar region of Indonesia (Fig. 18). Along the non-volcanic arc
from Timor through Tanimbar to Salem, a U-shaped non-volcanic
outer arc continues over 2000 km. Miocene glaucophane-bearing
regional metamorphic belts are exposed in southwestern part of this
region for about 700 km long and in the northwestern part for over
another 600 km, although not in the central part (Ishikawa et al.,
2007; Kaneko et al., 2007). To the west of the N–S trending non-
volcanic arc, a forearc basin (Weber Deep) is developed with a depth
of up to 7000 m. The region itself is underlain by strong extensional
Fig. 16. Jack-up model with ramping of the subducted continental crust, which explains state (Charlton et al., 1991). On the contrary, the domains to the north
the second stage uplift of the Himalayan orogen (Takada and Matsuura, 2002). HFT —
Himalayan Frontal Thrust, MBT — Main Boundary Thrust; MCT — Main Central Thrust.
and south are underlain by strong N–S compressional state (Fig. 18).
Note the large vertical vector around MCT for the uplift of MCT by the ramping of The metamorphic belts towards the SW part, particularly the
buoyant continental crust underneath, shown by a broken line. southwestern end, correspond to the late stage of collision orogeny
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 123

Fig. 17. World map showing ongoing orogeny along the active consuming plate boundaries of the Earth. Six examples are marked, the details of which are explained in later figures.

and the island itself is uplifted as a dome showing the development of north of the volcanic arc to the trench is underlain by the strong N–S
a number of normal faults on the island to uplift Quaternary reef compressional state (Fig. 18).
limestone above 1300 m. The top of the folded mountain in the Timor On the other hand, to the west of Timor, the Indo-Australian plate
reaches 2960 m and shows the highest elevation speed over the world, is completely separated from the descending oceanic slab, which is
0.4 to 0.6 cm/year for the last 0.5 million years on an average. To the east sinking down into deep mantle. The subduction angle of the
of the outer arc, the elevation rate tends to be small. In the case of Australian continental lithosphere has turned shallower thereby
Laibobar Island, 700 km away from Timor, the top boundary of pushing up the entire Timor Island. The continued uplift of Timor
glaucophane-bearing metamorphic belt is exposed above the sea level Island starting from 0.5 Ma has resulted in the present day peak of the
(which is a normal fault) and the Quaternary reef has not been uplifted mountains reaching over 2960 m.
for the last 1 million year. The difference in terms of uplift clearly shows We now consider ongoing example from the Pacific-type orogenic
the time difference in the stage of orogenic evolution, and confirms that belts. Whether continent collision-type or Pacific-type regional
domal uplift has not been initiated to the east. On Laibobar Island, minor metamorphic belts fundamentally record the same structural pattern;
normal faults were identified which indicate that the domal stage may that is, a sandwich structure and two-step exhumation process
have started very recently. However, the Timor Island has already been followed by extensive mountain building at the latest stage. This
uplifted vertically over 1300 m. Under the Tanimbar Island, which is the suggests that a similar mechanism operates to exhume the regional
easternmost portion of the non-volcanic arc, domal uplift has not yet metamorphic belts regardless of their category as collision or Pacific-
begun and the leading edge of the subducted Australian continental type. As mentioned in an earlier section, the fundamental control is
plate is being broken off from the foregoing descending oceanic exerted by the change in subduction angle to shallow values. In the
lithosphere (Osada and Abe, 1981) (Fig. 18b). The boundary of break- case of Pacific-type, the shallowing subduction angle and the
off between continental lithosphere and oceanic lithosphere is approach of mid-oceanic ridge to the trench are well-studied. It is
characterized by high seismicity, vertical to the seismicity in the Benioff considered that in the circum Pacific orogenic belts, mid-oceanic ridge
plane (Osada and Abe, 1981). The seismological characteristics suggest subduction has occurred once every one hundred million years on an
slab break-off and that the buoyant continental lithosphere cannot average (Engebretson et al., 1985).
subduct deeper than the present, a constraint posed by the more dense Some of the youngest Pacific-type orogenic belts occur in the
older oceanic lithosphere which exerted a higher slab-pull force leading Olympic Peninsula, southwest of Seattle near Vancouver (Fig. 19a)
to the slab break-off. To the west under the Timor Island, the continental (Maruyama et al., 1996). This is the region of Quaternary uplift cut to
lithosphere is already released from oceanic slab and hence the free the east by active normal fault, and to the west by active reverse fault.
buoyant subducted continental slab promotes exhumation of regional These two faults seem to be paired and sandwich the thin ‘platy’
metamorphic belts as shown in Fig. 18. geologic unit 20 km × 100 km in space. The core of this wedge-shaped
Thus, the process of exhumation is related to slab break-off and unit contains low-grade lawsonite-bearing schists formed at about
resultant change in the subduction angle from deep to shallow 11 km depth (Brandon and Calderwood, 1990). The wedge extrusion
thereby promoting the subducted wedge to remove to arc–trench gap here started in the Miocene to expose the deep-seated portion of an
(Fig. 18b) (Maruyama et al., 1996). Tanimbar illustrates the stage just accretionary complex. Apatite fission track age shows 7 to 11 Ma
after slab break-off and complete release of slab. On the other hand, (Brandon et al., 1988; Brandon and Calderwood, 1990).
the case of Timor shows a further advanced stage of second domal The Juan de Fuca plate subducting underneath the Mt. Olympus is
uplift. On the contrary, 100 km north of Timor, the volcanic arc has very young, ca. 5 Ma (Engebretson et al., 1985). To the west, the N–S
already shut down its magmatic factory to the north of volcanic arc, trending active Juan de Fuca ridge is present which moves eastward at
and a south dipping back thrust is developed (Hamilton, 1979; Harris, the rate of 3 cm/year and which will finally subduct after few million
2006; Nugroho et al., 2009). The active back thrust disappears to the years. Considering the last 20 million year movement of the mid-
northeastern end of Timor indicating that the entire region from the oceanic ridge, the age of subducting slab underneath is gradually
124 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

Fig. 18. Ongoing exhumation of continent collision type metamorphic belt in Timor–Tanimbar region of Indonesia (partly modified from Maruyama et al., 1996). (a) An outline of
tectonic environment of the Indonesian region. Note the contrasting stress field between the west (compressional) and the east (extensional). (b) Cross section along A–B in the
extensional environment. Forearc region is underlain by the compressional stress field whereas the back arc side is dominated by extension to form the 7000 m deep Weber Basin.
Beneath the arc, slab break-off is ongoing. (c) Western side N–S cross section of the Timor Island. Note the location shown on (a). The entire region is underlain by strong N–S
compressional stress field and the back thrust is developed on the back arc side. Oceanic slab has been disconnected with the resultant buoyancy of continental lithosphere. The
regional metamorphic belt exposed on the surface is in the second stage of exhumation by doming associated with a number of high angle normal faults.

younging through time. The subducted oceanic slab bends rather old portion of the oceanic plate (older than 100 Ma) together with 4
suddenly at 40 km depth of the Benioff plane, which would remnant island arcs are being subducted (Seno and Maruyama, 1984;
considerably promote the wedge to extrude upward as shown in a Maruyama et al., 2009). On the other hand, in the central Honshu and
present day cross section in Fig. 19. Kanto regions, the oceanic lithospheric is older than 48 Ma as deduced
History of plate subduction in this region supports an approaching from the age of boninite in the Ogasawara for arc (Seno and
ridge through time to remove the wedge to the surface. In the case of Maruyama, 1984). The Philippine Sea plate subducts at the interme-
the well-studied Pacific-type blueschist belts of Sambagawa belt in diate region from Kanto to Kyushu and the Shikoku Palece–Vera Basin
Japan, the emplacement of deep-seated eclogite blueschist belt to the has an age of 8 to 30 Ma. The youngest part is the region marked by
mid-crustal level occurred at 70 to 80 Ma which is almost identical to Kinan Sea Mount chain, which is being subducted just under the Kii
the ridge subduction of the Kula-Pacific. The transport to the surface Peninsula. Kii Peninsula is the region where most remarkable uplift is
occurred afterwards, at about 50 Ma. The example of the Sambagawa ongoing since the Quaternary. Therefore, the central part of the Kii
indicates that the transport to shallow surface level was after the Peninsula has been severely eroded exposing the structural bottom
event of ridge subduction. The Juan de Fuca plate has only a 20 to under the Sambagawa belt. Thus, the Shimanto belt is in direct contact
30 km thick lithosphere because of the very young age. Therefore at with the Median Tectonic Line to the north, with the Sambagawa belt
depths of 25 to 50 km, a number of small earthquakes occur within missing in the Central Kii Peninsula. An evaluation of the uplift rate in
the slab mantle. These earthquakes are derived by dehydration SW Japan shows that the eastern and western parts were less eroded
reaction, similar to the well-studied case in Japan (Hasegawa et al., and the central part of the Kii has been most uplifted. Therefore,
2009; Maruyama et al., 2009; Omori et al., 2009). The regional apparently, the age of subducting plate is critical; if it is young, the
metamorphic belts emplaced at mid-crustal level under the forearc hanging wall seems to be selectively uplifted and eroded.
region could have been infiltrated by abundant dehydrated fluid — In the case of NE Japan, the Cretaceous Pacific plate is being
both deep subducted sediments and from the downgoing slab. subducted under the arc. The E–W cross section of NE Japan arc at
We now examine the case of SW Japan. The age of the Philippine Sendai brings out subducted sediments from high-resolution of
Sea Plate subducting underneath SW Japan arc is highly variable seismic tomography made by Hasegawa et al. (2009). The subducted
depending on the place (Fig. 20). For example, under Kyushu, a very sediments are present below the Benioff plane at 40–80 km depth,
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 125

Fig. 19. Ongoing Pacific-type orogeny in the Cascade region in the western North America (partly modified by Brandon and Calderwood, 1990). (a) Tectonic environment of the
Cascade region showing the distribution of subduction zones and subducting young Juan de Fuca plate. Note the location of the youngest lawsonite-bearing schist at Mt. Olympus. (b)
E–W cross section showing the uplift of Mt. Olympus which is cut on the top by normal fault and on the bottom by reverse fault. Extensive hydration could be present underneath
shown by the nest of frequent seismicity from the underlying Juan de Fuca plate.

together with hanging wall of already-accreted sediments. The metamorphic belts are being extruded and emplaced at mid-crustal
hanging wall of metamorphosed unit under BS–EC conditions may level. These units absorb the dehydrated fluid released from
be promoted to uplift along the Benioff plane carrying fragments of underneath to produce hydrous minerals. On the contrary, in the
mantle wedge by anticlockwise convection in the wedge corner intermediate regions such as Kii Peninsula and the central part of
triangle (Fig. 21). If this process is ongoing, the exhumation of UHP– Shikoku, hydration completed to extract excess fluids above the depth,
HP belt may occur in the case of old-slab subduction. with the second stage of doming to form high angle secondary normal
In the area where younger plates are being subducted in SW Japan, faults. In future, the youngest blueschist belt may appear in both Kii
local differences of uplift are also observed. For example, along the and Shikoku if this interpretation is correct. We compare two cross
strike of orogenic belts, topographic depression appears periodically sections across eastern Shikoku with no tremors, and central Kii
from east to the west. The Ise Bay, Kii Strait and Bungo Strait were Peninsula with frequent tremors (Fig. 22b). Along the eastern Shikoku
tectonically subsided whereas the intermediate region in Kii Penin- section, a topographic anomaly is seen in the forearc called Tosabae.
sula and Shikoku were elevated. The origin of this topographic anomaly (high) close to the sea level, has
Recently, microearthquakes termed as tremors have been precisely been speculated to be due to the presence of subducted seamount
recorded by the recently established seismic network in Japan. This has underneath (Kodaira et al., 2000). An alternate interpretation is that
helped to carefully monitor the fluid derived from Moho depths and to the subducted accretionary complex was tectonically extruded after
trace the ‘day-to-day’ movement of fluids (Obara, 2002; 2009). The blueschist facies metamorphism up to mid-crustal level to push up the
results show two regions where no tremors were observed beneath foregoing portion, which is now close to the sea level (Fig. 22). The Kii
the Kii Strait and the Ise Bay (Fig. 22a). In spite of the fact that Peninsula and Central Shikoku are now underlain by second stage of
subduction-related dehydration occurs commonly underneath shown doming as supported by the tectonic uplift above 350 m for the last
by slab-seismicity, the fluid has not come out above the Moho depth. 0.17 Ma at Muroto cape (Yoshikawa et al., 1964). Average uplift rate is
The following is a speculation to explain this enigma. Above the Benioff 2 mm/year. The sharp V-shaped notch at the south of Muroto
plane in these two regions, high-pressure–low-temperature regional Peninsula probably was formed by high angle normal fault along the
126 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

Fig. 20. (a) Index map showing the age variation of Philippine Sea plate. Western Philippine Sea on the north is as old as 100–40 Ma. The eastern margin, which subducts under the
Kanto region, is older than 48 Ma. The central part of Philippine Sea that subducts underneath SW Japan is as young as 30–17 Ma. (b) Cross section along A–A’ (modified after Park
et al., 2002). Regional metamorphic belt seems to be exhumed along the splay faults, which may explain the recent remarkable uplift in the Kii Peninsula. Also shown is the top
boundary of subducting Philippine Sea plate. The presence of cold seep indicates low-temperature springs enriched in nutrients. Cold seep is the surface expression of splay fault, and
the abundant nutrients feed biological communities.
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 127

Fig. 21. A schematic illustration of 3 domains of metamorphic–metasomatic factory (MMF), subduction zone magma factory (SZMF) and big mantle wedge (BMW) under NE Japan
across Sendai (Maruyama et al., 2009, based on seismic tomography work by Hasegawa et al., 2009). The double seismic planes (hatched) cause a large pulse of dehydrated fluid from
the descending Pacific-slab leading to viscosity contrast among MMF, SZMF and BMW. Counter-flow convection in the mantle wedge begins immediately above 200 km depth and
moves toward a volcanic front obliquely to release an arc magma under the volcanic front. The boundary between MMF and SZMF corresponds to the culmination point of frequent
seismicity at 60 km depth. The enlarged figure shows a contrasting convection flow for MMF and SZMF, the former promoting exhumation of subducted sediments as shown by
yellow color including fragments of wedge mantle peridotites (Alpine peridotites).

coastal lines, supporting the idea of domal uplift. Such domal uplift cut 15 million years. From the time of initiation of metamorphism to
by high angle normal faults suggests the underplating of a buoyant mountain building, it took a total of 40 million years. Subsequently
mass of accretionary complex below to push up the above. The the plate boundary shifted leading to the rise of the Himalayan
buoyant mass could be younger spreading ridge, or rising large volume mountains since 1 Ma ago. Mountain building is still going on at
of quartzofeldspathic accretionary complex. A similar process seems to present (Fig. 23).
be ongoing under the Kii Peninsula. Thus, SW Japan might expose the In the case of Pacific, taking the Sambagawa belt as an example,
youngest blueschist belt in future. subduction and underplating of accretionary complex from trench up
to 60–70 km depth took only 1 million year, based on the computa-
10. Relationship to orogeny tions of relative plate motion during the Cretaceous time (Engebretson
et al., 1985; Maruyama and Seno, 1986). Regional metamorphism has
Regional metamorphic belts occupy the orogenic core and are its peak at 120 to 125 Ma as demonstrated by zircon SHRIMP age
hence the most important element of an orogen. The process to (Okamoto et al., 2004). Exhumation and emplacement at mid-crustal
produce regional metamorphic belts is complex including generation level occurred between 80 and 60 Ma, where extensive hydration
of the metamorphic belt, subsequent exhumation, emplacement at occurred. Therefore, it took 40 to 45 million years from mantle depths
mid-crustal levels, and finally mountain building which exposes the to the mid-crustal level. Thereafter, dome-like second step mountain
regional metamorphic units to the surface. Large-scale orogenic belts building began to expose the high-grade regional metamorphic rocks
regionally affect the already existing continental crust causing to the surface by 50 Ma (Isozaki and Maruyama, 1990). From the
widespread sedimentation in the surrounding regions, accompanied second stage mountain building, the uplift rate is estimated as 4 mm/
by arc magmatism to increase the volume of continental crust. Thus, year, converting 10 to 12 km uplift during 20 to 30 million years. Thus,
regional metamorphism is one of the most important processes in the Sambagawa orogeny has taken 70 million years from its beginning
orogeny. to the surface exposure of regional metamorphic rocks.
Orogeny is not an instantaneous phenomenon, but a process that Pacific-type orogeny in the Sambagawa metamorphic belt during
lasts over a long time of the order of a hundred million years. In the the Cretaceous is yet another process. In this case, the formation of the
case of Himalayas, the subduction started at 50 Ma. The beginning of folded belts may have occurred not only at late stage but also at the
UHP to HP metamorphism started at 48 Ma (Kaneko et al., 2003), initial stage. To record the earliest Cretaceous event, information from
which is very close to the timing of subduction initiation, and it took Asian continents is critical. In the Japanese Islands, information is
23 million years from UHP metamorphism at mantle depths to the limited to the presence of large batholith belts. The northern limit of
mid-crustal level exhumation at 25 Ma where extensive hydration– the batholith belt is missing which is now observed in China. As will be
recrystallization occurred. Furthermore, dome like uplift and moun- discussed later, in S. America and N. America the initial stage of
tain building started only after 8 Ma with a large time gap reaching Cretaceous orogeny is marked by a series of thrust belts developed on
128 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 129

Fig. 23. Tectonic evolution of the Himalayan region from 60 Ma to the present (modified after Kaneko, 1997). Before the Indian collision, Pacific-type orogeny prevailed to form paired belts,
Cretaceous huge batholith belt on the north associated with contemporaneous BS belt on the south. The Indian continent subducted underneath the southern margin of Eurasia at 50 Ma, the
regional metamorphism started from 48 Ma and proceeded to progressive HP–UHP metamorphism, followed by slab break-off to trigger the exhumation of UHP–HP belt into mid-crustal depth
by 25 Ma by wedge extrusion process. Asthenospheric upwelling by delaminated slab heated the bottom of thickened continental crust under Tibet, caused eclogite melting at the Moho depth
and erupted adakite lava-flows. Also shown is the top boundary of the UHP–HP unit called STDF (South Tibetan Detachment Fault). The bottom boundary is MCT (Main Central Thrust). An
extensive hydration occurred at 25 Ma at mid-crustal level. Second stage doming started around 8 Ma to cause mountain building to transport huge amounts of sediments to the Indian Ocean.

the northernmost foothill of the mountain belts consisting a major may have had two stages of mountain building at the initial stage and
portion of the batholith belts. If this is the case, Pacific-type orogeny at the final stage.

Fig. 22. (a) Distribution of tremors at the Moho depths in SW Japan (Obara, 2002). Note the absence of tremor in Kii Strait and easternmost part of Shikoku and in the small region
near Ise Bay. Also shown is the isodepth contour of the top of the Philippine Sea plate. (b) Schematic cross section along A–B and C–D. A–B is the cross section for the tremor-absent
region whereas C–D is the tremor present region. Absence of tremor can be explained by the dehydrated fluids were all absorbed to hydrate the exhumed and water unsaturated
regional metamorphic unit. Tremor is observed after the saturation of exhumed regional metamorphic belt shown along C–D cross section.
130 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

11. Himalayas and Andes: a comparison thrust. Therefore, there is no manifestation on the surface in terms of
vertical movements.
In the Himalayas, a 50 million years long history exists prior to the There is an essential difference between collision-type and Pacific-
final collision-type orogeny (Fig. 23). At present, within the Indian type orogens (Fig. 25). In the case of collision reflecting the different
Ocean near the spreading ridge, prominent seismicity records mechanism of uplift, only the pressure drops at near-constant
compressional stress. Therefore, the plate boundary is delineated temperature at temperature after the metamorphic pressure reaches
not along Main Frontal Thrust in the Himalayas, but within the Indian the peak. On the other hand, both pressure and temperature decrease
Ocean (Engdahl et al., 1998), which corresponds to the northern together in the case of Pacific-type. This may be due to the fact that
boundary of the Indo-Australian plate. If the plate boundary shifts oceanic plate subduction was simultaneously ongoing when the HP–
completely from MFT to Indian Ocean, the Himalayan Mountain UHP units are being removed tectonically to the mid-crust, although
Range would not be a mobile orogenic belt, which means the end of need to verify in future (Fig. 25).
collision orogeny.
On the other hand, in the central part of the Andean Mountain Belt,
mountain building has definitely begun since Miocene at about 25 Ma 12. Concept of paired metamorphic belts
(Ramos and Aleman, 2000). This was caused by continent subduction
of South America underneath the Andean mountain range known as A- In 1961, Miyashiro proposed metamorphic facies series and paired
type subduction to jack-up the Andean mountain. This process metamorphism based on the geology of Japan. These two concepts
corresponds to the mountain building at the initial stage of Pacific- helped the establishment of plate tectonics in 1968. However, the
type orogeny (Fig. 24). However, regional metamorphic belt has not concept of paired metamorphic belts has been considerably modified
yet been emplaced into the core of the Andean mountain, and since then (Santosh and Kusky, 2010; see also Brown, 2010-this
extensive scale of felsic volcano-plutonism has not yet occurred. issue). For example, collision-type regional metamorphic belts are not
Moreover, subduction zone volcanism itself is absent in about one paired at all. No counterpart of low P high T metamorphic belt is
third of the entire Andean region (Gutscher et al., 1999). These present. Also, granitic batholith belt is absent in general. In the case of
observations suggest that the Andean mountain range is underlain by a Alps and Himalayas, minor post-orogenic granites intruded at the end
mountain building stage, but has not yet reached the activated time of the orogeny. It should be noted that oceanic plate subduction is a
dominated by extensive volcano-plutonism and exhumation of necessary phenomenon prior to the final collision of continents in the
regional metamorphic belts in the forearc region. Presumably, the case of collision-type orogeny. Therefore, Pacific-type accretionary
Andean belt is still in the initial stage over the entire 70 million years complex must have been formed to the north of collision type orogeny
long Pacific-type orogeny. However, Andean mountain range is wide in the case of Himalayas. In this case, proportional to the total length
enough so that a regional provincialism might be present. For example, of subducted oceanic plate, size of accretionary complex and size of
at the southern end of the Andean belt, mid-oceanic ridge began to batholith belts must vary. To the north of the Himalayan Mountain
subduct underneath where TTG granite intruded near the trench only Range, a huge batholith belt and small glaucophane-bearing meta-
50 km from the trench. Therefore, this local area may have been in the morphic belts are observed both of which are ca. 100 Ma old and
stage of a peak Pacific-type orogeny (Fig. 24c). But the regional meta- formed prior to the Himalayan collision (Maruyama et al., 1996).
morphic belt has not yet been exposed on the surface. Comparing the These paired rocks would belong to Pacific-type orogen, and not to the
Sambagawa orogeny, the stage at the southern end of the Andean collision type. On the other hand, there is no batholith belt in the Alps.
belt is the stage before the doming uplift and after the exhumation of In the case of collision type orogen, there is no heat source to produce
regional metamorphic belts at mid-crustal level. Nevertheless, the high-T low-P type metamorphic belts.
overall stage of the whole Andean belt has not yet reached the second In the case of Pacific-type orogenic belts, the presence of low-P and
mountain building stage. In future, a few mid-oceanic ridges would high-T regional metamorphic belts is rather exceptional. If present,
reach to the trench to promote the orogenic activity (Fig. 24c). both belts were not exhumed to the surface at the same period. For
example, in the Sambagawa–Ryoke paired metamorphic belt,
regarded as the world standard, the Ryoke belt runs side by side
11.1. Mountain building is unrelated to the exhumation of HP–UHP belt with the Median Tectonic Line (MTL) on the map view, with Ryoke in
the north and Sambagawa in the south, in general. But in the three
In collision orogeny, formation of huge mountain belt is restricted dimensions, the Ryoke belt rests above the Sambagawa with a sub-
to the final stage. On the other hand, the Pacific-type is marked by two horizontal fault called as paleo-MTL (Isozaki and Maruyama, 1991).
stages of mountain building — the first stage is characterized by the The boundary fault was formed at 15 Ma when Japan Sea opened and
tectonic uplift of volcanic front with simultaneous HP–UHP meta- the upper Ryoke belt was thrust over the Shimanto–Sambagawa and
morphism by A-subduction from the backside down to 60 to 100 km Jurassic accretionary complex from north to southward. Thereafter, in
depth. The second mountain building occurs at the end of orogeny as Kii Peninsula and Shikoku, particularly the regions towards the
discussed before. eastern part were preferentially uplifted to erode out the upper Ryoke
Regardless of collision-type or Pacific-type, the mountain building belt. As a result, the apparent distribution of the Sambagawa to the
is unrelated to exhumation of regional metamorphic belts. Mountain south and Ryoke to the north appear to give an impression of paired
building to expose the orogenic core occurs by the domal uplift of the belts. Yet, on the western margin of Shikoku, whole Kyushu and Kanto
sandwiched units with the core of regional metamorphic belts at the plain, top and bottom relationship of the Ryoke and Sambagawa are
end of orogeny. However, at this period, tectonically eroded portion preserved very well. The apparent arrangement of paired belt is also
reaches only 10 km. Formation of folded mountain accompanied by recognized in New Zealand and Celebes. Hida, Sangun and Hidaka are
active erosion generates the remarkable topography of mountain also exceptional examples of paired belts over the world. In the case of
belts, but corresponding to only about 3 kbar change in metamorphic Pacific-type orogenic belts, huge TTG belts on the continent side are
pressure. It is important to note the extremely large pressure gradient paired with low-temperature high-pressure type regional metamor-
from 25–70 kbar down to 3–5 kbar recorded in regional metamorphic phic belts oceanward. In this case, batholith belt is paired with high-
rocks, are absent in terms of vertical movement on the surface. This is pressure low-temperature regional metamorphic belts, and not with
due to extremely slow exhumation of regional metamorphic rocks low-pressure high-temperature metamorphic belts. Finally we con-
from mantle depths to mid-crustal level and because of the very thin clude that Miyashiro's concept of paired metamorphism is correct, but
sheet-like solid intrusion of the metamorphic belt along Benioff incorrect for the simultaneous uplift to expose paired belts.
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 131

Fig. 24. Tectonic map of Andean orogenic belt showing E–W cross section (A–B, C–D). Also shown is the distribution of thrust on the eastern foothill of the Andean mountain belts
(Ramos, 1999). Subducting mid-oceanic ridge is present only at the southern margin of the Andean mountain. As shown in A–B cross section (Ramos et al., 1996), a series of thrusts
push up the Andean belts by A-subduction of South American continent underneath the Andean belt. Regional metamorphic belts may have been formed along the southern margin
of Andean belt, but have not yet been exposed to the surface (shown along cross section C–D). Red triangles denote volcanoes.

13. Fluids and regional metamorphism perhaps played the most crucial role (e.g., Santosh et al., 2009b). The
zircon crystals formed during metamorphism include high-T low-P
The major cause of regional metamorphism has long been con- minerals such as magmatic origin at the core e.g., quartz, albite which
sidered as an effect of the increase of P and T. However, the evolution are mantled by the newly grown zircon during the increased P–T
and migration of fluids have been highlighted in recent studies and conditions. This in turn includes coesite, jadeite and diamond in the
132 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

100 to 200 km depths and witnessed temperatures of 1000 °C, some


igneous protoliths composed of anhydrous minerals preserve mag-
matic texture and minerals in spite of the UHP and UHT conditions
(Fig. 26). Some gabbroic rocks in HP–UHP belts in China and Alps
preserve olivine and plagioclase despite their conversion to eclogitic
assemblages, because of absence of fluids in gabbro.
Garnetite transforms from gabbro and is completely recrys-
tallized under UHP conditions. At mid-crustal levels of 10–15 km
depths and at temperature of 400–650 °C, these rocks recrystallize
to amphibolite through external fluid infiltration. If the garnetite
is large enough, the core of the unit would remain as garnetite
successively mantled by garnet amphibolite and amphibolite.
However, if fractures cut through, the central garnetite domain is
also is amphibolitized. Similar examples are also seen in the case
of quartzofeldspathic gneisses surrounding eclogite where ex-
treme ductile deformation and completely hydration obliterating
the UHP event can be seen. All of these observations clearly
indicate that fluids play a more critical role in controlling mineral
stability and ductility for quartzofeldspathic country rocks than P
and T increase.

14. Earthquake at subduction zones

The discovery of UHP metamorphic rocks and their petrologic and


tectonic characteristics have revolutionized the concepts of regional
metamorphism, particularly focusing on the evolution and migration
of fluids. The link between earthquakes and metamorphism in con-
suming plate boundaries became increasingly apparent. In these
zones, subducted hydrated slab dehydrates with increasing pressure
and temperature and the fluid liberated would remarkably influence
the rheological properties and stress fields (Peacock, 1993, Omori et
al., 2002, 2009; Hacker et al., 2003 etc.). In other words, the birth of
fluid and its transportation translates to triggering of earthquake,
based on the concept of dehydration embrittlement (e.g., Meade and
Jeanloz, 1991; Omori et al., 2009). Metamorphic petrologists can
study the fossil evidence for earthquake within regional metamorphic
Fig. 25. Schematic diagram to show the slightly different exhumation processes belts and evaluate the dynamics of orogenic movement. Subduction
between Pacific- and collision-types. The fundamental process of two-step exhumation zones along which frequent earthquakes occur are in fact the location
is identical for both, but P–T time path is systematically different. There is much deeper of ongoing progressive metamorphic reactions (Maruyama et al.,
subduction in the case of continent-collision compared to the Pacific-type. The nearly
2004; Omori et al., 2009).
isothermal decompression for the collision type is slightly different from the Pacific-
type. The difference may be explained by subduction and refrigeration in the case of
Pacific-type.
15. Fluid circulation characterizes the plate boundary

Distribution of earthquakes is confined only to the plate


case of UHP rocks. The outermost margin includes low-P minerals boundary (Fig. 27a). Fluid circulation is restricted only to the
again such as plagioclase. This observation demonstrates that the plate margin of consuming plate boundary from which fluid moves
quartz and feldspar at the core were not converted to the high- deep into the earth's interior at least covering the whole upper
pressure and high-temperature minerals even though extremely mantle down to 660 km depth. In the divergent boundary, fluids
high-P and high-T conditions prevailed over geologic time. Moreover, come up from mid ocean ridge magma, which in turn originated at
UHT and UHP minerals included within zircon crystal have not 410 km through a curtain-like mantle upwelling (Zhao, 2004).
changed to low-P and low-T minerals even after their transport to Fluids may have been derived from the Earth's interior to the
shallow level. Shielding of those minerals within robust minerals like surface through superplumes (see Maruyama et al., 2009; Santosh
zircons prevented their phase transformation in spite of the drastic et al., 2009b). This observation indicates that regional metamor-
change of P–T conditions over geologic time. phism occurs only along the plate boundary where earthquakes
On the other hand, matrix minerals in UHP rocks have almost all occur. For the other regions such as under the continents and
completely transformed to low-pressure minerals. All of these within ocean plate, no free fluids are available except hotspots and
observations support the critical role of fluids in mineral phase hence no recrystallization proceeds. Even under the continents in
transformation. Absence of fluids was the major cause for the absence the lower crust or below the Moho, even if temperature conditions
of phase transformations of the inclusions within zircon crystals. A exceed 700 °C, no metamorphic recrystallization must prevail if
similar phenomenon has also been observed in the case of mineral fluid is absent. A contrasting picture emerges in the distribution of
inclusions in garnet and omphacite. These minerals appear to be earthquakes between mid-oceanic ridge and continental rift. In the
shielded and protected from the fluid flow along grain boundaries or former case, earthquakes are present down to 660 km depth, but in
cleavages and cracks within UHP rocks. the latter, only at depths shallower than 30 km along the transform
We will now examine the outcrop scale of fluid infiltration in UHP fault boundary, earthquakes are generated at depths shallower
terranes. In the case of UHP rocks, which were taken down to depths of than 15 km. These contrasting differences in depths correspond to
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 133

Fig. 26. Partially eclogitized metagabbro, which preserves igneous mineralogy and textures which survived UHP metamorphism (after Mork, 1985). (a) Sketch of the domain in
which igneous mineralogy and texture are relatively well preserved. (b) Sketch of more advanced stage of metamorphic recrystallization. (c) Schematic diagram to show the gradual
change from gabbro to eclogite in mineralogy and texture with progressive metamorphism and eclogitization.

the depths of circulating fluids. The physical differences of fluid ages remain as before. If this is true, the orogenic peridotites may retain
circulation depend on the plate boundary. old mantle convection history even at CMB level, preserved in within
mineral inclusions of zircons.

16. Metamorphic recrystallization in the whole mantle


17. Subduction zone geotherm through geologic history and the
If the circulating fluid is the only factor to cause regional meta- shrinking forbidden zone
morphism as documented clearly in UHP–HP metamorphic rocks and
zoned zircon which contains quartz, graphite, albite inclusions in the The subduction zone geotherm must have changed through time,
core and rim, with mantle including diamond and coesite and/or according to the cooling Earth since its birth at 4.6 Ga in the planetary
jadeite, the site of ongoing recrystallization in the whole mantle scale space. This issue has long been considered by several authors based on
could be envisaged as shown in Fig. 27. The cross section of the Earth the P–T estimates of regional metamorphic belts over the world, either
is schematically shown cross-cutting Pacific and African superplume Pacific-type or collision-type. Here we summarize the concept based
together with the western Pacific triangular zone with double-sided on the results presented in Maruyama et al. (1996) and Maruyama and
subduction zones, to make it clear for circulating fluids to recrystallize Liou (2005).
the mantle and related rocks. According to the double-sided Since the pioneering work by de Roever (1957), several authors
subduction, the upper mantle in the western Pacific triangular zone have tried to confirm the geothermal gradients at subduction zone
is being recrystallized most predominantly among whole upper through time, and all of them recognized a clear trend of secular
mantle region, and by the same reason the lowermost mantle could variation with time (Fig. 28). The maximum T limit for regional
preserve the lowest temperature of ca. 2300 °C at CMB (core–mantle metamorphism has been less than 900–1000 °C throughout Earth
boundary) (Maruyama et al., 2007). history. However, the maximum-P condition increased from 2.0 GPa
In addition, the two sites of rising superplume could bring recrys- before 1000 Ma to N6.0 GPa after 630 Ma. Most metamorphic terranes
tallization from the bottom to the top by the presence of fluids derived older than 1000 Ma show maximum-P conditions less than 1.0 GPa,
from liquid outer core, because of the presence of light elements such indicating that Precambrian metamorphism occurred within crustal
as C, H, O, and S, which occupy ca. 10% of the outer core, and are depths and slab-melting to produce TTG (tonalite–trondhjemite–
oversaturated in the growing solid inner core (Maruyama, 1994; granodiorite) rocks that dominate the continents (Martin, 1993). This
Maruyama et al., 2007). Below the mid-oceanic ridge down to ca.30– is because a wide P–T field of slab-melting lies above the wet solidus of
40 km depth, rising mantle flow would release magmatic melts to hydrated MORB (Fig. 28).
cause the recrystallization to reset the isotope equilibration. All recognized UHP belts are younger than 630 Ma, and the
On the other hand, the fluid-absent zone such as the bottom of subduction zone metamorphism to form UHP belt became possible
cratonic continents would not be recrystallized even at T over 700 °C, only after the temperatures became cold enough to inhibit slab-
nor will there be any isotopic re-equilibration. Hence the radiometric melting (Fig. 28). It should be pointed out that the subduction zone
134 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

Fig. 27. a): Global distribution of earthquakes (shown by dots) and plate boundaries (modified after Seno, 1995). The distribution of earthquakes shows an excellent correlation with
the plate boundary such as mid-oceanic ridges, continental rifts and transform faults. The upper mantle beneath the western Pacific shown as shaded region is dominated by water
supplied by double-sided subduction zones from the east and from the south (see Maruyama et al., 2007). b): The cross section of the Earth along the solid curve in a), indicating
ongoing recrystallization of mantle by fluid circulation (blue color). Recrystallization, which corresponds to ongoing regional metamorphism, may be highly restricted as shown by
the cross section. Even in the deep mantle recrystallization may be possible only by fluid migration from the liquid outer core. See more details in the text.

geotherm is nearly parallel to the wet solidus of MORB (Maruyama forbidden zone in P–T space at low-T and high-P side has reduced
and Liou, 1998). This probably caused the sudden appearance of UHP– through time. The region was wide enough to cover a major part of P–T
HP belts after 630 Ma. space below MORB wet solidus before 1000 Ma, but got rapidly reduced
Following the secular change of subduction zone geotherm, as towards the low-T side after 630 Ma. This change allowed lawsonite
recorded in regional metamorphic rocks shown in Fig. 28, the eclogite to appear first since 520 Ma along the eastern margin of
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 135

Unfortunately the Japanese scientific community lost him from Japan in


1965 as he spent the latter half of his academic years in U.S.A, developing
his field to cover whole Earth, particularly ocean-floor petrology. We
Japanese would like to appreciate his highly intelligent and thoughtful
works written in Japanese first before the articles in English were
published. All of these contributions educated Japanese, and developed
understanding of the consuming plate boundary around Japan and the
whole Earth.
We would like to thank Prof. M. Santosh for his valuable help
during the preparation and revision of this manuscript. We also thank
an anonymous referee for thoughtful comments.

References
Aoki, K., Iizuka, T., Hirata, T., Maruyama, S., Terabayashi, M., 2007. Tectonic boundary
between the Sanbagawa belt and the Shimanto belt in central Shikoku, Japan.
Journal of the Geological Society of Japan 113, 171–183.
Aoki, K., Itaya, T., Shibuya, T., Masago, H., Kon, Y., Terabayashi, M., Kaneko, Y., Kawai, T.,
Maruyama, S., 2008. The youngest blueschist belt in SW Japan: implication for the
exhumation process of the Cretaceous Sanbagawa high-P/T metamorphic belt.
Journal of Metamorphic Geology 26, 583–602.
Aoki, K., Kitajima, K., Masago, H., Terabayashi, M., Omori, S., Yokoyama, T., Takahata, N.,
Sano, Y., Maruyama, S., 2009. Metamorphic P–T–t history of the Sanbagawa belt in
central Shikoku, Japan and implications for retrograde metamorphism during
exhumation. Lithos 113, 393–407.
Aoki, K., 2009. Progressive and retrogressive metamorphism in subduction zone
revisited: Metamorphic history of the Sanbagawa high-P/T metamorphic belt in
Japan. Ph. D. Thesis, Tokyo Institute of Technology.
Aoya, M., Wallis, S., 1999. Structural and microstructural constraints on the mechanism of
eclogite formation in the Sambagawa belt, SW Japan. Journal of Structural Geology 21,
1561–1573.
Aoya, M., Uehara, S., Matsumoto, M., Wallis, S., Enami, M., 2003. Subduction stage
pressure–temperature path of eclogite from the Sambagawa belt: Prophetic record
for oceanic-ridge subduction. Geology 31, 1045–1048.
Fig. 28. P–T diagram showing the decrease of subduction zone geotherm through Banno, S., Sakai, C., 1989. Geology and metamorphic evolution of the Sambagawa belt,
geologic time (modified after Maruyama and Liou, 2005). The summary of P–T Japan. In: Day, J.F., Cliff, R.A., Yardley, B.W.D. (Eds.), Evolution of Metamorphic
conditions are estimated from about 300 metamorphic belts over the world through Belts: Geological Society of London Special Publication, 43, pp. 519–532.
time back to the Archean (ca. 4.3 Ga) (summarized after Maruyama and Liou, 2005). Banno, S., Higashiono, T., Otsuki, M., Itaya, T., Nakajima, T., 1978. Thermal structure of
the Sanbagawa metamorphic belt in central Shikoku. Journal of the Physics of the
Subduction zone is the coldest region in the mantle at anytime in the geologic past. The
Earth 26, 345–356.
coldest region shown in the P–T space has reduced towards the lower-T side through
Barrow, G., 1893. On an intrusion of muscovite–biotite gneiss in the South-Eastern
time, as shown by the shrinkage of the forbidden zone. The Archean-Proterozoic highlands of Scotland, and its accompanying metamorphism. Quaternary Journal of
forbidden zone was widest with an extensive slab-melting. The reference frames of P–T Geological Society of London 49, 330–358.
grids such as coesite-quartz, diamond-graphite and jadeite + quartz = low-albite or Brandon, M.T., Calderwood, A.R., 1990. High-pressure metamorphism and uplift of the
high-albite, MORB solidus under dry or wet condition, and metamorphic facies grids are Olympic subduction complex. Geology 18, 1252–1555.
after Maruyama et al. (1996). Brandon, M.T., Cowan, D.S., Vance, J.A., 1988. The Late Cretaceous San Juan thrust system,
San Juan Island, Washington. Geological Society of America Special Publication 21,
81.
Brown, M., 2010. Paired metamorphic belts revisited. Gondwana. Gondwana Research
Australia (Watanabe et al., 1997). Another example is from Guatemala 18, 46–59 (this issue).
(Tsujimori et al., 2006). Caby, R., 1994. Precambrian coesite from northern Mali; first record and implications
for plate tectonics in the trans-Saharan segment of the Pan-African belt. European
Journal of Mineralogy 6, 235–244.
18. Postscript Campos Neto, M.C., 2000. Orogenic systems from southwestern Gondwana. In: Cordani,
U.G., Milany, E.J., Thomaz Filho, A., Campos, D.A. (Eds.), Tectonic Evolution of South
America: 31st International Geological Congress. , pp. 335–365.
It has already passed nearly 30 years since coesite-bearing UHP Charlton, T.R., Barber, A.J., Barkham, S.T., 1991. The structural evolution of the Timor
rocks were first reported from collisional belts in Europe, but several collision complex, Eastern Indonesia. Journal of Structural Geology 13, 489–500.
debates continue. The UHP belts constitute only a very minor Chemenda, A.I., Burg, J.P., Mattauer, M., 2000. Evolutionary model of the Himalaya–
Tibet system; Geopoem based on new modeling, geological and geophysical data.
component among the various metamorphic belts of the world. Earth and Planetary Science Letters 174, 397–409.
However, this trivial “noise” has radically revised many of the Chopin, C., 1984. Coesite and pure pyrope in high grade blueschists of the western Alps:
concepts and framework of metamorphic petrology established over a first record and some consequences. Contributions to Mineralogy and Petrology
86, 107–118.
the last 100 years and expanded the realm of metamorphic petrology Copeland, P., Harrison, T.M., Hodges, K.V., Maruejol, P., Le Fort, P., Pecher, A.,., 1991. An
to include many subjects including deep Earth geophysics. One of the early thermal disturbance of the Main Central Trust, central Nepal. Implications for
major advancements in metamorphic petrology was the understand- Himalayan Tectonics. Journal of Geophysical Research 96-B5, 8475–8500.
Coward, M.P., Dietrich, D., 1989. Alpine tectonics— an overview. In: Coward, M.P., Dietrich, D.,
ing of the role fluids and the critical role they play in mineral
Park, R.G. (Eds.), Alpine Tectonics: Geological Society Special Publication, 45, pp. 1–29.
transformation. Indeed, crustal rocks subjected to extreme metamor- de Roever, W.P., Nijhuis, H.J., 1963. Plurifacial metamorphism in the eastern Betic
phic conditions were the key for this paradigm shift, with newer Cordilleras (SE Spain), with special references to the genesis of glaucophane.
Geologiosche Rundschau 53, 324–336.
information on ongoing processes in active plate boundaries adding
de Roever, W.P., 1957. Sind die alpinotypen peridotit-massen vielleicht tectonisch
more fuel to the debate. verfrachtete bruchstucke der peridotitschale? Geologiosche Rundschau 46, 137–146.
Defant, M.J., Drummond, M.S., 1990. Derivation of some modern magmas by melting of
young subducted lithosphere. Nature 347, 662–665.
Acknowledgments Derry, L.A., France-Lanord, C., 1996. Neogene Himalayan weathering history and river
87
Sr/86Sr: impact on the marine Sr record. Earth and Planetary Science Letters 142,
Prof. A. Miyashiro was a pioneering worker who systematized 59–74.
Drummond, M.S., Defant, M.J., 1990. A model for trondhjemite–tonalite–dacite genesis
metamorphic petrology, combined with tectonic settings, back in late and crustal growth via slab melting — Archean to modern comparisons. Journal of
1950s. He was born in Japan, and contributed to Earth Science in Japan. Geophysical Research 95, 21503–21521.
136 S. Maruyama et al. / Gondwana Research 18 (2010) 106–137

Enami, M., 1982. Oligoclase–biotite zone of the Bessi district of the Sanbagawa belt, Katayama, I., Maruyama, S., Parkinson, C.D., Terada, K., Sano, Y., 2001. Ion micro-probe
central Shikoku. Journal Geological Society of Japan 88, 887–900 (in Japanese with U–Pb zircon geochronology of peak and retrograde stages of ultrahigh-pressure
English abstract). metamorphic rocks from the Kokchetav massif, northern Kazakhstan. Earth and
Engdahl, E.R., van der Hilst, R., Buland, R., 1998. Global teleseismic earthquake Planetary Science Letters 188, 185–198.
relocation with improved travel times and procedures for depth determination. Kodaira, S., Takahashi, N., Nakanishi, A., Miura, S., Kaneda, Y., 2000. Subducted seamount
Seismological Society of America Bulletin 88, 722–743. imaged in the rupture zone of the 1946 Nankaido Earthquake. Science 289, 104–106.
Engebretson, D.C., Cox, A., Gordon, R.G., 1985. Relative plate motion between oceanic Komiya, T., Hayashi, M., Maruyama, S., Yurimoto, H., 2002. Intermediate-P over T type
and continental plates in the Pacific basin. Geological Society of America Special Archean metamorphism of the Isua suplacrustal belt: implications for secular
Paper, 206. 59 pp. change of geothermal gradients at subdction zones and for Archean plate tectonics.
England, P.C., Thompson, J.B., 1984. Pressure–temperature–time paths of regional meta- American Journal of Science 302, 804–826.
morphism. I. Heat transfer during the evolution of regions of thickened continental Komiya, T., Maruyama, S., Hirata, T., Yurimoto, T., Nohda, S., 2004. Geochemistrry of the
crust. Journal of Petrology 25, 894–928. oldest MORB and OIB in the Isua supracrustal belt, SouthernWest Greenland:
Fettes, D.J., 1979. The metamorphic map of the British and Irish Caledonides. In: Harris, implication and temperature of early Archean upper mantle. The Island arc 13, 47–72.
A.L., et al. (Ed.), The Caledonides in the British Isles — Reviewed. : Geological Society Lefort, P., 1975. Himalaya, the collided range, present knowledge of the continental arc.
of London Special Publication, 8. Scottish Academic Press, Edinburgh, pp. 307–321. American Journal of Science 75, 1–44.
Ganguly, J., Cheng, W., Chakraborty, S., 1998. Cation diffusion in alminosilicate garnets: Liou, J.G., Zhang, R.Y., Jahn, B.M., 2000. Petrological and geochemical characteristics of the
experimental determination in pyrope–almandine couples. Contributions to Mineralogy ultrahigh-pressure metamorphic rocks from the Dabie–Sulu terrane. International
and Petrology 131, 171–180. Geology Review 42, 328–352.
Gansser, A., 1964. Geology of the Himalayas. Willey, London. 289pp. Liou, J.G., Zhang, R.Y., Katayama, I., Maruyama, S., 2002. Global distribution and
Guillot, S., Hodges, K., Le Fort, P., Pecher, A., 1994. New constraints on the age of the petrotectonic characterizations of UHPM terranes. In: Parkinson, C.D., Katayama, I.,
Manaslu leucogranite: evidence for episodic tectonic denudation in the central Liou, J.G., Maruyama, S. (Eds.), Diamond-bearing from Kokchetav Massif, Kazakhstan.
Himalayas. Geology 22, 559–562. Universal Academy Press, Tokyo, pp. 15–35.
Gutscher, M.A., Olivet, J.L., Aslanian, D., Eissen, Z.P., Maury, R., 1999. The “Lost Inca Liu, F.L., Xu, Z.Q., Liou, J.G., Katayama, I., Masago, H., Maruyama, S., Yang, J.S., 2002.
Plateau” cause of flat subduction beneath Peru? Earth and Planetary Science Letters Ultrahigh-pressure mineral inclusions in zircons from gneissic core samples of the
171, 335–341. Chinese Continental Scientific Drilling Site in eastern China. European Journal of
Hacker, B.R., Peacock, S.M., Abers, G.A., Holloway, S.D., 2003. Subduction factory. 2. Are Mineralogy 14, 499–512.
intermediate-depth earthquakes in subducting slabs linked to metamorphic Martin, H., 1993. The mechanism of petrogenesis of the Archean continental crust. Lithos
dehydration reactions? Journal of Geophysical Research 108 (B1). doi:10.1029/2001 30, 373–388.
JB 00129. Maruyama, S., 1990. Exhumation mechanism of high-pressure metamorphic belt.
Hamilton, W., 1979, Tectonics of the Indonesian region. US Geol. Survey, Professional Abstract of Ann. Meeting of Geol. Soc. Japan, p. 484 (in Japanese).
Paper 1078, 345pp. Maruyama, S., 1994. Plume tectonics. Journal Geological Society of Japan 100, 24–49.
Harris, R., 2006. Rise and fall of the Eastern Great Indonesian arc recorded by the Maruyama, S., Liou, J.G., 1998. Initiation of UHP metamorphism and its significance on
assembly, dispersion and accretion of the Banda Terrane, Timor. Gondwana the Proterozoic/Phanerozoic boundary. The Island Arc 7, 6–35.
Research 10, 207–231. Maruyama, S., Liou, J.G., 2005. From snowball to Phanerozoic Earth. International
Hasegawa, A., Nakajima, J., Uchida, N., Okada, T., Zhao, D., Matsuzawa, T., Umino, N., 2009. Geology Review 47, 775–791.
Plate subduction, and generation of earthquakes and magmas in Japan as inferred Maruyama, S., Seno, T., 1986. Orogeny and relative plate motions: an example of the
from seismic observations: An overview. Gondwana Research 3–4, 370–400. Japanese Islands. Tectonophysics 127, 1–25.
Hashimoto, M., Tagiri, M., Kusakabe, K., Masuda, K., Yano, T., 1992. Geologic structure Maruyama, S., Cho, M., Liou, J.G., 1986. Experimental investigation of blueschist–
formed by tectonic stacking of sliced layers in the Sambagawa metamorphic greenschist transition equilibria: pressure dependence of Al2O3 contents in sodic
terrane, Kodama–Nagatoro area, Kanto Mountains. Journal of Geological Society of amphibole — a new geobarometer. In: Evans, B.W., Brown, E.H. (Eds.), Blueschists
Japan 98, 953–965 (in Japanese with English abstract). and Eclogites: Geological Society of America Memoir, 164, pp. 1–16.
Helgeson, H.C., Delany, J.M., Nesbitt, H.W., Bird, D.K., 1978. Summary and critique of the Maruyama, S., Liou, J.G., Zhang, R.Y., 1994. Tectonic evolution of the ultrahigh-pressure
thermodynamic properties of rock-forming minerals. American Journal of Science (UHP) and high-pressure (HP) metamorphic belts from central China. The Island
278-A 229 pp. Arc 3, 112–121.
Higashino, T., 1975. Biotite zone of the Sanbagawa belt in the Shiraga-yama district, Maruyama, S., Liou, J.G., Terabayashi, M., 1996. Blueschists and eclogites of the world
central Shikoku. Journal Geological Society of Japan 81, 653–670 (in Japanese with and their exhumation. International Geology Review 38, 485–594.
English abstract). Maruyama, S., Parkinson, C.D., Liou, J.G., 2002. In: Parkinson, C.D., Katayama, I., Liou, J.G.,
Higashino, T., 1990. Metamorphic zoning of the Sanbagawa belt, central Shikoku. Maruyama, S. (Eds.), Overview of the Tectonic Evolution of the Kokchetav Massif
Journal Geological Society of Japan 96, 703–718 (in Japanese with English abstract). and the Role of Fluid in Subduction and Exhumation. Universal Academy Press,
Holland, T.J.B., Powell, R., 1998. An internally-consistent thermodynamic data set for Tokyo, pp. 427–442.
phases of petrological interest. Journal of Metamorphic Geology 16, 309–343. Maruyama, S., Omori, S., Iwase, Y., 2004. On-going regional metamorphism under
Hubbard, M.S., Harrison, T.M., 1989. 40Ar–39Ar age constraints on deformation and the Japanese islands. Journal of Geography 113, 600–616 (in Japanese with English
metamorphism in the Main Central Thrust zone and Tibetan slab, eastern Nepal abstract).
Himalaya. Tectonics 8, 865–880. Maruyama, S., Santosh, M., Zhao, D., 2007. Superplume, supercontinent, and post-
Inui, M., Toriumi, M., 2002. Prograde pressure–temperature paths in the pelitic schists perovskite: mantle dynamics and anti-plate tectonics on the core–mantle boundary.
of the Sanbagawa metamorphic belt, SW Japan. Journal of Metamorphic Geology Gondwana Research 11, 7–37.
20, 563–580. Maruyama, S., Hasegwa, A., Santosh, M., Kogiso, T., Omori, S., Nakamura, H., Kawai, K.,
Irifune, T., Ringwood, A.E., Hibberson, W.O., 1994. Subduction of continental crust and Zhao, D., 2009. The dynamics of big mantle wedge, magma factory, and
terrigenous and pelagic sediments: an experimental study. Earth and Planetary Science metamorphic–metasomatic factory in subduction zones. Gondwana Research 16,
Letters 126, 351–368. 414–430.
Ishikawa, A., Kaneko, I., Kadarusman, A., Ota, T., 2007. Multiple generations of forearc mafic– Masago, H., Okamoto, K., Terabayashi, M., 2004. Exhumation tectonics of the
ultramafic rocks in the Timor–Tanimbar ophiolite, eastern Indonesia. Gondwana Sanbagawa high-pressure metamorphic belt, SW Japan — constraints from the
Research 11, 200–217. upper and lower boundary faults. International Geology Review 47, 1194–1206.
Isozaki, Y., Itaya, T., 1990. Chronology of Sanbagawa metamorphism. Journal of Masago, H., Omori, S., Maruyama, S., 2009. Counter-clockwise prograde P–T path in
Metamorphic Geology 8, 401–411. collisional orogeny and water subduction at the Precambrian–Cambrian boundary:
Isozaki, Y., Maruyama, S., 1991. Studies on orogeny based on plate tectonics in Japan the ultrahigh-pressure pelitic schist in the Kokchetav massif, northern Kazakhstan.
and new geotectonic subdivision of the Japanese Islands. Journal of Geography 100, Gondwana Research 15, 137–150.
697–761 (in Japanese with English abstracts). Masago, H., Omori, S., Maruyama, S., 2010. Significance of retrograde hydration in
Itaya, T., Takasugi, H., 1988. Muscovite K–Ar ages of the Sanbagawa schists, Japan and collisional metamorphism: a case study of water infiltration in the Kokchetav
argon depletion during cooling and deformation. Contributions to Mineralogy and ultrahigh-pressure metamorphic rocks, northern Kazakhstan. Gondwana Research
Petrology 100, 281–291. 18, 205–212 (this issue).
Kaneko, Y., 1997. Two-step exhumation model of the Himalayan metamorphic belt, Meade, C., Jeanloz, R., 1991. Deep-focus earthquakes and recycling of water into Earth's
central Nepal. Journal Geological Society of Japan 103, 203–226. mantle. Science 252, 68–72.
Kaneko, Y., Maruyama, S., Terabayashi, M., Yamamoto, H., Ishikawa, M., Anma, R., Miura, S., Kodaira, S., Nakanishi, A., Tsuru, T., Takahashi, N., Hirata, N., Kaneda, Y., 2003.
Parkinson, C.D., Ota, T., Nakajima, Y., Katayama, I., Yamauchi, K., 2000. Geology of Structural characteristics controlling the seismicity of southern Japan Trench
the Kokchetav UHP–HP metamorphic belt, northern Kazakhstan. The Island Arc 9, fore-arc region, revealed by ocean bottom seismographic data. Tectonophysics
264–283. 363, 79–102.
Kaneko, Y., Katayama, I., Yamamoto, H., Misawa, K., Ishikawa, M., Rehman, H.U., Kausar, Miyashiro, A., 1961. Evolution of metamorphic belts. Journal of Petrology 2, 277–311.
A.B., Shiraishi, K., 2003. Timing of Himalayan ultrahigh-pressure metamorphism: Miyashiro, A., 1965. Metamorphism and Metamorphic Belts. Iwanami-shoten, Tokyo.
sinking rate and subduction angle of the Indian continental crust beneath Asia. 458 pp. (in Japanese).
Journal of Metamorphic Geology 21, 589–599. Miyashiro, A., 1973. Metamorphism and Metamorphic belts. George Allen and Unwin,
Kaneko, Y., Maruyama, S., Kadarusman, A., Ota, T., Ishikawa, M., Tsujimori, T., Ishikawa, London. 492 pp.
A., Okamoto, K., 2007. On-going orogeny in the outer-arc of the Timor–Tanimbar Miyashiro, A., 1994a. Metamorphic Petrology. UCL Press, London. 404 pp.
region, eastern Indonesia. Gondwana Research 11, 218–233. Miyashiro, A., 1994b. Metamorphism. Iwanami-shoten, Tokyo. 256 pp. (in Japanese).
Katayama, I., Zayachkovsky, A., Maruyama, S., 2000. Progressive P–T records from zircon Miyashiro, A., 1998. What is Science Revolution? Iwanami-shoten, Tokyo. 347 pp. (in
in Kokchetav UHP–HP rocks, northern Kazakhstan. The Island Arc 9, 417–428. Japanese).
S. Maruyama et al. / Gondwana Research 18 (2010) 106–137 137

Mork, M.B.E., 1985. Incomplete high P-T metamorphic transitions within the Kvamsoy Santosh, M., Maruyama, S., Sato, K., 2009a. Anatomy of a Cambrian suture in Gondwana:
pyroxenite complex, west Norway: a case study of disequilibrium. Journal of Pacific-type orogeny in southern India? Gondwana Research 16, 321–341.
Metamorphic Geology 3, 245–264. Santosh, M., Maruyama, S., Omori, S., 2009b. A fluid factory in Solid Earth. Lithosphere
Nishimiya, Y., Tsunogae, T., Santosh, M., 2010. Sapphirine + quartz corona around (Geological Society of America) 1, 29–33.
magnesian (XMg 0.58) staurolite from the Palghat–Cauvery Suture Zone, southern Searle, M.P., Freyer, B.J., 1986. Garnet, tourmaline and muscovite bearing leucogranite,
India: evidence for high-pressure and ultrahigh-temperature metamorphism within gneisses and migmatites of the High Himalayas Zanskar, Kulu, Lahoul and Kashmir.
the Gondwana suture. Lithos 114, 490–502. In: Coward, M.P., Ries, A.C. (Eds.), Collision Tectonics: Geological Society London
Nishimura, Y., 1990. ‘Sangun metamorphic rocks’: terrane problem. In: Ichikawa, K., Special Publication, 19, pp. 185–201.
Mizutani, S., Hara, I., Hada, S., Yao, A. (Eds.), “Pre-Cretaceous Terranes in Japan”, Seno, T., 1995. Basis of plate tectonics. Asakura-shoten. 190pp. (in Japanese).
224. Pub. IGCP, Osaka, pp. 63–79. Seno, T., Maruyama, S., 1984. Paleogeographic reconstruction and origin of the Philippine
Nugroho, H., Harris, R., Lestariya, A.W., Maruf, B., 2009. Plate boundary reorganization Sea. Tectonophysics 102, 53–84.
in the active Banda arc–continent collision: insights from new GPS measurements. Smith, D.C., 1984. Coesite in clinopyroxene in the Caledonides and its implications for
Tectonophysics 479, 52–65. geodynamics. Nature 310, 641–644.
Obara, K., 2002. Nonvolcanic deep tremor associated with subduction in southeast Japan. Spear, F. S., 1993. Metamorphic phase equilibria and pressure–temperature–time paths.
Science 296, 1679–1681. Mineral. Soc. America, Monograph, Washington, D. C. 799 pp.
Obara, K., 2009. Inhomogeneous distribution of deep slow earthquake activity along the Spear, F.S., Kohn, M.J., Florence, F., Menard, T., 1990. A model for garnet and plagioclase
strike of the subducting Philippine Sea Plate. Gondwana Research 16, 512–526. growth in pelitic schists: implications for thermobarometry and P–T path determina-
Ogasawara, Y., Ohta, M., Fukasawa, K., Katayama, I., Maruyama, S., 2002. Petrology of tions. Journal of Metamorphic Geology 8, 683–696.
diamond-bearing dolomite marble from Kumdy-Kol. In: Parkinson, C.D., Katayama, I., Tabata, H., Maruyama, S., Shi, Z., 1998. Metamorphic zoning and thermobaric structure
Liou, J.G., Maruyama, S. (Eds.), Diamond-bearing from Kokchetav Massif, Kazakhstan. of the Dabie UHP–HP terrane, central China. The Island Arc 7, 142–158.
Universal Academy Press, Tokyo, pp. 191–211. Takada, Y., Matsuura, M., 2002. Structural evolution of the higher Himalaya; development
Okamoto, A., Toriumi, M., 2001. P–T paths of amphiboles by Gibbs method. Contributions of ramp structure through time and its mechanism to mountain building. Monthly
to Mineralogy and Petrology 141, 168–181. Magazine The Earth 24, 285–290 (in Japanese).
Okamoto, A., Toriumi, M., 2004. Optimal mixing properties of calcic and subcalcic Takasu, A., 1989. P–T histories of peridotite and amphibolite tectonic blocks in the
amphiboles: application of Gibbs method to the Sanbagawa schists, SW Japan. Sambagawa metamorphic belt, Japan. In: Daly, J.S., Cliff, R.A., Yardley, B.W.D. (Eds.),
Contributions to Mineralogy and Petrology 146, 529–545. The Evolution of Metamorphic Belts: Geological Society of London Special Publication,
Okamoto, K., Shinjoe, H., Katayama, I., Terada, K., Sano, Y., Johnson, S., 2004. SHRIMP U–Pb 43, pp. 533–538.
zircon dating of quartz-bearing eclogite from the Sanbagawa Belt, south-west Japan: Takasu, A., Dallmeyer, R.D., 1990. 40Ar/39Ar mineral age constraints for the tectonothermal
implications for metamorphic evolution of subducted protolith. Terra Nova 16, 81–89. evolution of the Sanbagawa metamorphic belt, central Shikou, Japan: a Cretaceous
Omori, S., Masago, H., 2004. Application of forward modeling to estimation of the accretionary prism. Tectonophysics 185, 11–39.
metamorphic P–T path. Journal of Geography 113, 647–663 (in Japanese with English Terabayashi, M., Ota, T., Yamamoto, H., Kaneko, Y., 2002. Contact metamorphism of the
abstract). Daulet Suite by solid-state emplacement of the Kokchetav UHP–HP metamorphic
Omori, S., Kamiya, S., Maruyama, S., Zhao, D., 2002. Morphology of the intraslab seismic slab. International Geology Review 44, 819–830.
zone and devolatilization phase equilibria of the subducting slab peridotite. Bull. Tirone, M., Ganguly, J., 2010. Garnet compositions as recorders of P–T–t history of
Earthquake Research Institute University of Tokyo 76, 455–478. metamorphic rocks. Gondwana Research 18, 138–146 (this issue).
Omori, S., Kita, S., Maruyama, S., Santosh, M., 2009. Pressure–temperature conditions of Toriumi, M., 1975. Petrological study of the Sanbagawa metamorphic rocks, the Kanto
ongoing regional metamorphism beneath the Japanese Islands. Gondwana Research Mountains, Central Japan. Univ. Museum, Univ. Tokyo Bull., 9. 99 pp.
16, 458–469. Toriumi, M., 1982. Strain, stress and uplift. Tectonics 1, 57–74.
Osada, M., Abe, K., 1981. Mechanism and tectonic implications of the Great Banda Sea Toriumi, M., 1985. Two types of ductile deformation/regional metamorphic belt.
earthquake of November 4, 1963. Physics of the Earth and Planetary Interiors 25, Tectonophysics 113, 307–326.
129–139. Toriumi, M., Kohsaka, Y., 1995. Cyclic P–T paths and plastic deformation of eclogite mass
Ota, T., Terabayashi, M., Katayama, I., 2004. Thermobaric structure and metamorphic in the Sambagawa metamorphic belt. Journal of Geology and Geography University
evolution of the Iratsu eclogite body in the Sanbagawa belt, central Shikoku, Japan. of Tokyo, Series III 22, 211–231.
Lithos 73, 95–126. Toriumi, M., Noda, H., 1986. The origin of strain patterns resulting from contemporaneous
Park, J.-P., Tsuru, T., Nakanishi, N., Hori, T., Kaneda, S., Nakanishi, A., Miura, S., Kaneda, Y., deformation and metamorphism in the Sambagawa metamorphic belt. Journal of
2002. A deep strong reflector of the Nankai accretionary wedge from multichannel Metamorphic Geology 4, 409–420.
seismic data: implications for underplating and interseismic shear stress release. Tsujimori, T., Sisson, V.B., Liou, J.G., Harlow, G.E., Sorena, S.S., 2006. Very-low-
Journal Geophysical Research 107 (ESE3-1), 3–16. temperature record of the subduction process: a review of worldwide lawsonite
Parkinson, C.D., Maruyama, S., Liou, J.G., Kohn, J.S., 2002. Probable prevalence of coesite- eclogites. Lithos 92, 609–662.
bearing metamorphism in collisional orogens and a reinterpretation of Barrovian Wang, X., Liou, J.G., Maruyama, S., 1992. Coesite-bearing ecologites from the Dabie
metamorphism. In: Parkinson, C.D., Katayama, I., Liou, J.G., Maruyama, S. (Eds.), The Mountains, Central China — petrogenesis, P–T paths, and implications for regional
Diamond-bearing Kokchetav Massif Kazakhstan. Universal Academic Press. tectonics. Journal of Geology 100, 231–250.
Peacock, S.M., 1993. Large-scale hydration of the lithosphere above subducting slabs. Watanabe, T., Leitch, E.C., Fanning, M., 1997. The age and tectonic significance of
Chemical Geology 103, 45–59. lawsonite eclogite and high temperature eclogite blocks in serpentinite mélange
Peacock, S.M., 1996. Thermal and petrologic structure of subduction zone. In: Bebout, G.E., from the southern New England Fold Belt, Eastern Australia. International Eclogite
Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), “Subduction : Top to Bottom”: AGU, Geophysical Conference, abstract volume. Terra Nova, p. 110.
Monograph, 96, pp. 119–133. Yamamoto, H., Okamoto, K., Kaneko, Y., Terabayashi, M., 2004. Southward extrusion of
Ramos, V.A., 1999. Plate tectonics setting of the Andean Cordillera. Episodes 22, 183–190. eclogite-bearing mafic–ultramafic bodies in the Sanbagawa belt, central Shikoku,
Ramos, V.A., Aleman, A., 2000. Tectonic evolution of Andes. In: Cordani, U.G., Milani, E.J., Japan. Tectonophysics 387, 151–168.
Thomaz, F.A., Campos, D.A. (Eds.), Tectonic Evolution of South America: 31st Yamamoto, S., Nakajima, J., Hasegawa, A., Maruyama, S., 2009. Izu-Bonin arc subduction
International Geological Congress Special Issue, pp. 635–685. under the Honshu island, Japan: evidence from geological and seismological aspect.
Ramos, V.A., Cegarra, M., Cristallini, E., 1996. Cenozic tectonics of the High Andes of Gondwana Research 16, 572–580.
west-central Argentina (30°–36°S latitude). Tectonophysics 259, 185–200. Yoshikawa, T., Kaizuka, S., Ohta, Y., 1964. Coastal terrace and crustal movement along
Sachan, H., Mukherjee, B., Ogasawara, Y., Maruyama, S., Ishida, H., Muko, A., Yoshioka, the northeast coast of Tosa Bay. Geographical Review of Japan 37, 627–648 (in
N., 2004. Discovery of coesite from Indus suture zone (ISZ), Ladakh, India; evidence Japanese).
for deep subduction. European Journal of Mineralogy 16, 235–240. Zhao, D., 2004. Global tomographic images of mantle plumes and subducting slabs:
Sakai, H., 2002. Rapid uplift of fore-land basin in front of the Himalaya ca. 1 Ma. Monthly Insight into deep Earth dynamics. Physics of the Earth and Planetary Interiors 146,
Magazine The Earth 24, 279–284 (in Japanese). 3–34.
Santosh, M., Kusky, T., 2010. Origin of paired high pressure–ultrahigh-temperature Zhang, R.Y., Liou, J.G., Ernst, W.G., 2009. The Dabie–Sulu continental collision zone: a
orogens: a ridge subduction and slab window model. Terra Nova 22, 35–42. comprehensive review. Gondwana Research 16, 1–26.

You might also like