You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/369803026

Mechanisms to generate ultrahigh-temperature metamorphism

Article in Nature Reviews Earth & Environment · April 2023


DOI: 10.1038/s43017-023-00403-2

CITATION READS

1 1,364

10 authors, including:

Shujuan Jiao Michael Brown


Institute of Geology and Geophysics, Chinese Academy of Sciences University of Maryland, College Park
38 PUBLICATIONS 1,217 CITATIONS 219 PUBLICATIONS 13,767 CITATIONS

SEE PROFILE SEE PROFILE

Priyadarshi Chowdhury Chris Clark


National Institute of Science Education and Research Curtin University
29 PUBLICATIONS 454 CITATIONS 178 PUBLICATIONS 7,771 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Compilation of Metamorphic History information in Western Australia View project

Metamorphic Evolution of the South West Yilgarn View project

All content following this page was uploaded by Priyadarshi Chowdhury on 12 April 2023.

The user has requested enhancement of the downloaded file.


nature reviews earth & environment https://doi.org/10.1038/s43017-023-00403-2

Review article Check for updates

Mechanisms to generate ultrahigh-


temperature metamorphism
Shujuan Jiao 1,2 , Michael Brown 3, Ross N. Mitchell 1,2,4, Priyadarshi Chowdhury 5
, Chris Clark6, Lin Chen ,
1,2

Yi Chen 1,2, Fawna Korhonen7, Guangyu Huang 1,2 & Jinghui Guo1,2
Abstract Sections

Ultrahigh-temperature (UHT; Tmax ≥ 900 °C) metamorphism requires Introduction

unusually high heat in continental crust at depths of 15–55 km, but Main characteristics of UHT
how such extreme thermal conditions are achieved is enigmatic. In metamorphism

this Review, we investigate UHT metamorphism, based on advances Competing mechanisms


in metamorphic petrology and numerical modelling, to identify the Implications for
tectonic settings where UHT metamorphism occurs. UHT rocks are metamorphism today

spatially related to convergent plate margins and spatially correlate Summary and future
perspectives
with the assembly of supercontinents, such as the formation of
Rodinia (1,350–850 Ma). Commonly, UHT occurrences are linked
to arc–backarc systems, thinned lithosphere or orogenic plateaus.
Elevated mantle heat in younger arc and backarc systems is related
to slab rollback, whereas thinned lithosphere in ancient orogens is
related to lithospheric peeling or shallow slab breakoff. By contrast,
UHT metamorphism in orogenic plateaus is a result of radiogenic
heating during thickening, sometimes with elevated mantle heat during
orogenic collapse. Geophysical mapping of Moho temperature and
depth beneath present-day orogens reveals the locations where UHT
metamorphism is occurring, such as in the Tibetan Plateau and the
North American Cordillera. Future research should include improved
geodynamic modelling of UHT metamorphism and the respective
tectonic settings to establish quantitative correlations between a viable
heat source and the spatial extent of UHT metamorphism.

1
State Key Laboratory of Lithospheric Evolution, Institute of Geology and Geophysics, Chinese Academy of
Sciences, Beijing, China. 2Innovation Academy of Earth Science, Chinese Academy of Sciences, Beijing, China.
3
Laboratory for Crustal Petrology, Department of Geology, University of Maryland, College Park, MD, USA.
4
College of Earth and Planetary Sciences, University of Chinese Academy of Sciences, Beijing, China. 5School of
Earth and Planetary Sciences, National Institute of Science Education and Research, HBNI, Bhubaneswar, Odisha,
India. 6School of Earth and Planetary Sciences, Curtin University, Perth, WA, Australia. 7Geological Survey of
Western Australia, East Perth, WA, Australia. e-mail: jiaoshujuan0215@126.com

Nature Reviews Earth & Environment


Review article

Key points Temperature exerts a first-order control on lithospheric strength


and profoundly influences lithosphere stability and evolution. The
thermal state of the crust primarily depends on internal heat produc-
•• Occurrences of ultrahigh-temperature (UHT) metamorphism tion, mainly by radiogenic heating (with secondary contributions
are widespread in the geological record since about 3,100 Ma. from viscous dissipation), and mantle heat flux, with or without heat
After the Archaean, since about 2,200 Ma, the mechanisms of UHT advection via magmatism11. The relative importance of these heat
metamorphism are closely related to the supercontinent cycle and sources varies in different tectonic settings, and their respective con-
tectonic processes at convergent plate margins. tributions to UHT metamorphism is a matter of ongoing debate. Dur-
ing plate convergence, thickening of continental crust during plate
•• The most likely tectonic settings to generate UHT metamorphism collisions12–14 and lithospheric thinning in arc–backarc systems15,16 have
are at depth in arc–backarc systems, in continental collisional belts, commonly been proposed as potential tectonic settings for regional
and in thinned orogens associated with lithospheric extension or UHT metamorphism (Fig. 1b–c). In addition, various tectonic processes
delamination. have been suggested that have the potential to achieve UHT metamor-
phic conditions in specific circumstances, including slab rollback17,
•• In arc–backarc settings and/or thinned orogens, UHT metamorphism ridge subduction and the formation of slab windows18, slab breakoff19
is characterized by isobaric or decompression heating and then and lithospheric delamination20–22.
cooling-dominated pressure–temperature (P–T) paths, and short Advances in petrological techniques since the early 2000s, espe-
durations of heating (≤30 Myr). Heating is dominated by elevated cially the development of thermometers based on slow-diffusing
mantle heat flux beneath thinned lithosphere, probably due to slab elements23–28, the maturation of phase equilibrium modelling29–32,
rollback in Phanerozoic orogens, and due to lithospheric peeling or and the introduction of chemically and texturally constrained petro­
shallow slab breakoff in Precambrian orogens. chronology33–37, have allowed for better characterization of UHT meta-
morphic pressure–temperature–time (P–T–t) paths and determination
•• In continental collisional belts, eclogite- or HP granulite-facies of the duration of metamorphic events38–43. Improvements have also
metamorphism is commonly followed by UHT metamorphism been made to understanding the geodynamic processes that have been
generated by heating during decompression; thus, the event is involved in generating UHT metamorphism, based on comparisons
characterized by a clockwise P–T path. Such UHT metamorphism is between P–T–t and T–t paths retrieved from natural occurrences and
most likely caused by a combination of radiogenic heating in thickened those predicted from numerical models21,42,44. Importantly, geophysical
crust and then elevated mantle heat flux during orogenic collapse. data for typical orogenic settings have been synthesized to determine
and map the present-day Moho depth and temperature45,46, which
•• Alternatively, UHT metamorphism can be generated in continental has enabled occurrences of ongoing UHT metamorphism and their
crust that has unusually high radiogenic heat production — although tectonic settings to be inferred. As a result of these developments, the
radiogenic heating alone is unlikely to be the sole cause of UHT first-order features of UHT metamorphism are better understood as it
metamorphism. In these rare cases, the duration of heating is generally is now possible to synthesize the key petrological attributes with results
longer (≥30 Myr), and the retrograde P–T paths are commonly from numerical and geophysical models. For example, the duration
decompression dominated. of heating during UHT is more tightly constrained, which is a critical
part in discriminating between different heat sources and tectonic
••Secular variation in the T/P ratios and durations of UHT settings1,3,4,8,47. In addition, the spatiotemporal distribution of types of
metamorphism correlate with cooling rate, with the highest T/P ratios, metamorphism based on supercontinent reconstructions48 provides a
longest durations and lowest cooling rates occurring in the mid- basis for discussing the relationship between mechanisms to generate
Proterozoic (1,350–850 Ma) during the assembly of Rodinia. Orogenic UHT metamorphism and plate tectonics.
style in the mid-Proterozoic favoured the formation of large-scale, In this Review, we first summarize the characteristics of UHT meta-
long-lived UHT metamorphic terranes. morphic localities worldwide, and then evaluate the proposed heat
sources and tectonic settings based on insights from numerical model-
ling. Using these empirical and theoretical constraints, and considering
Introduction the distribution of localities where UHT metamorphism is inferred to
Ultrahigh-temperature (UHT) metamorphism represents the most be occurring on the Earth today, we discuss the most likely mechanisms
extreme thermal conditions that the continental crust is capable of to generate regional UHT metamorphism and the implications for the
sustaining at crustal depths of 15–55 km (refs. 1–4) (Fig. 1a–c). The evolution of plate tectonics. Finally, we summarize the primary features
continental crust records evidence of UHT metamorphism since of the major tectonic settings in which UHT metamorphism occurs and
the Mesoarchaean, approximately 3,100 million years ago (Ma), most the changes over time, consider areas of continued uncertainty
commonly in refractory crustal rocks that are the residue left after and suggest future research opportunities.
partial melting and melt loss5. Moreover, the spatiotemporal distribu-
tion of UHT granulites correlates well with convergent plate margins Main characteristics of UHT metamorphism
and the supercontinent cycle since about 2,200 Ma (refs. 2,6). Under- In this section, the P−T conditions and the spatiotemporal distribu-
standing how UHT metamorphic conditions are generated is critical to tion of UHT metamorphism, and its relationship to the late Archaean
interpreting the evolution of orogenic belts and interactions between (pre ~2500 Ma) formation of supercratons and the Proterozoic–
the crust and mantle1–4,7–10. However, there is currently no consensus Phanerozoic (since ~2200 Ma) supercontinent cycle are reviewed. Dis-
concerning the mechanisms by which UHT metamorphic conditions tinct microstructures in UHT granulites record two endmember types
are achieved. of post-peak metamorphic evolution, which are cooling dominated

Nature Reviews Earth & Environment


Review article

a P–T conditions of metamorphism b Continent–continent collision

d Preserved material of
on ite
30 Uncertainty iam hite UHP Coes z prior subduction UHT
D ap t
Gr Quar

100
z
rt
ua
25 + Q
ite

te
de Gn Am
bi
s

Ja UHT
Al
olidu

E-HPG
80 Am Gn E-HPG
E-HPG
Wet s

20
ite
an te
Ky ani

Depth (km)
m c Subduction zone and accretionary region
P (kbar)

li
Sil 1 60
Bl r–
15 °C kba
77.5

Fluid
40
10 Exhumation of rocks
of E and/or UHP Am, Gn, UHT

kbar E, UHP
–1

150 °C
5 20
UHT Muscovite-breakdown UHT granulites
Gr Am
An melting (peak–T ≥900 °C)
da Gn
lus Biotite-breakdown Common Am to
ite
melting Gn rocks
0
300 500 700 900 1,100
T (°C)

Fig. 1 | P–T conditions and possible tectonic settings of UHT and other high- whereas Am to Gn metamorphism occurs in the middle of the thickened crust;
grade metamorphism. a, The range of pressure–temperature (P–T) conditions UHT metamorphism can occur at any depth in the lower half of the thickened crust.
for ultrahigh-temperature (UHT) and common amphibolite- to granulite-facies Red teardrops represent mantle-derived melts and white teardrops represent
metamorphism. Each P–T datum records a point — the metamorphic peak — on a crust-derived melts. c, In orogenesis related to modern-style subduction, in
geotherm crossed during a dynamic evolution, and a convenient way to represent particular Phanerozoic orogens, UHT and common Am to Gn metamorphism
the metamorphic peak is using the thermobaric ratio (T/P). Uncertainties on the generally occur in the arc–backarc region of the overriding plate, while eclogite
P–T data (±1 kbar and ±50 °C) are discussed in the main text. UHT metamorphism (E) and ultrahigh-pressure metamorphism occur along the subducting slab. Bl,
generally occurs at temperatures above the biotite-breakdown melting reaction, blueschist facies; Gr, greenschist facies. Configuration of P–T space, metamorphic
is generally limited to the sillimanite stability field, and has similar T/P to common facies, and the location of the wet solidus in pelitic rocks in panel a taken from
amphibolite (Am) to granulite (Gn) facies rocks but achieves more extreme peak ref. 220. The temperature range for muscovite- and biotite-breakdown melting in
temperatures. b, In continental collisions, eclogite–high-pressure-granulite panel a is modified from ref. 221. Datasets are updated from ref. 11 and provided
facies (E-HPG) metamorphism occurs close to the base of the thickened crust, in Supplementary Data 1. Panels b and c are adapted from ref. 222, Springer Nature.

and decompression dominated. The durations of metamorphic events 5−15% at 10 kbar and 1,000 °C. UHT metamorphism spans a similar
that reach UHT metamorphic conditions are variable (10–160 Myr), range of T/P ratios to common amphibolite- to granulite-facies meta­
but the prograde heating stage is generally short (≤30 Myr). In this sec- morphism, but at higher peak temperatures (Fig. 1a). The highest
tion, the main characteristics of UHT metamorphism are summarized. metamorphic T/P values were recorded during the amalgamation of
These provide constraints on the tectonic settings and mechanisms of supercontinent Rodinia during the period 1,350–850 Ma (Box 1). There
formation discussed in the following section. are 19 high-grade terranes where outcrop is sufficient to demonstrate
areal extents of UHT metamorphism that exceed ~500 km2 (Supple-
P−T and spatiotemporal distribution mentary Data 2); the largest of these is the Mesoproterozoic Musgrave
There are approximately 110 localities worldwide where the peak P−T Province of central Australia with UHT metamorphism covering an area
conditions of UHT metamorphism have been quantified, with meta- of ~120,000 km2 (ref. 52).
morphic ages that range from about 3,100 Ma to the present day42,49,50 Since the early Palaeoproterozoic (about 2,200 Ma), the occur-
(Supplementary Data 1; Fig. 2a). These P−T conditions were determined rence of both UHT and common amphibolite- to granulite-facies
using various thermobarometric methods, which have differing uncer- metamorphism correlates well with the supercontinent cycle, that is,
tainties on both P and T across the range of P−T conditions shown in the successive assembly of Columbia, Rodinia, Pangaea and Amasia, the
Fig. 1a (refs. 2,31,51). Notwithstanding, for the P−T data determined last of which is still in progress (Fig. 2b–e). The formation of the old-
from common amphibolite- to granulite-facies and UHT rocks used est supercontinent from rifted cratonic fragments is evidence of
herein (Supplementary Data 1 and Fig. 1a), uncertainties of ±1 kbar and global horizontal displacements and provides a minimum age for the
±50 °C are recommended as appropriate minima, which lead to uncer- linking up of the Earth’s global plate-tectonic network53,54. Assuming
tainties on the thermobaric ratio (T/P in °C kbar−1) of approximately no substantial exposure bias, most preserved UHT metamorphism

Nature Reviews Earth & Environment


Review article

a 220

UHT granulites
(peak-T ≥900 °C)
180
Common amphibolite-
to granulite-facies
rocks
Thermobaric ratio (°C kbar–1)

140

100

60

20
4,000 3,500 3,000 2,500 2,000 1,500 1,000 500 0
Metamorphic age (Ma)

b Columbia (2,200–1,350 Ma) d Pangaea (850–200 Ma) (including Gondwana megacontinent)

60°

60°
30°

30°

0° 0°

–30°

–30°
–60°

–60°

c Rodinia (1,350–850 Ma) e Amasia (200–0 Ma) (including Eurasia megacontinent)

60°

60°
30°

30°


–30°

–30°

–60°

–60°

Nature Reviews Earth & Environment


Review article

Fig. 2 | Spatiotemporal distribution of UHT metamorphism and correlation Pangaea, which includes the early-formed Gondwana megacontinent223.
with the supercontinent cycle. a, Secular change of thermobaric ratio (T/P, e, As b, for metamorphic ages of 200–0 Ma plotted on the nascent supercontinent
in °C kbar−1) for ultrahigh temperature (UHT) and common amphibolite- to Amasia, which includes the early-formed Eurasia megacontinent223. Note that
granulite-facies metamorphism through time. b, Spatial distribution of UHT the highest thermobaric ratios occurred in the mid-Proterozoic. Supercontinent
and common amphibolite- to granulite-facies localities with metamorphic reconstruction data in panels b–e from ref. 48. Datasets updated from ref. 11,
ages of 2,200–1,350 Ma plotted on the supercontinent Columbia. c, As b, for shown in Supplementary Data 1. In each case, ultrahigh-temperature (UHT)
metamorphic ages of 1,350–850 Ma plotted on the supercontinent Rodinia. metamorphism is widely distributed and is generally located along or close
d, As b, for metamorphic ages of 850–200 Ma plotted on the supercontinent to a continental margin, representing a probable convergent plate boundary.

is located close to the edges of continental fragments in supercon- of fluid ingress, retrograde reactions were inhibited. Examples of
tinents (Fig. 2b–e), suggesting a close relationship with convergent microstructures developed during high-temperature cooling include
plate margins48, although some common amphibolite- to granulite- the reactions of sapphirine + quartz to orthopyroxene + sillimanite
facies metamorphism occurs in intraplate settings, most notably in (Fig. 3e) and spinel + quartz to garnet + sillimanite (Fig. 3f), and
the Proterozoic supercontinent Columbia (Fig. 2b). Therefore, since the replacement of osumilite by a characteristic symplectite of
at least about 2,200 Ma, mechanisms to generate UHT metamorphism cordierite + orthopyroxene + K-feldspar7,72,73 (Fig. 3g).
are generally related to processes associated with plate tectonics, UHT localities with post-peak decompression-dominated paths
particularly at convergent plate boundaries. generally record a pressure drop of >2 kbar, although the associated
Twenty localities record UHT metamorphism of Archaean age, decrease in temperature is variable (Fig. 3h), implying distinct exhuma-
from about 3,100 Ma to 2,500 Ma: that is, during a period of craton tion mechanisms following the UHT peak. There are numerous exam-
stabilization and formation of several supercratons before the start ples of symplectitic mineral assemblages that replace peak minerals
of the supercontinent cycle54–56 (Fig. 2a). Based on a variety of proxies during decompression, such as sapphirine or spinel and cordierite
in the geological record57–63 and numerical modelling studies22,64, this replacing sillimanite (Fig. 3i–j), and sapphirine and orthopyroxene
period could represent the transition to plate tectonics, which followed replacing garnet74–77 (Fig. 3k).
after a stagnant-to-sluggish lid tectonic regime dominated by mantle Compound P–T paths have also been recovered, where a first
plumes, lithospheric delamination and intermittent lithospheric over- decompression heating stage was followed by cooling, and then either a
turn events65–68. Therefore, the dominant mechanisms that generated second decompression or a second episode of decompression heating
UHT metamorphism in the late Archaean might differ from those (Fig. 3l). Such compound P–T paths either indicate multiple tectonic
since the Archaean, and lithospheric delamination and mantle plumes events or the evolution of the thermal structure of an orogen with
probably were more important at that time. time, such as the addition of mantle heat during orogenic collapse
after initial radiogenic heating during thickening. Sillimanite replac-
P–T–t evolution ing kyanite is one example of petrological evidence that supports
P–T–t paths are regarded as diagnostic of specific tectonic settings such compound P–T paths78–80 (Path 5 in Fig. 3l; Fig. 3m). In samples
for the formation of UHT metamorphism. Before the turn of the from the Xiaoshizi locality in the Khondalite Belt, North China craton,
century, the UHT metamorphic peak was generally identified in the lit- the change of CaO (a pressure-sensitive indicator81,82) with yttrium
erature by diagnostic equilibrium mineral assemblages, such as (Y) in garnet indicates initial growth under isobaric conditions fol-
sapphirine + quartz + orthopyroxene (Fig. 3a) and orthopyroxene + lowed by growth during decreasing pressure towards the rim before a
sillimanite + quartz (Fig. 3b). However, these assemblages are restricted kick-up in pressure close to the rim83 (Fig. 3n). Similarly, the change of
to specific bulk rock compositions (largely magnesian metapelites), titanium (Ti; a temperature-sensitive indicator84,85) with Y indicates a
limiting their utility. More recently, since about 2000, the peak P–T temperature increase during early growth of garnet, while tempera-
conditions and P–T paths retrieved from UHT metamorphic rocks are ture decreased during subsequent growth83 (Fig. 3n). The decrease
constrained by various thermobarometric methods (Box 1). For exam- of heavy rare-earth elements (HREEs) and Y from garnet core to rim
ple, the coexistence of mesoperthite with metamorphic orthopyroxene indicates garnet growth without any resorption or additional source
requires high temperatures (Fig. 3c), which can be quantified using the of HREEs + Y, because HREEs and Y are preferentially incorporated into
two-feldspar thermometer to establish the presence of UHT metamor- garnet during growth86. Combining these data reveals a decompression
phism69–71. Two commonly referenced endmember types of post-peak heating stage followed by a cooling-dominated retrograde stage83,87,
P–T path are one dominated by post-peak cooling (sometimes referred similar to Path 6 in Fig. 3l.
to as isobaric cooling or an IBC path) and another dominated by decom- The prograde heating (pre-peak) stage of the metamorphic evolu-
pression (sometimes referred to as isothermal decompression or an ITD tion is rarely well constrained. First, reaction rates and elemental diffu-
path). An IBC path indicates that thickening and subsequent exhuma- sion in minerals are both faster at higher temperatures, and therefore
tion were limited and the UHT granulites remained in the lower crust earlier microstructural and chemical records are generally overprinted
for a protracted period, whereas an ITD path shows that the rocks have before the metamorphic peak is achieved. However, in some localities,
been exhumed soon after the thermal peak, probably because of thick- evidence, such as the preservation of cordierite and/or staurolite as
ening followed by orogenic collapse. Nevertheless, many retrograde inclusions in garnet88–90, demonstrates increasing P and T along the
P–T paths lie between these two endmembers. prograde path prior to reaching UHT conditions. Second, retrograde
UHT localities with post-peak cooling-dominated paths generally overprinting reactions, particularly those related to crystallization of
record a pressure drop of <2 kbar (Fig. 3d), implying limited exhuma- residual melt, can obliterate evidence of the prograde P–T path.
tion after the UHT peak. Thus, these localities remained in the deep In summary, based on compiled P–T–t data from the literature
crust during slow cooling to the solidus, after which, in the absence (Supplementary Data 1), there is no clear secular change in the style

Nature Reviews Earth & Environment


Review article

Box 1

Determining P–T–t paths and timescales


Pressure–temperature–time (P–T–t) paths record information diffusion chronometry, can also provide constraints on the heating–
related to the burial, residence at depth and exhumation of rocks, cooling and/or burial–exhumation rates of metamorphic rocks229,230.
the associated heating and cooling, and the timescale from Using these methods, the duration of ultrahigh-temperature (UHT)
initial burial to final exhumation of rocks. The P–T conditions can metamorphic events and the length of the prograde heating stages
be calculated either by using the compositions of minerals in can be determined (box figure, parts a,b).
combination with the equilibrium thermodynamics of balanced A compilation of geographically distributed P–T–t data has allowed
reactions among mineral endmembers (that is, thermobarometry), an assessment of secular variation for multiple variables, grouped
or by using phase equilibrium modelling for a given rock by supercontinent or supercraton. In the box figure (parts c–f),
composition. For the latter method, combining the results with box-and-whisker plots are shown versus age of supercontinent or
observed mineral assemblages, proportions and compositions supercraton for the duration of the UHT metamorphic events, the
provides information related to the P–T path in addition to the duration of heating, the T/P ratios of UHT granulites, and cooling
peak P–T conditions31. rates. Gyr, billion years. The cooling rate data are from ref. 108.
It is also important to know the timing, duration and rates Numerical methods can be used to forward model P–T–t paths and
of processes and events. To determine these variables requires timescales on the basis of Fourier’s law and the heat flow equation.
geochronology on both rock-forming (for example garnet, micas) Fourier’s law is:
and accessory (for example zircon, monazite and titanite) minerals.
dT
These minerals have different closure temperatures for diffusion that Q = −k
dz
can be combined with microstructural and compositional information
so that an age can be linked to a specific P–T condition228. Diffusion where Q is heat flow in W m−2, k is thermal conductivity in J s−1 m−1 K−1,
modelling of compositional zoning preserved in minerals, termed z is the depth in metres, and dT/dz is the geothermal gradient.

a UHT metamorphic events record durations of 10–160 Myr, with one b UHT metamorphic events record durations of prograde heating of
locality at 250 Myr 5–30 Myr, with two localities of 40 and 60 Myr
200 50

An outlier: 250 Ma An outlier: 60 Ma


Duration of heating

40
150
Duration (Myr)

stage (Myr)

30
100
20
50 10

0 0
4,000 3,500 3,000 2,500 2,000 1,500 1,000 500 0 4,000 3,500 3,000 2,500 2,000 1,500 1,000 500 0
Metamorphic age (Ma) Metamorphic age (Ma)

c Duration of UHT metamorphism (Myr) d Duration of heating of UHT metamorphism (Myr)


300 60
Duration of heating

250 50
Duration (Myr)

Outlier
stage (Myr)

200 40
Median
150 30
100 Mean 20
50 10
0 0

e T/P of UHT granulites (°C kbar–1) f log10 cooling rate (°C Gyr–1)
250 3.5
3
Thermobaric ratio

log10 cooling rate


(°C billion year–1)

200
2.5
(°C kbar–1)

150 2
1.5
100 1
0.5
50
0
0 –0.5
s bia ia a a s bia ia a a
on din ) ae asi ) on din ) ae asi )
rat a) lum a) Ro Ma ng a) Am Ma rat a) lum a) Ro Ma ng a) Am Ma
p erc 00 M Co 50 M 0 Pa 0 M 0 p erc 00 M Co 50 M 0 Pa 0 M 0
5 20 – 5 20 –
Su >2,2 1,3 –8 0–
0 Su >2,2 1,3 –8 0–
0
( 0– 50 (20 ( 0– 50 (20
,20 (1,3 (85 ,20 (1,3 (85
(2 (2

Nature Reviews Earth & Environment


Review article

(continued from previous page)


The heat flow equation combines the contributions from conduction where T is temperature in K, t is time in seconds, κ is thermal
of mantle heat, radiogenic heat production and heat advection116, diffusivity in m2 s−1, A is heat production (from radioactive decay)
which is: in W m−3, ρ is density in kg m−3, c is heat capacity in J kg−1 K−1, and Uz
2 is velocity in m s−1. The temperature and pressure (or depth) history
dT ∂ T A dT at any time can be calculated for any point in the thermal model.
=κ 2 + − Uz
dt ∂z ρc dz

of P–T–t paths for UHT metamorphism, although IBC is more common both can provide better, if imperfect, constraints on the duration of an
than ITD in the Archaean record. In some high-grade metamorphic ter- event or a segment of the P–T–t evolution38–43. However, problems such
ranes, the post-peak P–T–t paths record either cooling-dominated or as isotopic resetting due to metamictization, volume diffusion and/or Pb
decompression-dominated trajectories from the UHT peak at different mobilization at submicrometre scale, and possible decoupling of zircon
locations, such as in the Eastern Ghats Province, India91, and the Rayner ages from their compositions at UHT conditions, could make both the
Complex, East Antarctica92. metamorphic age and duration of the event difficult to constrain104–107.
Such problems can be recognized by following systematic isotopic
Duration of UHT metamorphism and trace element analysis, and using high-spatial-resolution (~2 μm)
The duration of a metamorphic event is not well defined but is gener- scanning ion imaging of dated grains105–107.
ally considered to extend from diagenesis through to a metamorphic The reported durations of UHT metamorphic events are mostly
peak, and continue until the exhumation and cooling related to the in the range 10–160 Myr (Box 1). There is secular variation in the data,
associated orogenic event have ceased. As a result, it is generally not particularly in terms of supercontinent cycle, as the crust with the
feasible to determine when a metamorphic event began or finished highest T/P values correlate with the longest durations of UHT meta-
using geochronology4,47, even though it is clear that the duration of an morphic events, which occurred during amalgamation of the Rodinia
event provides a minimum estimate for the longevity of the heat source supercontinent (Box 1). Indeed, the mid-Proterozoic (1,350–850 Ma)
and the type of orogenesis involved4,21, both of which are important to period was characterized by the slowest post-peak cooling rates21,108
identifying mechanisms of formation. In one view, the duration of UHT (Box 1), orogens with relatively thin crust and generally low eleva-
metamorphism was defined as the time spent above 900 °C (that is, tions109,110, extensive massif-type anorthosites111, and only a limited
Δt900), and that of the granulite-facies metamorphism as Δt800 (ref. 47). number of long-lived passive margins112. These features, which require
Long-lived UHT metamorphism generally has Δt900 > 30 Myr, and short- a substantial contribution of mantle heat and elevated crust–mantle
lived UHT metamorphism has Δt900 < 10 Myr (ref. 47). Although such interaction110, have been explained by stagnation and/or a slowdown
definitions provide a useful measure of how long the crust has spent of plate tectonics109,113,114, and/or the incomplete breakup of the Colum-
under UHT conditions, which relates to specific tectonic settings47, the bia supercontinent in transforming into Rodinia115. By contrast, the
duration of the prograde heating versus the post-peak cooling and/or reported durations of the suprasolidus prograde heating stage lead-
decompression stages remain largely unknown. ing to UHT metamorphism are generally in the range 5–30 Myr, with
In most cases, the timing and duration of UHT metamorphism have two localities with durations of 40 and 60 Myr, respectively (Box 1).
been constrained by zircon and/or monazite U–Pb geochronology, However, the geochronological constraints on which these estimates
which means both are controlled by the mechanisms of zircon and are based are poor enough that the timescales of heating could be
monazite formation during the P–T evolution and the interpretation considered effectively unknown in many cases. Even so, the apparently
of the respective U–Pb ages4,93–95. Thus, both the duration of a UHT short durations of prograde stage suggest that mantle heat contributed
metamorphic event and the duration of the prograde heating segment to the development of UHT metamorphism at many of the localities
will be model-dependent. Zircon and monazite grow substantially in summarized herein.
melt-bearing systems93–98. As a result, in this Review, the duration of a UHT metamorphism is closely spatially related to the assembly of
metamorphic event leading to UHT conditions is defined as the period supercratons and supercontinents. The peak P–T conditions record
during which the crust was suprasolidus, that is, from the prograde secular variation, with the highest T/P ratios occurring in the mid-
crossing of the muscovite-breakdown reaction in rocks of appropriate Proterozoic (1350–850) during the formation of Rodinia, while the
composition at ~700 °C until the post-peak retrograde crossing of the retrograde P–T–t paths are variable between IBC and ITD types. Secu-
elevated solidus for the residual composition after melt loss99 (Fig. 1a). lar variation in the duration of UHT metamorphism correlates with
The duration of the suprasolidus prograde heating segment is then cooling rate, with the longest durations and the lowest cooling rates
defined as the period from the beginning of partial melting at ~700 °C occurring in the mid-Proterozoic. Based on limited data, the durations
to the peak temperature. of prograde heating seem to be short.
Through the use of microstructures, elemental mapping, chemi-
cal composition, and especially the partitioning of the HREEs between Competing mechanisms
minerals such as garnet and zircon33,100–102, advances have been made There are two primary sources of heat to generate regional metamor-
in linking the behaviour of zircon and monazite with the rock-forming phism: radiogenic heating (with secondary contributions from viscous
minerals generally used for thermobarometry, which has substantially dissipation), and mantle heat (including heat advected with magmas).
improved the interpretation of U–Pb ages41,99,103. Separate periods of Generating UHT metamorphic conditions requires that one or both
zircon and monazite growth are commonly recognized, although the primary heat sources are prevalent during orogenesis. Radiogenic
age ranges can overlap to some extent. Therefore, using age ranges from heating is dominant in thickened orogens because of the enriched

Nature Reviews Earth & Environment


Review article

Equilibrated mineral assemblages at UHT peak

a b c

Opx

Qz Opx
Spr
Qz

Sil Mesoperthite
Opx
500 µm 100 µm 400 µm

Post-peak cooling-dominated
d Cooling-dominated
an
ite
e
e Qz f g
retrograde P–T path Ky anit 60 Crd
Sil Qz
lim Sil
15 Sil
Grt
Depth (km)

Spr
P (kbar)

40 Crd+Opx+Kfs
10
Spl
1
Opx Sil
5 2 20
Qz
Qz
And 500 µm 200 µm 200 µm
alu
0 site
600 800 1,000 1,200
T (°C)

Post-peak decompression-dominated
h Decompression-dominated ite
i j
retrograde P–T path an
Ky anit
e k
4 60
illim Crd
15 S
Sil Spr+Opx
Grt
Depth (km)

40 Spr+Crd
P (kbar)

10 3
Crd

20 Spl
5 Spl+Crd
400 µm 20 µm 500 µm
Grt
0
600 800 1,000 1,200
T (°C)

Pre-peak decompression heating


l n 8.5 kbar
P stable to
Pre-peak decompression-
an
ite
e m Bt
1.2
Ky anit 60
Grt (wt%)

heating P–T path decreased


CaO in

8.0 kbar
lim P stable or
15 Sil 1.1
Sil p 7.5 kbar slightly
5 6 seud increased
Depth (km)

o 1
40 afte morph
P (kbar)

10 r Ky 900 °C
200
Ti in Grt (ppm)

Bt Spr+Crd
150 850 °C T increased
5 20 800 °C T decreased
Grt 100
750 °C
2,000 µm 50 700 °C
0 0
600 800 1,000 1,200 0 200 400 600 800
T (°C) Y in Grt (ppm)

concentrations of radioactive elements (U, Th and K) in the continen- in residual minerals during mantle melting, and therefore are concen-
tal crust, whereas mantle heat flux dominates at sites of lithospheric trated in the crust as compared with the mantle118. However, radiogenic
thinning and/or removal. This section assesses the mechanisms that heat production in the continental crust is variable over two orders of
can generate the heat required to achieve UHT metamorphism in magnitude, from 0.1 to 10 μW m−3, owing to the heterogeneous dis-
the continental crust and compares competing tectonic settings for the tribution of radioactive elements from lithology to lithology, and to
formation of UHT metamorphic rocks. metamorphic and magmatic processes117,119–122. Granites and metasedi-
mentary rocks have much higher radiogenic heat production (mean
Radiogenic heating value of 2.0–3.3 μWm−3) compared to that of metabasites (mean value
Earth’s surface heat flow comprises approximately equal amounts of ~0.08 μW m−3)119–121. Although partial melting and granulite-facies
derived from primordial heat left over from the formation of Earth metamorphism generally do not reduce the whole-rock Th budget,
and radiogenic heat produced by the radioactive decay of 238U, 235U, UHT metamorphism and substantial melt loss can deplete the crust in
232
Th and 40K (refs. 116,117). These elements are strongly incompatible Th121. The average radiogenic heat production in the continental crust

Nature Reviews Earth & Environment


Review article

Fig. 3 | Representative metamorphic features of UHT metapelitic granulites. by a symplectite of sapphirine and cordierite, Itabuna–Salvador–Curaçá Block;
a–c, Representative equilibrium ultrahigh-temperature (UHT) mineral ( j) multilayer corona of cordierite, spinel and symplectite of cordierite plus
assemblages: (a) sapphirine (Spr), quartz (Qz) and orthopyroxene (Opx), spinel separating garnet (Grt) from sillimanite, Khondalite Belt; (k) symplectite
Napier Complex, East Antarctica; (b) orthopyroxene, sillimanite (Sil) and of sapphirine and orthopyroxene replacing garnet, Southern Granulite Terrane,
quartz, Highland Complex, Sri Lanka; (c) mesoperthite in equilibrium with India. l, Pre-peak decompression-heating P–T paths80,87,226. m, Pseudomorph
orthopyroxene, Itabuna–Salvador–Curaçá Block, Brazil. d, Post-peak cooling- of sillimanite after kyanite (Ky) replaced by a symplectite of sapphirine and
dominated pressure–temperature (P–T) paths7,224. e–g, Photomicrographs cordierite, Palghat Cauvery Shear System, India. n, Change in CaO records the
showing representative cooling-dominated reaction microstructures: change in P, and change in Ti records the change in T during garnet growth,
(e) sapphirine and quartz reacted to form sillimanite and orthopyroxene, Khondalite Belt. Different symbols refer to different garnet grains. See text for
Eastern Ghats Province, India; (f) spinel and quartz reacted to form sillimanite explanation of the trends. Data in panel n are from ref. 83. Mineral abbreviations
and garnet, Khondalite Belt, North China craton; (g) pseudomorph of osumilite adapted from ref. 227. In summary, there are two endmember types of P–T path
completely replaced by a symplectite of cordierite (Crd), orthopyroxene and retrieved from UHT metamorphic rocks, those dominated by decompression
K-feldspar (Kfs), Napier Complex. h, Post-peak decompression-dominated and those dominated by cooling, and less commonly pre-peak temperature paths
P–T paths75,225. i–k, Photomicrographs showing representative decompression- characterized by decompression and heating that link an early high-pressure
dominated reaction microstructures: (i) pseudomorph of sillimanite replaced metamorphism with a late UHT overprint. Bt, biotite.

is about 1–1.2 μW m−3, but is less in the lower continental crust at about Long-lived UHT metamorphism, especially in terranes without
0.4–0.8 μW m−3 (ref. 119), which can be replicated with 20–30 vol% meta- coeval voluminous mantle-derived magmatism, has generally been
sedimentary rocks and 70–80 vol% metabasite121. The relatively high interpreted as a result of orogenic thickening of the crust and radio-
heat production of the lower crust in some high-grade metamorphic genic heating, such as the Napier Complex of East Antarctica7,47, the
terranes, such as 2.3 μW m−3 in the Ivera–Verbano Zone, northern Italy, Madurai Block of southern India129, the Namaqualand Complex, South
can be explained by a high proportion of metasedimentary rocks120. By Africa132, and the Reynolds–Anmatjira Ranges of central Australia133.
contrast, radiogenic heat production in the mantle is around 0.2 μW m−3 More generally, the moderate heat production (~2.0–2.5 μW m−3)
(ref. 116). After correcting for radioactive decay, it is apparent that and/or short-lived (≤40 Myr) nature of the pre-peak heating stage, as
crustal heat production remained relatively constant throughout Earth found in the Kerala Khondalite Belt of southern India134 and the Khon-
history, with median values for felsic igneous rocks (SiO2 > 60 wt%) of dalite Belt of the North China craton135, suggest that radiogenic heating
2.53 μW m−3 and for mafic igneous rocks (SiO2 ≤ 60 wt%) of 0.61 μW m−3 alone was insufficient to generate the UHT metamorphic conditions, and
(ref. 122). Enhanced radiogenic heating by thickening of the continental an elevated mantle heat flux was probably required as well. However, the
crust has long been regarded as a dominant mechanism for common relative importance of these heat sources can be difficult to pin down.
amphibolite- to granulite-facies44,123–126, and UHT metamorphism3,4,7,8,47. For example, initially, thermal models of an orogenic plateau setting for
One-dimensional thermal models suggest that the attainment of the Anosyen domain of southern Madagascar suggested that heating to
UHT conditions requires a much higher than average heat production UHT conditions could be achieved in about 65 Myr with the presence of
(>3 μW m−3), a thickened crust by a factor of 2–3, long-lived orogenesis a crustal layer at least 25 km thick with heat production of 3.5–5 μW m−3
(>60 Myr) and/or a low erosion rate (0–0.7 mm yr−1)127–129 (Fig. 4a). located at a depth of >20 km (ref. 136). However, new geochronology
Thermomechanical models have been used to investigate crustal-scale has limited the period of suprasolidus prograde heating to between
lateral transport of heat and mass in large hot orogens12–14,130 (Fig. 4b). 20 and 40 Myr, which is insufficient time to achieve UHT conditions by
Two specific models can reproduce UHT conditions: first, a model radiogenic heating alone, and elevated mantle heat flux or advection of
of the Himalayan–Tibetan Orogen12,13, hereafter referred to as HT1, heat with mantle-derived magmas was probably required73,90.
which allows laterally homogeneous ductile flow and was designed to To generate UHT metamorphism by radiogenic heating alone
investigate channel flow in the middle crust of the orogen; second, a requires either unusual initial conditions or higher than average radio-
model designed to investigate the evolution of the western Grenville genic heat production in the crust and long durations of heating, which
Orogen14, hereafter referred to as GO-ST87, in which heterogeneous contrasts with the short durations of prograde heating reported in the
ductile flow occurs because of the variable strength of the crustal current dataset (≤30 Myr). Thus, we conclude that in many cases it is
units14. In these models, the heat production was set at 2 μW m−3 for likely that UHT metamorphism in collisional orogens was generated by
the upper and middle crust, and 0.75 μW m−3 for the lower crust, with a combination of early radiogenic heating in thickened crust followed
a mantle heat flux of 20 mW m−2(Fig. 4b). In both the HT1 and GO-ST87 by an elevated mantle heat flux during exhumation from eclogite- or
models, radiogenic heating is dominant, and the time taken to reach HP granulite-facies conditions.
UHT conditions is long (in the range of 30–75 Myr; Fig. 4c). The resultant
P–T–t paths are clockwise. The pre-peak stage varies from isobaric Elevated mantle heat flux
heating or heating with decompression during fast convergence The mantle heat flux in stable continental shields is generally no less
to heating with compression if convergence is slow, whereas the post- than ~20 mW m−2 (ref. 137). Elevated mantle heat flux is caused by
peak stage is decompression-dominated owing to erosion128,129 or exhu- litho­spheric thinning due to extension or delamination that ensures
mation by lateral expulsion12–14 (Fig. 4d). Eclogite- or high-pressure (HP) astheno­spheric mantle upwelling. Several tectonic settings or concep-
granulite-facies metamorphism can occur prior to decompression tual models have been proposed to explain UHT metamorphism in this
heating to the UHT peak (Fig. 4d). Lastly, it is noted that in these two way, including arc–backarc systems (Fig. 5a–b), slab rollback (Fig. 5c),
models the Moho temperature based on the initial conditions is 704 °C slab breakoff (Fig. 5d), lithospheric delamination (including lithospheric
at 35 km depth (Fig. 4b), which is ~200 °C higher than that based on a foundering and lithospheric peeling (Fig. 5e)), ridge subduction with
steady-state geothermal gradient in continental crust131. the formation of slab windows (Fig. 5f), and mantle plumes. In these

Nature Reviews Earth & Environment


Review article

cases, the thermal effect of enhanced mantle heat flux can be supple- it is possible to achieve UHT metamorphic conditions associated with
mented by heat advected with mantle-derived magmas138,139. However, widespread ductile deformation and igneous activity in arc–backarc
magmatism alone, such as the emplacement of an anorthosite complex, systems. Convergence of active backarcs by lateral wedging of adjacent
generates only local and transient contact UHT metamorphism140,141, cold continents can form an orogenic root143 (Fig. 5b). The conver-
and hence is not discussed here. gence accelerates cooling of the backarc region (Fig. 5g) and generates
counter-clockwise P–T–t paths, with cooling from UHT conditions to
Arc–backarc systems. There is commonly a 200–1,000-km-wide eclogite- or HP granulite-facies conditions143 (Fig. 5h).
uniformly hot and weak backarc region that develops during ocean There is a relatively large number of UHT terranes proposed to
plate subduction15 (Fig. 5a). The Moho temperature of the backarc have been formed in arc–backarc settings (Supplementary Data 1).
region is 800–1,000 °C at a depth of ~35 km, and the surface heat flow In these cases, the reported prograde heating stages are generally
is 60–100 mW m−2 as compared to 30–50 mWm−2 in normal stable isobaric or involve decompression heating, with durations of heating
crust15,142. Such an extremely hot environment could be due to vigor- ≤30 Myr, although the durations of the whole UHT event could be long
ous small-scale mantle convection underneath the thin lithospheric (>70 Myr) owing to slow cooling (Supplementary Data 1). These features
mantle (60–70 km), which can be sustained for several tens of millions are consistent with an arc–backarc system as a common and generally
of years after the termination of subduction15,142 (Fig. 5g). As a result, suitable tectonic setting to generate UHT metamorphism.

a Homogeneous crustal thickening b Crustal channel flow


Pre-thickening Post-thickening Post-erosion
0
T = 0 °C 5 km Thermal properties
Upper crust
A = 3 µW m–3
20 Qs = 71.25 mW m–2
35 km Tsurface = 0 °C

Depth (km)
35 0 0
Lower crust 40 km A1
25
Depth (km)

A = 0 µW m–3 S A2 A1 = 2.0 µW m–3


35

Depth (km)
70 km Vp = 5 cm yr–1 TMoho = 704 °C
Mantle Mantle 20
A = 0 µW m–3
A2 = 0.75 µW m–3

T = 1,300 °C 150 km 35
150 QM = 20 mW m–2

185 km

c Temperature–time paths reaching UHT conditions d Pressure–temperature paths reaching UHT conditions
1,100 20
Duration to reach UHT conditions: 30–75 Myr ite
an
Ky ite
a an
illim 60
900
b 15
S
b
a 0
b
Depth (km)
40
P (kbar)
T (°C)

700 10

10
Homogeneous thickening 30 a UHT
(initial depth at 35 km) 10 b
Homogeneous thickening 20
500 5
(initial depth at 50 km) a
Homogeneous channel flow
Heterogeneous channel flow

300 0
0 20 40 60 80 100 120 140 400 600 800 1,000 1,200
t (Myr) T (°C)

Fig. 4 | UHT metamorphism at thickened continental orogens. a, Model of conditions correspond to those in panel c. Markers are shown at 10-Myr
homogeneous thickening of the crust (by a factor of 2), inspired by ref. 129. intervals on each path, with the number at the first marker indicating millions
Note that the columns are not drawn to scale. b, Initial lithosphere structure of years from the beginning of the experiment. Panels c and d updated from
and properties for each lithosphere layer for crustal channel flow models of large refs. 13,14,129. Properties for these models are provided in the Supplementary
hot orogens with laterally homogenous (HT1) or heterogeneous (GO-series) Note 1. UHT metamorphism caused by radiogenic heating is characterized by
viscosity, simplified from ref. 14. S is the initial model singularity. Q, heat flow; long-lived (30–75 Myr) duration of heating (the period of time from 700 °C to
T, temperature; A, heat production. c, T–t paths that reach UHT metamorphic the temperature peak for each path), and clockwise P–T paths with increasing
conditions based on the models in panels a (homogeneous thickening, paths T during thickening, or isobaric heating followed by decompression heating,
for makers at post-thickening depths of 50 km and 35 km) and b (homogeneous before variable post-peak decompression.
and heterogeneous channel flow). d, P–T paths that achieve UHT metamorphic

Nature Reviews Earth & Environment


Review article

a Hot arc–backarc b Thermal regime of arc–backarc 8 Myr after arc–backarc


before thickening gradual thickening
Potential region to reach Moho
UHT conditions Continent Arc–backarc Continent
20–40 km; 800–1,000 °C
Oceanic Upper
crust Continental crust 400 °C crust
500 °C >700 °C Lower crust
Lithospheric 600 °C
mantle Fluid Lithospheric mantle
700 °C, 35 km Lower crust
Moho Moho

>700 °C
1,350 °C, 1,350 °C,
70 km 70 km
Asthenospheric Asthenospheric
Mantle convection mantle mantle

c Slab rollback d Shallow slab breakoff (< 50 km) e Lithospheric peeling f Ridge subduction or slab window

g T–t paths reaching UHT conditions h P–t paths reaching UHT conditions
1,200 20
Duration to reach UHT conditions: 1.5–10 Myr
ite
an e
Ky anit
m
lli 60
1,000 Si
15

800 b 25
b

Depth (km)
40
P (kbar)

30
T (°C)

10 20
e a
UHT
600

Slow cooling of backarc 0


Thickening of arc–backarc 20
5 e
400 Slab rollback
Lithospheric peeling
f Ridge subduction

200 0
0 20 40 60 80 100 120 140 200 400 600 800 1,000 1,200
t (Myr) T (°C)

Fig. 5 | UHT metamorphism at subduction-related and collisional orogens. models in panels b–f. Updated from refs. 21,143,148. The path for slab rollback
In each model, regions with the potential to generate ultrahigh-temperature corresponds to the path in panel g, but the early evolution before exhumation
(UHT) metamorphism are marked with a red dashed oval. a, Schematic cross- to the lower crust is not shown. The paths for thickening of arc–backarc and
section showing a hot arc–backarc system 16. b, Model of a thickened hot lithospheric peeling correspond to those in panel g. Markers are shown at 10-Myr
arc–backarc143. c, Model of slab rollback. d, Model of shallow slab breakoff 154. intervals on each path, with the number at the first marker indicating millions of
e, Model of lithospheric peeling21. f, Model of ridge subduction with a slab window. years from the beginning of the experiment (except slab rollback, where markers
g, Temperature–time (T–t) paths that reach UHT metamorphic conditions, based are shown at 25 and 30 Myr from the beginning). Properties for these models are
on models in panels a–f. Updated from refs. 15,18,21,143,148. However, note provided in the Supplementary Note 2. UHT metamorphism caused by elevated
that the path for slab rollback is based on a model of continental subduction, mantle heat is characterized by a heating period (the time taken from 700 °C to
not ocean plate subduction as shown in panel c. h, Corresponding pressure– the temperature peak for each path) of 1.5–10 Myr and P–T paths involving close
temperature (P–T) paths that reach UHT metamorphic conditions, based on to isobaric heating and cooling.

Nature Reviews Earth & Environment


Review article

Slab rollback and breakoff. Slab rollback causes extension in the Lithospheric peeling occurs where a weak lower crust (due to a
overriding lithosphere and can induce the formation of a backarc hot geotherm, partial melting or composition) remains attached to the
basin144,145 (Fig. 5c). Rollback can cause rapid heating of the lower crust underlying negatively buoyant and dense lithospheric mantle22,64,161–163
underneath the accretionary region and/or in the extending overriding (Fig. 5e). Numerical models show that peeling can lead to thinning
plate17. Unfortunately, there is no thermomechanical model yet for slab of the overriding lithospheric plate, mantle upwelling and juvenile
rollback during ocean plate subduction to document the detailed P–T magmatism, heating of the thinned orogenic crust, syn-convergence
and T–t paths of any UHT metamorphism. However, the very young extension and topographic uplift.
UHT metamorphism from the island of Seram (about 16 Ma), part Large-scale lithospheric peeling was suggested to be more com-
of the Banda arc, has been interpreted as exhumed during slab roll- mon in the Archaean-to-early Proterozoic, because the mantle was
back77. Here, decompression from ~8 kbar to <7 kbar occurred at UHT hotter (100–250 °C) than today155, and hot and weak lithospheres were
metamorphic conditions, which is consistent with rollback-driven core more common21,22,64,164,165. Numerical geodynamic models show that
complex–style exhumation by extension. The duration of the UHT met- under hotter mantle conditions, convergence between two continen-
amorphic event is <10 Myr, with a period of prograde heating of <5 Myr tal lithospheres can generate a long-lived (ultra-)hot accretionary
(ref. 146). Similarly, UHT metamorphism at about 92 Ma recorded by orogen characterized by a thin crust with hot and laterally extensive
felsic granulite xenoliths in the Kakanui Mineral Breccia, located along flat Moho residing atop hot upwelling mantle (Fig. 5e). The thinning of
the New Zealand margin of Gondwana, has also been explained as a the orogenic crust and mantle upwelling were caused by peeling off
result of slab rollback147. By contrast, some HP and ultrahigh-pressure of the incoming lithospheric mantle (and/or parts of the lower crust).
(UHP) metamorphic rocks preserve beta-shaped P–T paths, indicating The orogenic crust within the peeled-off zone would have high mantle
a two-stage evolution in which deeply subducted continental crust was heat flux, which could have caused UHT metamorphism (900–1,100 °C)
exhumed to the base of the crust before isobaric heating to a UHT peak at mid-to-lower crustal depths22,64,165. The geodynamic models further
and post-peak close-to-isobaric cooling148 (Fig. 5h). These paths have predict that the UHT metamorphic conditions could persist for long
been modelled as due to rollback of a subducting continental plate, periods (>50–150 Myr) owing to the stabilization of the peeling process
which leads to fast exhumation of the deeply subducted rocks to the by the convective mantle flow. However, the prograde heating stages
base of the crust, where asthenospheric upwelling allows the exhumed last for a much shorter time (<15 Myr)20,21,64,165, and the P–T–t paths
HP and UHP metamorphic rocks to be conductively heated to UHT for crustal markers are mostly cooling dominated. As an example,
metamorphic conditions within ~10 Myr (Fig. 5g). the marker tracked in Fig. 5g–h records ~10 Myr of isobaric heating
Slab breakoff could be preceded by and even accelerated by slab to the UHT peak followed by another ~10 Myr of cooling from the peak
rollback148,149. The depth of slab breakoff is strongly controlled by tem- temperature of ~1,000 °C to ~700 °C (ref. 21).
perature, and by the subduction velocity and the age and strength of In summary, lithospheric peeling is a favourable mechanism for
the subducting slab150–152. In most Phanerozoic orogens, slab breakoff the generation of UHT metamorphism in Precambrian terranes, par-
occurs too deep to substantially affect the thermal structure of the ticularly for those with long-lived UHT events and large-scale spatial
overriding lithosphere19,148,150,153,154. However, in the Archaean and early distribution of UHT metamorphism21. The largest long-lived Mesopro-
Palaeoproterozoic the ambient mantle potential temperature was terozoic UHT metamorphic event, which was associated with extensive
probably 100–250 °C hotter than today155. As a result, shallow slab coeval high-temperature magmatism, is related to the Musgrave Orog-
breakoff (that is, breakoff that allows asthenosphere upwelling to eny (1,260–1,110 Ma) in the Musgrave Province, central Australia166,167. It
depths <50–60 km) was more likely in Archaean-to-early Proterozoic has been proposed that this event was caused by lithospheric peeling,
orogens19,151,156. It leads to extension of the overriding plate and astheno- in either an ultrahot orogen or an intraplate setting20,166,167. However, the
spheric upwelling, which could have generated UHT metamorphism19. relationship between lithospheric foundering and UHT metamorphism
Shallow slab breakoff is a plausible model to explain some natural needs to be further investigated using both geodynamic modelling and
examples of UHT metamorphism in Precambrian orogens, such as that data from the geological record.
in the Palaeoproterozoic Helanshan Terrane of the Khondalite Belt of
the North China craton157, in the Palaeoproterozoic Bakhuis Granulite Ridge subduction and slab windows. Ridge subduction (short for
Belt of the Guiana Shield of Surinam158, and in the Mesoproterozoic mid-ocean ridge subduction) can occur on Earth, as exemplified by the
Eastern Ghats Province of India19. Kula–Pacific Ridge, which subducted beneath Japan, and the Farallon–
In summary, slab rollback is a favourable mechanism to generate Pacific Ridge, which subducted beneath western North America, both
UHT metamorphism in the Phanerozoic. By contrast, shallow slab since about 80 Ma (ref. 168). During ridge subduction, an expanding
breakoff was more likely prior to the Neoproterozoic and is a favourable window might form between the two slabs and is filled with upwelling
mechanism to generate UHT metamorphism during the Precambrian. asthenospheric mantle owing to divergence of the two ocean plates169
(Fig. 5f). In an obliquely convergent subduction system, more than one
Lithospheric delamination. Two specific types of lithospheric slab window can form at each site of ridge–trench intersection, as the
delamination are considered here: foundering and peeling. intersection migrates along the plate motion vector170,171.
Lithospheric foundering refers to the removal of a thickened At the initiation of ridge subduction, thermal modelling shows that
cold and dense mantle root due to the negative buoyancy of the litho- temperature increases abruptly in the overriding continental plate,
sphere leading to gravitational instability159–161. Numerical modelling by 50–500 °C, within a few millions of years18,172. Therefore, high-T/P
shows that when an entire mantle lithosphere gets removed via found- metamorphism, and even UHT metamorphism, could occur during
ering, a thermal pulse could generate UHT conditions within ~2 Myr ridge subduction and the formation of one or more slab windows173,174.
at the base of the crust (at 42.6 km depth), although the removal is However, in numerical models, UHT conditions are barely achieved
too rapid for any thermal contribution by conduction of heat into or, if achieved, are limited to within a few tens of kilometres from the
the lower crust161. slab–crust interface, and then only in cases where the subduction

Nature Reviews Earth & Environment


Review article

velocity is rapid (≥6.3 cm yr−1) and the slab age is young (≤10 Myr old)18,175 Implications for metamorphism today
(Fig. 5g). For example, in one model, temperature increased rapidly Geophysical mapping of Moho temperature and depth beneath Ceno-
from ~300 °C to ~1,000 °C within 10 Myr for a subduction velocity of zoic collisional belts and orogens that experienced lithospheric thin-
6.3 cm yr−1, but the thermal pulse sustained UHT conditions for <1 Myr ning and/or removal provides new insight about regions where UHT
(Fig. 5g); therefore, heat advected with magma would be necessary if metamorphism could be occurring currently on the Earth. Using this
UHT conditions were to last for long enough to be registered in the insight combined with appropriate examples of UHT metamorphism,
crust18. However, these models did not consider the thermal effect of a synoptic view of two plausible tectonic settings for generating UHT
the formation of one or more slab windows, which could enhance the metamorphic conditions is offered: first, beneath thickened crust,
thermal perturbation. particularly orogenic plateaus, where the major source of heating was
One example of a short-lived (~2–3 Myr) amphibolite- to granulite- radiogenic (and/or elevated mantle heat); second, in an arc–backarc
facies metamorphic event that has been related to ridge subduction and/or a thinned orogen, where each setting was dominated by mantle
and/or a slab window is the Chugach metamorphic complex of south- heat flux.
ern Alaska176. However, there are no examples of UHT metamorphism
demonstrably related to ridge subduction. This mechanism remains The Tibetan Plateau example
speculative until examples of UHT metamorphism explicitly related It has been proposed that the core of the large orogenic plateau in Tibet
to ridge subduction and the formation of one or more slab windows is currently experiencing ongoing UHT metamorphism49,50. Based on
are demonstrated and further investigated by numerical modelling. the Moho temperature and depth in central–eastern Tibet45, UHT meta-
morphism could be occurring currently under the Qilian orogenic belt
Mantle plumes. Mantle plumes, which are thought to originate from and the Songpan–Ganzi Terrane, beneath the eastern Lhasa Terrane and
the mantle transition zone and/or the core–mantle boundary, cause at depth in the southern Chuandian Block (Fig. 6a–b). This potential
short-lived episodes of extensive flood basalt volcanism, and generate UHT metamorphism occurs at depths between 40 and 75 km, with that
volcanic plateaus or hotspot tracks177,178. A plume head is predicted beneath the southern Chuandian Block occurring at the shallowest
to be 1,500–2,500 km across and 100–200 km thick as it spreads level (40–55 km), and that beneath the eastern Lhasa Terrane occurring
beneath the lithosphere, which is important for crustal reworking, at the deepest level (55–75 km) (Fig. 6c). These sites of potential UHT
high-grade metamorphism and partial melting at a regional scale177. metamorphism are in close association with the potential occurrence
Mantle plumes have high mantle potential temperatures, in the range of eclogite- to HP granulite-facies metamorphism, because the Tibetan
of 1,450–1,600 °C, and are generally 100–250 °C hotter than the ambi- crust is generally at least 60–70 km thick (Fig. 6c). Moreover, Moho
ent mantle178,179. A plume head event can last for several to tens of temperatures are positively correlated with Moho depths, suggesting
millions of years, and the plume tail can persist for up to 200 Myr177, that radiogenic heating and crustal thickening are the driving forces
where the duration in any one locality in the crust is controlled to first to achieve UHT metamorphic conditions.
order by the rate of the overlying plate motion. A hot mantle plume Nevertheless, the uplift of Tibet, the east–west extensional struc-
head might generate UHT metamorphism, considering the abundant tures and the extensive Cenozoic volcanism have all been explained by
mantle-derived magma generated by a plume, in which case the timing post-thickening lithospheric delamination182,183, which is supported
of any UHT metamorphism should coincide with the formation of a by geochemical signatures of the magmatic rocks184–186, helium iso-
large igneous province. tope data from geothermal springs187, numerous seismic tomographic
The deep crustal granulite xenoliths brought up in kimberlites studies188–192 and numerical modelling193,194. The potential UHT meta-
in the central Kaapvaal craton, South Africa, experienced short-lived morphism beneath the eastern Lhasa Terrane, in the core of the cen-
(5–10 Myr) UHT metamorphism at about 2,720 Ma180,181. The timing and tral Tibetan Plateau, and also that beneath the Qilian orogen and the
duration of UHT metamorphism are consistent with the Ventersdorp Songpan–Ganzi Terrane, at the northeastern margin of the plateau,
flood basalt magmatism180,181. As a result, the UHT metamorphism associated with extensive volcanism in western–central Tibet, could
could have been generated by a mantle plume. However, no correla- have been situated in a continental arc–backarc system due to the
tion between the temporal distribution of UHT metamorphism and northward subduction of the Indian plate, and have probably also been
large igneous province and/or komatiite magmatism has yet been affected by an elevated mantle heat flux due to lithospheric delamina-
established. Therefore, more evidence from both numerical modelling tion195. For comparison, the potential UHT metamorphism beneath the
and other natural examples is required to further document a possible southern Chuandian Block at the eastern edge of the plateau, coupled
relationship between mantle plumes and UHT metamorphism. with the Tengchong volcano to the southwest of this block, might be
Radiogenic heating alone is unlikely to generate UHT metamor- related to a continental arc–backarc system generated by the eastward
phism unless the crustal heat production is substantially higher than subduction of the Indian plate beneath the Burmese arc195–197, and
average, and an elevated mantle heat flux represents the more general probably also affected by lithospheric delamination198,199.
case to achieve UHT conditions. Given a high enough crustal heat pro- Cenozoic UHT metamorphism (at 25–15 Ma) overprinting prior
duction and low erosion rates, sufficient time for radiogenic heating eclogite-facies metamorphism (at about 30 Ma) by decompression
during thickening UHT metamorphism can be achieved in an orogenic heating has been reported from the central Himalaya200–202. It has been
plateau maintained by ongoing plate collision. Also, orogenic thickening proposed that partial melting of the crust and the early eclogite-facies
followed by collapse during extension could lead to UHT overprints on metamorphism were related to thickening of the crust and radiogenic
earlier eclogite- or HP granulite-facies metamorphism. In the Phanero- heating200,203, while UHT metamorphism was caused by elevated mantle
zoic, arc–backarc systems and/or thinned orogens due to slab rollback heat flux due to lithospheric thinning200. In northern Qiangtang, UHT
probably generated UHT metamorphic conditions. By contrast, in the metamorphism at about 2.3 Ma is associated with deeply sourced
Precambrian, shallow slab breakoff or lithospheric peeling during plate dacites. Here, the driving mechanism for metamorphism and crustal
convergence could have generated UHT metamorphic conditions. melting is thought to have been 20–40 Myr of radiogenic heating in

Nature Reviews Earth & Environment


Review article

a Longitude (E) d Longitude (W)


90° 95° 100° 105° 110° 125° 120° 115° 110° 105° 100° 95° 90°
40° 50°
Central–eastern Tibet Western United States

des
Qili Ordos
an o

Casca
rog Potential region for
en
UHT metamorphism
Qaidam 45° Sna Yellowstone volcano
Riv ke
er P
35° lain
So AKM Wyoming
JR ng S craton
S pa
Qiangtang

Sie
n-G Basin and Southern
Latitude (N)

40°
an

rra
zi Qinling Range Rocky Mountains
BNS

Ne
vad
Colorado Great Plains
Sichuan

a
30° Lhasa basin 35° Plateau
Rio Grande Rift
IYS
Himalaya Yangtze
EHS Chuandian
Mogok block
30° Basin and
metamorphic
belt Range
25°
Tengchong volcano
25°

30 40 50 Moho depth (km) 70 80 90 20 30 Moho depth (km) 40 50

90° 95° 100° 105° 110° 125° 120° 115° 110° 105° 100° 95° 90°
b 40° e 50°
Central–eastern Tibet Western United States

45°

35°
Latitude (N)

40°

30° 35°

30°

25°
25°

400 600 Moho temperature (°C) 1,000 1,200 500 600 Moho temperature (°C) 900 1,000

c 30
UHP
f 30
Central–eastern Tibet Western United States
E-HPG
100 100
site
tz

Forbidden Coe tz Eastern


te uar

r
zone Qua Lhasa
Al + Q

ite
ite
bi

an ite 80 80
de
Wet solidus

Ky an
Ja

20 lim 20
Si
Depth (km)

Depth (km)
P (kbar)

P (kbar)

60 60
BI

40 40
10 Southern 10
Chuandian block

UHT 20 20
Am Gn
Lg
Gr
0 0
200 400 600 800 1,000 1,200 1,400 200 400 600 800 1,000 1,200 1,400
T (°C) T (°C)

Nature Reviews Earth & Environment


Review article

Fig. 6 | Thermal regime at the Moho beneath central–eastern Tibet e, Moho temperature (°C) beneath the western United States. f, P–T conditions
and the western United States. Moho temperatures higher than 900 °C at Moho depth beneath the western United States. The purple colour in panels c
in the sillimanite-stable field are marked by solid yellow in panels a,b,d & e, and f represents all collected Moho temperature and pressure datasets in the
representing likely locations where ultrahigh-temperature (UHT) metamorphism regions of respective panels a and d. AKMS, Animaque–Kunlun Mountain
could be occurring at the present day. a, Moho depth (km) beneath central– suture; BNS, Bangong–Nujiang suture; EHS, eastern Himalayan syntaxis; IYS,
eastern Tibet, a representative collisional orogen. b, Moho temperature (°C) Indo-Yarlung suture, JRS, Jinsha River suture; Lg, low-grade metamorphism.
beneath central–eastern Tibet. c, Pressure–temperature (P–T) conditions at Panels a–b adapted with permission from ref. 45, Elsevier. Panel c produced from
Moho depth beneath central–eastern Tibet. P–T conditions of potential UHT data in ref. 45; Panels d–f produced from data in ref. 46. At present on Earth, UHT
metamorphism beneath Eastern Lhasa are shown in light blue, those at depth in metamorphism is associated with convergent plate margins, either in continental
the southern Chuandian Block in red, and the remainder in yellow. d, Moho depth collisional belts or associated with subduction-related extensional orogens.
(km) beneath the western United States, a representative extensional orogen.

felsic Tibetan crust ≥60 km thick with heat production of 2.5 μW m−3 metamorphism (Fig. 6d), which is not limited to the plume-affected
(ref. 204). The occurrence of this young UHT metamorphism is compa- region, the sinking Farallon slab would have influenced a large part of
rable to the ongoing UHT metamorphism in the core of Tibet. Moreo- the western United States.
ver, UHT metamorphism occurring at about 25 Ma was identified from Large-scale active UHT metamorphism was postulated to be occur-
the Mogok metamorphic belt in Myanmar205. The Mogok metamorphic ring beneath the southern segment of the Basin and Range42,216,217, which
belt extends over 1,500 km along the western margin of the Shan–Thai could be caused by lithospheric extension after Laramide orogeny42.
Block, and from the Andaman Sea north to the eastern Himalayan syn- Granulite xenoliths from Prindle volcano, Alaska, recorded UHT meta-
taxis206; the Tengchong volcano, which has erupted since the late Mio- morphism at 63–48 Ma, indicating similar lithospheric thinning in the
cene, is located in the north of this metamorphic belt207 (Fig. 6a). The Canadian Cordillera218.
mechanism to generate Cenozoic UHT metamorphism in the Mogok To summarize, the dominant mechanism generating UHT meta-
metamorphic belt is comparable to that for the potential ongoing UHT morphic conditions in the western United States is most likely to be
metamorphism beneath the southern Chuandian Block. the lithospheric extension due to the detachment of the subducted
To summarize, potential UHT metamorphism in central–eastern Farallon slab.
Tibet is likely to be a result of the combined effects of radiogenic heat-
ing in thickened crust and elevated mantle heat flux due to lithospheric Summary and future perspectives
thinning by delamination. To conclude, the most likely tectonic settings to generate UHT meta-
morphism are at depth in arc–backarc systems, in continental col-
The Cordilleran example lisional belts, and in thinned orogens associated with lithospheric
Moho temperatures in the North American Cordillera, which rep- extension. In arc–backarc settings, UHT metamorphism is character-
resents a modern arc–backarc system associated with subduction ized by isobaric or decompression heating and then cooling-dominated
of the Farallon and Juan de Fuca plates, are commonly in the range P–T paths, and short durations of heating (≤30 Myr). Metamorphism
800 to 850 °C, and even reach UHT conditions, whereas Moho depths is driven by elevated mantle heat flux beneath thinned lithosphere,
are shallow, in the range 30 to 35 km (refs. 46,142). UHT metamorphism probably due to slab rollback in Phanerozoic orogens, where deep
could be occurring under the Sierra Nevada volcanic arc and the eastern subduction can occur. In continental collisional belts, eclogite- or HP
Snake River Plain, along the western edge of the Colorado Plateau, granulite-facies metamorphism commonly can be followed by UHT
and to the east of the Rio Grande Rift, areas that are closely associated metamorphism that was generated by decompression heating. Thus,
with the occurrence of volcanism since about 30 Ma46,208 (Fig. 6d,e). the metamorphic event is characterized by a clockwise P–T path. Such
Moho temperatures under the Cordillera are independent of the Moho UHT metamorphism is most probably caused by a combination of
depths (Fig. 6f), which contrasts with the case of the Tibetan orogenic radiogenic heating in thickened crust and then elevated mantle heat
plateau (Fig. 6c). The lithosphere is thin, in the range 60–70 km, and the flux during decompression. Alternatively, UHT metamorphism can be
elevated mantle heat flux was responsible for generating the UHT meta- generated in continental crust that has unusually high heat production.
morphism and the extensive volcanism. However, the mechanism(s) In these cases, the duration of heating is generally longer (≥30 Myr) and
that caused the lithospheric thinning and/or asthenospheric upwelling the P–T paths are commonly decompression dominated. In thinned
remain controversial. and hot Precambrian orogens, UHT metamorphism was probably caused
In one proposal, following subduction of the Farallon plate, the by lithospheric peeling or shallow slab breakoff during plate tectonics.
Columbia River flood basalts, the Yellowstone hotspot and a volcanic Overall, radiogenic heating is probably seldom the sole cause of UHT
track in the Snake River Plain were thought to be related to a mantle metamorphism, while mantle heat is likely to be dominant in exten-
plume starting at about 17 Ma209,210. However, this interpretation might sional settings where lithospheric mantle is thinned and/or delami-
be an over-simplification, owing to plume–slab interactions in the nated. Since the Archaean, the mechanisms of UHT metamorphism
region, and it has also been argued that the plume activity was related have been closely related to the supercontinent cycle and tectonic
to complexities in the subduction-driven mantle flow generated by processes at convergent plate margins, which have evolved over time.
ridge subduction and the formation of multiple slab windows, and the Numerical models designed to investigate the generation of UHT
detachment of the Farallon slab211–215. Moreover, the latter case can also metamorphism are generally limited in number and scope. In terms
explain the uplift of the Colorado Plateau and the extensive volcanism of numerical modelling of specific tectonic settings, sometimes the
younger than about 30 Ma throughout the western United States208. metamorphic evolution has not been fully investigated, and no quan-
From the perspective of the spatial distribution of potential UHT titative correlation between a viable heat source and spatial extent

Nature Reviews Earth & Environment


Review article

Glossary

Accretionary region Lithospheric foundering Prograde heating stage Stable continental shields
A region at the convergent plate margin A type of delamination in which a cold The period during which temperature The large stable areas of low relief in the
between a subducting ocean plate and and dense mantle lithosphere root generally increases with time; in the Earth’s continents that are composed
overriding continental plate, generally founders into the hot mantle beneath. case of UHT metamorphism, the of Precambrian crystalline igneous
including forearc, magmatic arc and duration is measured from ~700 °C and high-grade metamorphic rocks.
backarc components. Lithospheric peeling to the peak temperature.
A type of delamination in which the Steady-state
Channel flow dense lithospheric mantle and/or Radiogenic heating A condition in which the variables that
Protracted flow of a weak, viscous portions of the lower continental The generation of heat via radioactive define the behaviour of the system or
crustal layer between relatively rigid yet crust peel off from the mid-to-upper decay during the production of the process do not change over time.
deformable bounding crustal layers. continental crust along a plane that radiogenic nuclides.
runs parallel to the Moho. Supercontinent
Closure temperature Retrograde reaction A single large landmass that includes
During cooling, the temperature at Mantle convection A metamorphic reaction that occurs all or most of the existing continents
which there is no longer any substantial The movement of the mantle after the metamorphic peak during the on the Earth.
diffusion of the parent or daughter as it transfers heat from the interior decrease of temperature with time.
isotopes out of the system; this to Earth’s surface. Supercratons
temperature is dependent on cooling Ridge subduction Large late Archaean landmasses before
rate and grain size. Mantle plumes The subduction of a mid-ocean ridge the supercontinent cycle since about
A mechanism of convection within the beneath the trench at a convergent 2,200 Ma.
Geothermal gradient Earth’s mantle; hypothesized to explain boundary.
The rate of temperature change with anomalous volcanism. Terrane
respect to increasing depth in the Slab breakoff A fault-bounded block containing
Earth’s interior. Mantle potential temperature A process whereby the oceanic a group of related rocks that have a
The temperature the mantle would lithosphere detaches from the distinct geological history compared
Incompatible have at Earth’s surface if extrapolated continental lithosphere as subduction with contiguous blocks.
Incompatible elements are those that along an adiabat without melting. locks up during continental collision.
prefer to be in the liquid phase rather Ultrahigh-temperature (UHT)
than the solid phase during melting. Petrochronology Slab rollback metamorphism
The discipline that links geochronology The trench migrates in the direction An extreme subset of granulite-
Laramide orogeny with petrology. opposite to the plate motion of the facies metamorphism that occurs at
Mountain building in western North subducting slab, also known as trench temperatures above the breakdown
America, starting at 70–80 Ma and Phase equilibrium modelling retreat. of biotite in metapelitic rocks, with peak
ending at 35–55 Ma. The forward modelling of phase temperatures that exceed 900 °C in the
equilibria involving solid solutions Slab windows sillimanite stability field.
Lithospheric delamination in complex systems either by direct Gaps that form in a subducting ocean
Detachment of the continental mantle minimization of Gibbs energy or by plate when a mid-ocean ridge is Viscous dissipation
lithosphere and/or a portion of the solution of simultaneous nonlinear subducted. The irreversible process by means
lowermost continental crust. equations. of which the work done by deformation
is transformed into heat.

of any UHT metamorphism has been established. Mechanisms such We encourage the development of new ion mapping approches to
as lithospheric foundering, slab rollback, ridge subduction and the identify and avoid problems associated with isotopic resetting at
formation of multiple slab windows (particularly for the generation of UHT conditions and decoupling of zircon U–Pb dates from their trace
Phanerozoic UHT metamorphism), and mantle plumes (particularly for element compositions. An increased number of natural examples with
Archaean UHT metamorphism) demand a wider range of systematic better constrained P–T–t paths and estimates of metamorphic dura-
thermomechanical modelling. tions is required for comparison with results from numerical modelling,
For the future, we encourage comprehensive investigations of particularly to investigate secular change in UHT metamorphism in
UHT metamorphic localities in terms of petrology, multimineral petro- relation to geodynamic models.
chronology and phase equilibrium modelling to better determine the The processes that generate UHT metamorphism are closely
peak P–T conditions, P–T–t paths and duration of UHT metamorphic correlated with the formation and evolution of continental crust,
events, especially the duration of the suprasolidus prograde heating changes of tectonic mode and the evolution of plate tectonics. Addi-
stage. We acknowledge that for UHT rocks in particular, the petrochro- tional comparisons between Phanerozoic and Precambrian examples
nology required to constrain the duration of the event and different of UHT metamorphism would deepen understanding of the thermal
stages of the evolution is the most challenging part of the research. evolution of the lithosphere and geodynamic processes during the

Nature Reviews Earth & Environment


Review article

transition to plate tectonics. For example, global mapping of Moho 7. Harley, S. L. in Encyclopedia of Geology 2nd edition (eds Alderton, D. & Elias, S. A.)
522–552 (Elsevier, 2021).
depth and temperature using geophysical datasets to determine A comprehensive guidebook to the study of UHT metamorphism, especially for
the spatial distribution of potential sites of UHT metamorphism at the students and beginners.
present day will provide an overview of relationships with other types 8. Kelsey, D. E. On ultrahigh-temperature crustal metamorphism. Gondwana Res. 13, 1–29
(2008).
of metamorphism and/or magmatic events, yielding new insights into 9. Lei, H. C. & Xu, H. J. A review of ultrahigh temperature metamorphism. J. Earth Sci. 29,
favourable tectonic settings and geodynamic processes that gener- 1167–1180 (2018).
ate UHT metamorphism. For ancient examples, whether there are 10. Sengupta, P. et al. Petro-tectonic imprints in the sapphirine granulites from Anantagiri,
Eastern Ghats Mobile Belt, India. J. Petrol. 31, 971–996 (1990).
occurrences of UHT metamorphism older than about 3,100 Ma should 11. Brown, M. & Johnson, T. Time’s arrow, time’s cycle: granulite metamorphism and
be explored. Systematic comparisons of possible ancient examples geodynamics. Miner. Mag. 83, 323–338 (2019).
with geodynamic models will constrain the dominant tectonic mode 12. Beaumont, C., Jamieson, R. A., Nguyen, M. H. & Medvedev, S. Crustal channel flows:
1. Numerical models with applications to the tectonics of the Himalayan–Tibetan orogen.
before the transition to plate tectonics. The role of mantle plumes in J. Geophys. Res. Solid Earth https://doi.org/10.1029/2003jb002809 (2004).
generating UHT metamorphism in the Archaean should be addressed, 13. Jamieson, R. A., Beaumont, C., Medvedev, S. & Nguyen, M. H. Crustal channel flows:
2. Numerical models with implications for metamorphism in the Himalayan–Tibetan
because plumes are widely considered to have been responsible for
orogen. J. Geophys. Res. Solid Earth https://doi.org/10.1029/2003jb002811 (2004).
the numerous extremely hot komatiites that are characteristic of the This was one of the first articles to report results of modelling the formation of a
Archaean219. Himalaya-type large hot orogen and to document channel flow in the mid-crust
underneath the Tibetan Plateau.
Lastly, the unique nature of the lithospheric conditions and
14. Jamieson, R. A. & Beaumont, C. Coeval thrusting and extension during lower crustal
tectonics recorded by mid-Proterozoic orogens should be better ductile flow — implications for exhumation of high-grade metamorphic rocks.
explored. These orogens record the highest T/P and longest durations J. Metamorph. Geol. 29, 33–51 (2011).
15. Currie, C. A. & Hyndman, R. D. The thermal structure of subduction zone back arcs.
of UHT metamorphism, as well as the lowest post-peak cooling rates
J. Geophys. Res. Solid Earth https://doi.org/10.1029/2005jb004024 (2006).
compared with those formed before and after this period. Several 16. Hyndman, R. D. Mountain building orogeny in precollision hot backarcs: North American
conceptual models have been proposed to explain the tectonic regime Cordillera, India-Tibet, and Grenville Province. J. Geophys. Res. Solid Earth 124, 2057–2079
(2019).
during the mid-Proterozoic, such as a plate slowdown or a shutoff of 17. Collins, W. J. Hot orogens, tectonic switching, and creation of continental crust. Geology
plate tectonics, or single-lid tectonics109,110,114,115, but as yet, there is no 30, 535–538 (2002).
consensus. 18. Iwamori, H. Thermal effects of ridge subduction and its implications for the origin of granitic
batholith and paired metamorphic belts. Earth Planet. Sci. Lett. 181, 131–144 (2000).
Overall, future research on UHT metamorphism and its links with 19. Sizova, E., Gerya, T. & Brown, M. Contrasting styles of Phanerozoic and Precambrian
the thermal evolution of the lithosphere and the secular evolution of continental collision. Gondwana Res. 25, 522–545 (2014).
plate tectonics demands a comprehensive multidisciplinary investiga- This paper highlighted the difference between the styles of Phanerozoic and
Precambrian continental collision due to secular mantle cooling, and proposed
tion. The combination of metamorphic petrology with palaeogeog- that shallow slab breakoff during the Precambrian precluded the formation of
raphy, various geophysical datasets and geodynamic modelling will deeply subducted rocks.
lead the way. 20. Gorczyk, W., Smithies, H., Korhonen, F., Howard, H. & De Gromard, R. Q. Ultra-hot
Mesoproterozoic evolution of intracontinental central Australia. Geosci. Front. 6,
23–37 (2015).
Data availability 21. Chowdhury, P., Chakraborty, S. & Gerya, T. V. Time will tell: secular change in
The data for Figs. 1 and 2 are available in the online Supplementary metamorphic timescales and the tectonic implications. Gondwana Res. 93, 291–310
(2021).
Data file. This paper discussed a valid model (lithospheric peeling underneath a thinned orogen)
to explain Archaean UHT metamorphism and juvenile magmatism when the mantle
Published online: xx xx xxxx was warmer than today.
22. Chowdhury, P., Chakraborty, S., Gerya, T. V., Cawood, P. A. & Capitanio, F. A. Peel-back
controlled lithospheric convergence explains the secular transitions in Archean
References metamorphism and magmatism. Earth Planet Sci. Lett. https://doi.org/10.1016/
1. Harley, S. L. in Geology Society Special Publications Vol. 138 (eds Treloar, P. J. j.epsl.2020.116224 (2020).
& O’Brien, P. J.) 81–107 (Geological Society of London, 1998). 23. Kroll, H., Evangelakakis, C. & Voll, G. Two-feldspar geothermometry — a review and
This article defined UHT metamorphism and recognized two types of post-peak revision for slowly cooled rocks. Contrib. Mineral. Petrol. 114, 510–518 (1993).
pressure–temperature (P–T) path, which are close to isobaric cooling and close 24. Benisek, A., Kroll, H. & Cemic, L. New developments in two-feldspar thermometry.
to isothermal decompression. Am. Mineral. 89, 1496–1504 (2004).
2. Brown, M. Metamorphic conditions in orogenic belts: a record of secular change. 25. Zack, T., Moraes, R. & Kronz, A. Temperature dependence of Zr in rutile: empirical
Int. Geol. Rev. 49, 193–234 (2007). calibration of a rutile thermometer. Contrib. Mineral. Petrol. 148, 471–488 (2004).
This was one of the first articles to document the secular change of metamorphism 26. Watson, E. B., Wark, D. A. & Thomas, J. B. Crystallization thermometers for zircon and
through Earth history, argue that the occurrence of UHT metamorphism is rutile. Contrib. Mineral. Petrol. 151, 413–433 (2006).
closely related to the supercontinent cycle and propose the generation of UHT 27. Ferry, J. M. & Watson, E. B. New thermodynamic models and revised calibrations for
metamorphism in backarc settings. the Ti-in-zircon and Zr-in-rutile thermometers. Contrib. Mineral. Petrol. 154, 429–437
3. Harley, S. L. Refining the P–T records of UHT crustal metamorphism. J. Metamorph. Geol. (2007).
26, 125–154 (2008). 28. Tomkins, H. S., Powell, R. & Ellis, D. J. The pressure dependence of the zirconium-in-rutile
4. Kelsey, D. E. & Hand, M. On ultrahigh temperature crustal metamorphism: phase thermometer. J. Metamorph. Geol. 25, 703–713 (2007).
equilibria, trace element thermometry, bulk composition, heat sources, timescales 29. Holland, T. J. B. & Powell, R. An internally consistent thermodynamic data set for phases
and tectonic settings. Geosci. Front. 6, 311–356 (2015). of petrological interest. J. Metamorph. Geol. 16, 309–343 (1998).
This paper provided a systematic review of UHT metamorphism, summarizing major 30. Kelsey, D. E., White, R. W., Holland, T. J. B. & Powell, R. Calculated phase equilibria in
advances including the quantification of UHT conditions by phase equilibrium modelling K2O-FeO-MgO-Al2O3-SiO2-H2O for sapphirine-quartz-bearing mineral assemblages.
and trace element thermometry, linking metamorphic ages with P–T points/paths by J. Metamorph. Geol. 22, 559–578 (2004).
using HREE partitioning, and providing insight into the heat sources and tectonic settings. 31. Powell, R. & Holland, T. J. B. On thermobarometry. J. Metamorph. Geol. 26, 155–179
5. White, R. W. & Powell, R. Melt loss and the preservation of granulite facies mineral (2008).
assemblages. J. Metamorph. Geol. 20, 621–632 (2002). 32. Powell, R., White, R. W., Green, E. C. R., Holland, T. J. B. & Diener, J. F. A. On parameterizing
6. Brown, M. & Johnson, T. Secular change in metamorphism and the onset of global plate thermodynamic descriptions of minerals for petrological calculations. J. Metamorph. Geol.
tectonics. Am. Miner. 103, 181–196 (2018). 32, 245–260 (2014).
This work established the relationship between changes in thermal gradients evident 33. Rubatto, D. Zircon trace element geochemistry: partitioning with garnet and the link
in the metamorphic rock record and the evolution of geodynamic regimes, and between U–Pb ages and metamorphism. Chem. Geol. 184, 123–138 (2002).
proposed that the emergence of plate tectonics began during the late Mesoarchaean, 34. Whitehouse, M. J. & Platt, J. P. Dating high-grade metamorphism — constraints from
based on the widespread record of metamorphism since that time. rare-earth elements in zircon and garnet. Contrib. Mineral. Petrol. 145, 61–74 (2003).

Nature Reviews Earth & Environment


Review article

35. Kohn, M. J. & Malloy, M. A. Formation of monazite via prograde metamorphic reactions 63. Brown, M., Johnson, T. & Gardiner, N. J. Plate tectonics and the Archean Earth. Annu. Rev.
among common silicates: implications for age determinations. Geochim. Cosmochim. Acta Earth Planet. Sci. 48, 291–320 (2020).
68, 101–113 (2004). 64. Chowdhury, P., Gerya, T. & Chakraborty, S. Emergence of silicic continents as the lower
36. Kelly, N. M. & Harley, S. L. An integrated microtextural and chemical approach to zircon crust peels off on a hot plate-tectonic Earth. Nat. Geosci. 10, 698–703 (2017).
geochronology: refining the Archaean history of the Napier Complex, East Antarctica. 65. Bédard, J. H. A catalytic delamination-driven model for coupled genesis of Archaean
Contrib. Mineral. Petrol. 149, 57–84 (2005). crust and sub-continental lithospheric mantle. Geochim. Cosmochim. Acta 70, 1188–1214
37. Kelly, N. M., Clarke, G. L. & Harley, S. L. Monazite behaviour and age significance in (2006).
poly-metamorphic high-grade terrains: a case study from the western Musgrave Block, 66. Johnson, T. E., Brown, M., Kaus, B. J. P. & VanTongeren, J. A. Delamination and recycling
central Australia. Lithos 88, 100–134 (2006). of Archaean crust caused by gravitational instabilities. Nat. Geosci. 7, 47–52 (2014).
38. Guevara, V. E. et al. Polyphase zircon growth during slow cooling from ultrahigh 67. Gerya, T. Precambrian geodynamics: concepts and models. Gondwana Res. 25, 442–463
temperature: an example from the Archean Pikwitonei granulite domain. J. Petrol. (2014).
https://doi.org/10.1093/petrology/egaa021 (2020). 68. Rozel, A. B., Golabek, G. J., Jain, C., Tackley, P. J. & Gerya, T. Continental crust formation
39. Jiao, S. et al. Texturally-controlled U–Th–Pb monazite geochronology reveals on early Earth controlled by intrusive magmatism. Nature 545, 332–335 (2017).
Paleoproterozoic UHT metamorphic evolution in the Khondalite Belt, North China 69. Hokada, T. Feldspar thermometry in ultrahigh-temperature metamorphic rocks:
craton. J. Petrol. https://doi.org/10.1093/petrology/egaa023 (2020). evidence of crustal metamorphism attaining ~1100 °C in the Archean Napier Complex,
40. Jiao, S. J. et al. The timing and duration of high-temperature to ultrahigh-temperature East Antarctica. Am. Miner. 86, 932–938 (2001).
metamorphism constrained by zircon U–Pb–Hf and trace element signatures in the 70. Jiao, S. J. & Guo, J. H. Application of the two-feldspar geothermometer to ultrahigh-
Khondalite Belt, North China craton. Contrib. Mineral. Petrol. https://doi.org/10.1007/ temperature (UHT) rocks in the Khondalite belt, North China craton and its implications.
s00410-020-01706-z (2020). Am. Mineral. 96, 250–260 (2011).
41. Durgalakshmi et al. The timing, duration and conditions of UHT metamorphism in remnants 71. Benbatta, A. et al. Ternary feldspar thermometry of Paleoproterozoic granulites from
of the former eastern Gondwana. J. Petrol. https://doi.org/10.1093/petrology/egab068 In-Ouzzal terrane (Western Hoggar, southern Algeria). J. Afr. Earth Sci. 127, 51–61 (2017).
(2021). 72. Korhonen, F. J., Brown, M., Clark, C. & Bhattacharya, S. Osumilite-melt interactions in
42. Cipar, J. H., Garber, J. M., Kylander-Clark, A. R. C. & Smye, A. J. Active crustal differentiation ultrahigh temperature granulites: phase equilibria modelling and implications for the
beneath the Rio Grande Rift. Nat. Geosci. 13, 758–763 (2020). P–T–t evolution of the Eastern Ghats Province, India. J. Metamorph. Geol. 31, 881–907
This work documented the mechanism of crustal differentiation and UHT metamorphism (2013).
by combining petrological investigation with geophysical mapping and numerical 73. Holder, R. M., Hacker, B. R., Horton, F. & Rakotondrazafy, A. F. M. Ultrahigh-temperature
modelling. osumilite gneisses in southern Madagascar record combined heat advection and high
43. Wyatt, D. C., Smye, A. J., Garber, J. M. & Hacker, B. R. Assembly and tectonic evolution of rates of radiogenic heat production in a long-lived high-T orogen. J. Metamorph. Geol.
continental lower crust: monazite petrochronology of the Ivrea-Verbano zone (Val Strona 36, 855–880 (2018).
di Omegna). Tectonics 41, e2021TC006841 (2022). 74. Ouzegane, K., Guiraud, M. & Kienast, J. R. Prograde and retrograde evolution in high-
44. Cenki, B., Rey, P. F., Arcay, D. & Giordani, J. Timing of partial melting and granulite temperature corundum granulites (FMAS and KFMASH systems) from In Ouzzal terrane
formation during the genesis of high to ultra‐high temperature terranes: insight from (NW Hoggar, Algeria). J. Petrol. 44, 517–545 (2003).
numerical experiments. Terra Nova 34, 193–200 (2022). 75. Osanai, Y. et al. Permo-Triassic ultrahigh-temperature metamorphism in the Kontum
45. Li, L., Zhang, X. Z., Liao, J., Liang, Y. L. & Dong, S. X. Geophysical constraints on the nature massif, central Vietnam. J. Mineral. Petrol. Sci. 99, 225–241 (2004).
of lithosphere in central and eastern Tibetan plateau. Tectonophysics https://doi.org/ 76. Sajeev, K., Osanai, Y. & Santosh, M. Ultrahigh-temperature metamorphism followed
10.1016/j.tecto.2021.228722 (2021). by two-stage decompression of garnet-orthopyroxene-sillimanite granulites from
46. Schutt, D. L., Lowry, A. R. & Buehler, J. S. Moho temperature and mobility of lower crust Ganguvarpatti, Madurai Block, southern India. Contrib. Minerol. Petrol. 148, 29–46
in the western United States. Geology 46, 219–222 (2018). (2004).
47. Harley, S. L. A matter of time: the importance of the duration of UHT metamorphism. 77. Pownall, J. M., Hall, R., Armstrong, R. A. & Forster, M. A. Earth’s youngest known ultrahigh-
J. Min. Pet. Sci. 111, 50–72 (2016). temperature granulites discovered on Seram, eastern Indonesia. Geology 42, 279–282
This paper highlighted the importance of metamorphic duration, distinguished (2014).
short-lived from long-lived UHT metamorphism and proposed different mechanisms 78. Jons, N., Schenk, V., Appel, P. & Razakamanana, T. Two-stage metamorphic evolution of
accordingly. the Bemarivo Belt of northern Madagascar: constraints from reaction textures and in situ
48. Liu, Y. B., Mitchell, R. N., Brown, M., Johnson, T. E. & Pisarevsky, S. Linking metamorphism monazite dating. J. Metamorph. Geol. 24, 329–347 (2006).
and plate boundaries over the past 2 billion years. Geology 50, 631–635 (2022). 79. Tsunogae, T. & Van Reenen, D. D. High-pressure and ultrahigh-temperature granulite-
An important demonstration of the spatiotemporal correlation between different types facies metamorphism of Precambiran high-grade terranes: case study of the Limpopo
of metamorphism and convergent plate boundaries as a function of distance from the Complex. Geol. Soc. Am. Mem. 207, 107–124 (2011).
trench, with high T/P metamorphism generally occurring in the orogenic hinterland. 80. Dharmapriya, P. L. et al. Hybrid phase equilibria modelling with conventional and trace
49. Hacker, B. R. et al. Hot and dry deep crustal xenoliths from Tibet. Science 287, 2463–2466 element thermobarometry to assess the P–T evolution of UHT granulites: an example
(2000). from the Highland Complex, Sri Lanka. J. Metamorph. Geol. 39, 209–246 (2021).
50. Hacker, B. R., Ritzwoller, M. H. & Xie, J. Partially melted, mica-bearing crust in Central Tibet. 81. Holdaway, M. J. Recalibration of the GASP geobarometer in light of recent garnet
Tectonics 33, 1408–1424 (2014). and plagioclase activity models and versions of the garnet-biotite geothermometer.
51. Palin, R. M., Weller, O. M., Waters, D. J. & Dyck, B. Quantifying geological uncertainty in Am. Mineral. 86, 1117–1129 (2001).
metamorphic phase equilibria modelling; a Monte Carlo assessment and implications 82. Wu, C. M. Original calibration of a garnet geobarometer in metapelite. Minerals
for tectonic interpretations. Geosci. Front. 7, 591–607 (2016). https://doi.org/10.3390/Min9090540 (2019).
52. Walsh, A. K. et al. P-T-t evolution of a large, long-lived, ultrahigh-temperature Grenvillian 83. Jiao, S. J., Guo, J. H., Harley, S. L. & Peng, P. Geochronology and trace element
belt in central Australia. Gondwana Res. 28, 531–564 (2015). geochemistry of zircon, monazite and garnet from the garnetite and/or associated other
53. Wan, B. et al. Seismological evidence for the earliest global subduction network at 2 Ga high-grade rocks: Implications for Palaeoproterozoic tectonothermal evolution of the
ago. Sci. Adv. 6, eabc5491 (2020). Khondalite Belt, North China craton. Precambrian Res. 237, 78–100 (2013).
54. Mitchell, R. N. et al. The supercontinent cycle. Nat. Rev. Earth Environ. 2, 358–374 (2021). 84. Kohn, M. J., Corrie, S. L. & Markley, C. The fall and rise of metamorphic zircon. Am. Mineral.
55. Bleeker, W. The late Archean record: a puzzle in ca. 35 pieces. Lithos 71, 99–134 (2003). 100, 897–908 (2015).
56. Liu, Y. et al. Archean geodynamics: ephemeral supercontinents or long-lived 85. Jiao, S. et al. Establishing the PT path of UHT granulites by geochemically distinguishing
supercratons. Geology 49, 794–798 (2021). peritectic from retrograde garnet. Am. Mineral. J. Earth Planet. Mater. 106, 1640–1653 (2021).
57. Dhuime, B., Hawkesworth, C. J., Cawood, P. A. & Storey, C. D. A change in the 86. Otamendi, J. E., de la Rosa, J. D., Patiño Douce, A. E. & Castro, A. Rayleigh fractionation of
geodynamics of continental growth 3 billion years ago. Science 335, 1334–1336 (2012). heavy rare earths and yttrium during metamorphic garnet growth. Geology 30, 159–162
58. Tang, M., Chen, K. & Rudnick, R. L. Archean upper crust transition from mafic to felsic (2002).
marks the onset of plate tectonics. Science 351, 372–375 (2016). 87. Jiao, S. J. et al. Establishing the P–T path of UHT granulites by geochemically distinguishing
59. Cawood, P. A. et al. Geological archive of the onset of plate tectonics. Phil. Trans. R. Soc. peritectic from retrograde garnet. Am. Mineral. 106, 1640–1653 (2021).
A 376, 20170405 (2018). 88. Kihle, J., Harlov, D. E., Frigaard, O. & Jamtveit, B. Epitaxial quartz inclusions in corundum
60. Holder, R. M., Viete, D. R., Brown, M. & Johnson, T. E. Metamorphism and the evolution from a sapphirine-garnet boudin, Bamble Sector, SE Norway: SiO2-Al2O3 miscibility at high
of plate tectonics. Nature 572, 378–381 (2019). P-T dry granulite facies conditions. J. Metamorph. Geol. 28, 769–784 (2010).
This paper established the temporal correlation of the appearance of bimodal 89. Liu, T. & Wei, C. Metamorphic P–T paths and zircon U–Pb ages of Archean ultra-high
metamorphism and the global emergence of plate tectonics since the Neoarchaean. temperature paragneisses from the Qian’an gneiss dome, East Hebei terrane, North China
61. Dien, G. E., Doucet, L. S., Li, Z.-X., Cox, G. & Mitchell, R. Global geochemical craton. J. Metamorph. Geol. 38, 329–356 (2020).
fingerprinting of plume intensity suggests coupling with the supercontinent cycle. 90. Horton, F., Holder, R. M. & Swindle, C. R. An extensive record of orogenesis recorded
Nat. Commun. 10, 5270 (2019). in a Madagascar granulite. J. Metamorph. Geol. 40, 287–305 (2022).
62. Hawkesworth, C. J., Cawood, P. A. & Dhuime, B. The evolution of the continental 91. Mitchell, R. J. et al. Neoproterozoic evolution and Cambrian reworking of ultrahigh
crust and the onset of plate tectonics. Front. Earth Sci. https://doi.org/10.3389/ temperature granulites in the Eastern Ghats Province, India. J. Metamorph. Geol. 37,
feart.2020.00326 (2020). 977–1006 (2019).

Nature Reviews Earth & Environment


Review article

92. Halpin, J. A., Clarke, G. L., White, R. W. & Kelsey, D. E. Contrasting P–T–t paths for 126. Copley, A. & Weller, O. The controls on the thermal evolution of continental mountain
Neoproterozoic metamorphism in MacRobertson and Kemp Lands, east Antarctica. ranges. J. Metamorph. Geol. https://doi.org/10.1111/jmg.12664 (2022).
J. Metamorph. Geol. 25, 683–701 (2007). 127. McKenzie, D. & Priestley, K. The influence of lithospheric thickness variations on continental
93. Harley, S. L., Kelly, N. M. & Moller, A. Zircon behaviour and the thermal histories evolution. Lithos 102, 1–11 (2008).
of mountain chains. Elements 3, 25–30 (2007). 128. Clark, C., Fitzsimons, I. C. W., Healy, D. & Harley, S. L. How does the continental crust
94. Yakymchuk, C. & Brown, M. Behaviour of zircon and monazite during crustal melting. get really hot? Elements 7, 235–240 (2011).
J. Geol. Soc. 171, 465–479 (2014). This work highlighted that a long-lived mountain plateau with high radiogenic heat
95. Kohn, M. J. & Kelly, N. M. in Petrology and Geochronology of Metamorphic Zircon, 35–62 production and low erosion rate can lead to protracted UHT metamorphism.
(Wiley, 2018). 129. Clark, C. et al. Hot orogens and supercontinent amalgamation: a Gondwanan example
96. Kelsey, D. E., Clark, C. & Hand, M. Thermobarometric modelling of zircon and monazite from southern India. Gondwana Res. 28, 1310–1328 (2015).
growth in melt-bearing systems: examples using model metapelitic and metapsammitic 130. Jamieson, R. A., Beaumont, C., Warren, C. J. & Nguyen, M. H. The Grenville orogen
granulites. J. Metamorph. Geol. 26, 199–212 (2008). explained? Applications and limitations of integrating numerical models with geological
97. Spear, F. S. & Pyle, J. M. Theoretical modeling of monazite growth in a low-Ca metapelite. and geophysical data. Can. J. Earth Sci. 47, 517–539 (2010).
Chem. Geol. 273, 111–119 (2010). 131. Stüwe, K. in Geodynamics of the Lithosphere 337–367 (Springer, 2007).
98. Vorhies, S. H., Ague, J. J. & Schmitt, A. K. Zircon growth and recrystallization during 132. Andreoli, M. A., Hart, R. J., Ashwal, L. D. & Coetzee, H. Correlations between U, Th content
progressive metamorphism, Barrovian zones, Scotland. Am. Mineral. 98, 219–230 (2013). and metamorphic grade in the western Namaqualand Belt, South Africa, with implications
99. Korhonen, F. J., Clark, C., Brown, M., Bhattacharya, S. & Taylor, R. How long-lived is ultrahigh for radioactive heating of the crust. J. Petrol. 47, 1095–1118 (2006).
temperature (UHT) metamorphism? Constraints from zircon and monazite geochronology 133. Alessio, K. L. et al. Thermal modelling of very long-lived (>140 Myr) high thermal gradient
in the Eastern Ghats orogenic belt, India. Precambrian Res. 234, 322–350 (2013). metamorphism as a result of radiogenic heating in the Reynolds Range, central Australia.
100. Taylor, R. J. M. et al. Interpreting granulite facies events through rare earth element Lithos 352, 105280 (2020).
partitioning arrays. J. Metamorph. Geol. 35, 759–775 (2017). 134. Nandakumar, V. & Harley, S. L. Geochemical signatures of mid-crustal melting processes
101. Taylor, R. J. M. et al. Experimental determination of REE partition coefficients between zircon, and heat production in a hot orogen: the Kerala Khondalite Belt, Southern India. Lithos
garnet and melt: a key to understanding high-T crustal processes. J. Metamorph. Geol. 33, 324, 479–500 (2019).
231–248 (2015). 135. Huang, G., Guo, J., Jiao, S. & Palin, R. What drives the continental crust to be extremely
102. Rubatto, D. & Hermann, J. Experimental zircon/melt and zircon/garnet trace element hot so quickly? J. Geophys. Res. Solid Earth 124, 11218–11231 (2019).
partitioning and implications for the geochronology of crustal rocks. Chem. Geol. 241, 136. Horton, F., Hacker, B., Kylander-Clark, A., Holder, R. & Jöns, N. Focused radiogenic heating
38–61 (2007). of middle crust caused ultrahigh temperatures in southern Madagascar. Tectonics 35,
103. Clark, C., Taylor, R. J. M., Kylander-Clark, A. R. C. & Hacker, B. R. Prolonged (>100 Ma) 293–314 (2016).
ultrahigh temperature metamorphism in the Napier Complex, East Antarctica: 137. England, P. C. & Thompson, A. B. Pressure–temperature–time paths of regional
a petrochronological investigation of Earth’s hottest crust. J. Metamorph. Geol. 36, metamorphism. 1. Heat-transfer during the evolution of regions of thickened
1117–1139 (2018). continental-crust. J. Petrol. 25, 894–928 (1984).
104. Kusiak, M. A. et al. Changes in zircon chemistry during Archean UHT metamorphism 138. Lux, D. R., Deyoreo, J. J., Guldotti, C. V. & Decker, E. R. Role of plutonism in low-pressure
in the Napier Complex, Antarctica. Am. J. Sci. 313, 933–966 (2013). metamorphic belt formation. Nature 323, 794–797 (1986).
105. Kusiak, M. A., Whitehouse, M. J., Wilde, S. A., Nemchin, A. A. & Clark, C. Mobilization of 139. Hanson, R. B. & Barton, M. D. Thermal development of low-pressure metamorphic belts —
radiogenic Pb in zircon revealed by ion imaging: implications for early Earth geochronology. results from two-dimensional numerical models. J. Geophys. Res. Solid Earth Planets 94,
Geology 41, 291–294 (2013). 10363–10377 (1989).
106. Kunz, B. E., Regis, D. & Engi, M. Zircon ages in granulite facies rocks: decoupling from 140. Mitchell, R. K., Indares, A. & Ryan, B. High to ultrahigh temperature contact metamorphism
geochemistry above 850 °C? Contrib. Mineral. Petrol. 173, 26 (2018). and dry partial melting of the Tasiuyak paragneiss, Northern Labrador. J. Metamorph. Geol.
107. Kusiak, M. A. et al. Metallic lead nanospheres discovered in ancient zircons. Proc. Natl 32, 535–555 (2014).
Acad. Sci. USA 112, 4958–4963 (2015). 141. Laurent, A. T., Duchene, S., Bingen, B., Bosse, V. & Seydoux-Guillaume, A. M. Two
108. Brown, M., Johnson, T. E. & Spencer, C. J. Secular changes in metamorphism and successive phases of ultrahigh temperature metamorphism in Rogaland, S. Norway:
metamorphic cooling rates track the evolving plate tectonic regime on Earth. J. Geol. Soc. evidence from Y-in-monazite thermometry. J. Metamorph. Geol. 36, 1009–1037 (2018).
https://doi.org/10.1144/jgs2022-050 (2022). 142. Hyndman, R. D. Lower-crustal flow and detachment in the North American Cordillera:
109. Tang, M., Chu, X., Hao, J. & Shen, B. Orogenic quiescence in Earth’s middle age. Science a consequence of Cordillera-wide high temperatures. Geophys. J. Int. 209, 1779–1799
371, 728–731 (2021). (2017).
110. Spencer, C. J., Mitchell, R. N. & Brown, M. Enigmatic Mid-proterozoic orogens: hot, thin, 143. Thompson, A. B., Schulmann, K., Jezek, J. & Tolar, V. Thermally softened continental
and low. Geophys. Res. Lett. 48, e2021GL093312 (2021). extensional zones (arcs and rifts) as precursors to thickened orogenic belts. Tectonophysics
111. Ashwal, L. D. & Bybee, G. M. Crustal evolution and the temporality of anorthosites. 332, 115–141 (2001).
Earth Sci. Rev. 173, 307–330 (2017). 144. Nakakuki, T. & Mura, E. Dynamics of slab rollback and induced back-arc basin formation.
112. Bradley, D. C. Passive margins through Earth history. Earth Sci. Rev. 91, 1–26 (2008). Earth Planet. Sci. Lett. 361, 287–297 (2013).
113. O’Neill, C., Brown, M., Schaefer, B. & Gazi, J. A. Earth’s anomalous middle-age 145. Chen, Z. H., Schellart, W. P., Strak, V. & Duarte, J. C. Does subduction-induced mantle flow
magmatism driven by plate slowdown. Sci. Rep. 12, 10460 (2022). drive backarc extension? Earth Planet. Sci. Lett. 441, 200–210 (2016).
114. Stern, R. J. The Mesoproterozoic single-lid tectonic episode: prelude to modern plate 146. Pownall, J. M. et al. Miocene UHT granulites from Seram, eastern Indonesia:
tectonics. GSA Today 30, 4–10 (2020). a geochronological–REE study of zircon, monazite and garnet. Geol. Soc. Lond.
115. Roberts, N. M. et al. On the enigmatic mid-proterozoic: single-lid versus plate tectonics. Spec. Publ. https://doi.org/10.1144/sp478.8 (2018).
Earth Planet. Sci. Lett. 594, 117749 (2022). 147. Jacob, J. B., Scott, J. M., Turnbull, R. E., Tarling, M. S. & Sagar, M. W. High- to ultrahigh-
116. Spear, F. S. in Metamorphic Phase Equilibria and Pressure-Temperature-Time Paths, 25–71 temperature metamorphism in the lower crust: an example resulting from Hikurangi
(Mineralogical Society of America, 1993). Plateau collision and slab rollback in New Zealand. J. Metamorph. Geol. 35, 831–853
117. Jaupart, C., Mareschal, J.-C. & Iarotsky, L. Radiogenic heat production in the continental (2017).
crust. Lithos 262, 398–427 (2016). 148. Sizova, E., Hauzenberger, C., Fritz, H., Faryad, S. W. & Gerya, T. Late orogenic heating
118. Rudnick, R. L. & Gao, S. in Treatise on Geochemistry (eds Holland, H. D. & Turekian, K. K.) of (ultra)high pressure rocks: slab rollback vs. slab breakoff. Geosciences 9, 499 (2019).
1–64 (Pergamon, 2003). 149. van Hunen, J. & Miller, M. S. Collisional processes and links to episodic changes in
119. Bea, F. The sources of energy for crustal melting and the geochemistry of heat-producing subduction zones. Elements 11, 119–124 (2015).
elements. Lithos 153, 278–291 (2012). 150. Davies, J. H. & von Blanckenburg, F. Slab breakoff: a model of lithosphere detachment and
120. Alessio, K. L. et al. Conservation of deep crustal heat production. Geology 46, 335–338 its test in the magmatism and deformation of collisional orogens. Earth Planet. Sci. Lett. 129,
(2018). 85–102 (1995).
121. Williams, M. A., Kelsey, D. E. & Rubatto, D. Thorium zoning in monazite: a case study from 151. van Hunen, J. & Allen, M. B. Continental collision and slab break-off: a comparison
the Ivrea–Verbano zone, NW Italy. J. Metamorph. Geol. https://doi.org/10.1111/jmg.12656 of 3-D numerical models with observations. Earth Planet. Sci. Lett. 302, 27–37 (2011).
(2022). 152. van de Zedde, D. M. A. & Wortel, M. J. R. Shallow slab detachment as a transient source
122. Gard, M., Hasterok, D., Hand, M. & Cox, G. Variations in continental heat production from of heat at midlithospheric depths. Tectonics 20, 868–882 (2001).
4 Ga to the present: evidence from geochemical data. Lithos 342, 391–406 (2019). 153. Magni, V., Faccenna, C., van Hunen, J. & Funiciello, F. Delamination vs. break-off: the fate
123. Chamberlain, C. P. & Sonder, L. J. Heat-producing elements and the thermal and baric of continental collision. Geophys. Res. Lett. 40, 285–289 (2013).
patterns of metamorphic belts. Science 250, 763–769 (1990). 154. Freeburn, R., Bouilhol, P., Maunder, B., Magni, V. & van Hunen, J. Numerical models of
124. Huerta, A. D., Royden, L. H. & Hodges, K. V. The thermal structure of collisional orogens the magmatic processes induced by slab breakoff. Earth Planet. Sci. Lett. 478, 203–213
as a response to accretion, erosion, and radiogenic heating. J. Geophys. Res. Solid Earth (2017).
103, 15287–15302 (1998). 155. Herzberg, C., Condie, K. & Korenaga, J. Thermal history of the Earth and its petrological
125. Goffé, B., Bousquet, R., Henry, P. & Le Pichon, X. Effect of the chemical composition of the expression. Earth Planet. Sci. Lett. 292, 79–88 (2010).
crust on the metamorphic evolution of orogenic wedges. J. Metamorph. Geol. 21, 123–141 156. van Hunen, J. & van den Berg, A. P. Plate tectonics on the early Earth: limitations imposed
(2003). by strength and buoyancy of subducted lithosphere. Lithos 103, 217–235 (2008).

Nature Reviews Earth & Environment


Review article

157. Gou, L. L. et al. Ultrahigh-temperature metamorphism in the Helanshan complex of the 189. Chen, L. et al. Seismically constrained thermo-rheological structure of the eastern
Khondalite Belt, North China craton: petrology and phase equilibria of spinel-bearing Tibetan margin: implication for lithospheric delamination. Tectonophysics 627, 122–134
pelitic granulites. J. Metamorph. Geol. 36, 1199–1220 (2018). (2014).
158. Beunk, F. F., de Roever, E. W. F., Yi, K. & Brouwer, F. M. Structural and tectonothermal 190. Chen, M. et al. Lithospheric foundering and underthrusting imaged beneath Tibet.
evolution of the ultrahigh-temperature Bakhuis Granulite Belt, Guiana Shield, Surinam: Nat. Commun. 8, 15659 (2017).
Palaeoproterozoic to recent. Geosci. Front. 12, 677–692 (2021). 191. Bao, X. & Shen, Y. Early-stage lithospheric foundering beneath the eastern Tibetan Plateau
159. Houseman, G. A., Mckenzie, D. P. & Molnar, P. Convective instability of a thickened revealed by full-wave Pn tomography. Geo. Res. Lett. https://doi.org/10.1029/2019gl086469
boundary-layer and its relevance for the thermal evolution of continental convergent (2020).
belts. J. Geophys. Res. Solid Earth 86, 6115–6132 (1981). 192. Wu, Y. K., Bao, X. W., Zhang, B. F., Xu, Y. X. & Yang, W. C. Seismic evidence for stepwise
160. Houseman, G. A. & Molnar, P. Gravitational (Rayleigh–Taylor) instability of a layer with lithospheric delamination beneath the Tibetan Plateau. Geophys. Res. Lett. 49,
non-linear viscosity and convective thinning of continental lithosphere. Geophys. J. Int. e2022GL098528 (2022).
128, 125–150 (1997). 193. Jiménez-Munt, I. & Platt, J. P. Influence of mantle dynamics on the topographic evolution
161. Göğüş, O. H. & Pysklywec, R. N. Near-surface diagnostics of dripping or delaminating of the Tibetan Plateau: Results from numerical modeling. Tectonics https://doi.org/
lithosphere. J. Geophys. Res. Solid Earth https://doi.org/10.1029/2007jb005123 (2008). 10.1029/2006tc001963 (2006).
162. Bird, P. Continental delamination and the Colorado Plateau. J. Geophys. Res. Solid Earth 194. Huangfu, P. et al. Multi-terrane structure controls the contrasting lithospheric evolution
84, 7561–7571 (1979). beneath the western and central-eastern Tibetan Plateau. Nat. Commun. 9, 3780 (2018).
163. Göğüş, O. H. & Ueda, K. Peeling back the lithosphere: controlling parameters, surface 195. Hearn, T. M., Ni, J. F., Wang, H., Sandvol, E. A. & Chen, Y. J. Depth-dependent Pn velocities
expressions and the future directions in delamination modeling. J. Geodyn. 117, 21–40 and configuration of Indian and Asian lithosphere beneath the Tibetan Plateau. Geophys.
(2018). J. Int. 217, 179–189 (2019).
164. Chardon, D., Gapais, D. & Cagnard, F. Flow of ultra-hot orogens: a view from the 196. Li, C., Van der Hilst, R. D., Meltzer, A. S. & Engdahl, E. R. Subduction of the Indian lithosphere
Precambrian, clues for the Phanerozoic. Tectonophysics 477, 105–118 (2009). beneath the Tibetan Plateau and Burma. Earth Planet. Sc. Lett. 274, 157–168 (2008).
165. Perchuk, A. L. et al. Precambrian ultra-hot orogenic factory: making and reworking 197. Lei, J. S., Zhao, D. P. & Su, Y. J. Insight into the origin of the Tengchong intraplate volcano
of continental crust. Tectonophysics 746, 572–586 (2018). and seismotectonics in southwest China from local and teleseismic data. J. Geophys.
166. Smithies, R. H. et al. The Mesoproterozoic thermal evolution of the Musgrave Province Res. Solid Earth https://doi.org/10.1029/2008jb005881 (2009).
in central Australia — plume vs. the geological record. Gondwana Res. 27, 1419–1429 198. Huang, Z. C. et al. P and S wave tomography beneath the SE Tibetan Plateau: evidence
(2015). for lithospheric delamination. J. Geophys. Res. Solid Earth 124, 10292–10308 (2019).
167. Smithies, R. H. et al. High-temperature granite magmatism, crust–mantle interaction 199. Feng, J., Yao, H., Chen, L. & Wang, W. Massive lithospheric delamination in southeastern
and the Mesoproterozoic intracontinental evolution of the Musgrave Province, Central Tibet facilitating continental extrusion. Natl Sci. Rev. 9, nwab174 (2022).
Australia. J. Pet. 52, 931–958 (2011). 200. Wang, J. M. et al. First evidence of eclogites overprinted by ultrahigh temperature
168. Uyeda, S. & MiyashiroI, A. Plate tectonics and the Japanese islands: a synthesis. GSA Bull. metamorphism in Everest East, Himalaya: implications for collisional tectonics on early
85, 1159–1170 (1974). Earth. Earth Planet. Sci Lett. https://doi.org/10.1016/j.epsl.2021.116760 (2021).
169. Dickinson, W. R. & Snyder, W. S. Geometry of subducted slabs related to San-Andreas 201. Wu, C. G. et al. Tectonothermal transition from continental collision to post-collision:
transform. J. Geol. 87, 609–627 (1979). insights from eclogites overprinted in the ultrahigh-temperature granulite facies
170. Thorkelson, D. J. & Taylor, R. P. Cordilleran slab windows. Geology 17, 833–836 (1989). (Yadong region, central Himalaya). J. Metamorph. Geol. 40, 955–981 (2022).
171. Thorkelson, D. J. Subduction of diverging plates and the principles of slab window 202. Dong, X., Zhang, Z., Tian, Z., Niu, Y. & Zhang, L. Protoliths and metamorphism of the
formation. Tectonophysics 255, 47–63 (1996). central Himalayan eclogites: zircon/titanite U–Pb geochronology, Hf isotope and
172. DeLong, S. E., Schwarz, W. M. & Anderson, R. N. Thermal effects of ridge subduction. geochemistry. Gondwana Res. 104, 39–53 (2022).
Earth Planet. Sci. Lett. 44, 239–246 (1979). 203. Chen, L., Song, X., Gerya, T. V., Xu, T. & Chen, Y. Crustal melting beneath orogenic plateaus:
173. Santosh, M. & Kusky, T. Origin of paired high pressure-ultrahigh-temperature orogens: insights from 3-D thermo-mechanical modeling. Tectonophysics 761, 1–15 (2019).
a ridge subduction and slab window model. Terra Nova 22, 35–42 (2010). 204. Zhang, X.-Z. et al. Tibetan Plateau insights into >1100°C crustal melting in the Quaternary.
174. Peng, P., Guo, J. H., Windley, B. F. & Li, X. H. Halaqin volcano-sedimentary succession in Geology 50, 1432–1437 (2022).
the central-northern margin of the North China craton: products of Late Paleoproterozoic 205. Chen, S. et al. Cenozoic ultrahigh-temperature metamorphism in pelitic granulites from
ridge subduction. Precambrian Res. 187, 165–180 (2011). the Mogok metamorphic belt, Myanmar. Sci. China Earth Sci. 64, 1873–1892 (2021).
175. Uehara, S. I. & Aoya, M. Thermal model for approach of a spreading ridge to subduction 206. Searle, M. P. et al. Tectonic evolution of the Mogok metamorphic belt, Burma (Myanmar)
zones and its implications for high-P/high-T metamorphism: importance of subduction constrained by U–Th–Pb dating of metamorphic and magmatic rocks. Tectonics
versus ridge approach ratio. Tectonics https://doi.org/10.1029/2004TC001715 (2005). https://doi.org/10.1029/2006tc002083 (2007).
176. Gasser, D., Rubatto, D., Bruand, E. & Stuwe, K. Large-scale, short-lived metamorphism, 207. Wang, Y., Zhang, X., Jiang, C., Wei, H. & Wan, J. Tectonic controls on the late Miocene-
deformation, and magmatism in the Chugach metamorphic complex, southern Alaska: Holocene volcanic eruptions of the Tengchong volcanic field along the southeastern
a SHRIMP U–Pb study of zircons. Geol. Soc. Am. Bull. 124, 886–905 (2012). margin of the Tibetan Plateau. J. Asian Earth Sci. 30, 375–389 (2007).
177. Hill, R. I., Campbell, I. H., Davies, G. F. & Griffiths, R. W. Mantle plumes and continental 208. Karlstrom, K. E. et al. Tectonics of the Colorado Plateau and its margins. Annu. Rev. Earth
tectonics. Science 256, 186–193 (1992). Planet. Sci. 50, 295–322 (2022).
178. Sleep, N. H. Mantle plumes from top to bottom. Earth Sci. Rev. 77, 231–271 (2006). 209. Schmandt, B., Dueker, K., Humphreys, E. & Hansen, S. Hot mantle upwelling across the
179. Herzberg, C. Understanding the Paleoproterozoic Circum-Superior Large Igneous Province 660 beneath Yellowstone. Earth Planet. Sc. Lett. 331, 224–236 (2012).
constrains the thermal properties of Earth’s mantle through time. Precambrian Res. 375, 210. Nelson, P. L. & Grand, S. P. Lower-mantle plume beneath the Yellowstone hotspot revealed
106671 (2022). by core waves. Nat. Geosci. 11, 280–284 (2018).
180. Dawson, J. B., Harley, S. L., Rudnick, R. L. & Irel, T. R. Equilibration and reaction in 211. James, D. E., Fouch, M. J., Carlson, R. W. & Roth, J. B. Slab fragmentation, edge flow and
Archaean quartz-sapphirine granulite xenoliths from the Lace kimberlite pipe, South the origin of the Yellowstone hotspot track. Earth Planet. Sci. Lett. 311, 124–135 (2011).
Africa. J. Metamorph. Geol. 15, 253–266 (1997). 212. Leonard, T. & Liu, L. J. The role of a mantle plume in the formation of Yellowstone volcanism.
181. Schmitz, M. D. & Bowring, S. A. Ultrahigh-temperature metamorphism in the lower crust Geophys. Res. Lett. 43, 1132–1139 (2016).
during Neoarchean Ventersdorp rifting and magmatism, Kaapvaal craton, southern 213. Zhou, Q., Liu, L. & Hu, J. Western US volcanism due to intruding oceanic mantle driven
Africa. Geol. Soc. Am. Bull. 115, 533–548 (2003). by ancient Farallon slabs. Nat. Geosci. 11, 70–76 (2017).
182. Harrison, T. M., Copeland, P., Kidd, W. S. F. & Yin, A. Raising Tibet. Science 255, 1663–1670 214. Zhou, Y. Anomalous mantle transition zone beneath the Yellowstone hotspot track.
(1992). Nat. Geosci. 11, 449–453 (2018).
183. Ding, L. et al. Timing and mechanisms of Tibetan Plateau uplift. Nat. Rev. Earth Environ. 215. Hawley, W. B. & Allen, R. M. The fragmented death of the Farallon Plate. Geophys. Res. Lett.
https://doi.org/10.1038/s43017-022-00318-4 (2022). 46, 7386–7394 (2019).
184. Turner, S. et al. Timing of Tibetan uplift constrained by analysis of volcanic-rocks. Nature 216. Hayob, J. L., Essene, E. J., Ruiz, J., Ortega-Gutiérrez, F. & Aranda-Gómez, J. J. Young
364, 50–54 (1993). high-temperature granulites from the base of the crust in central Mexico. Nature 342,
185. Liu, C. Z., Wu, F. Y., Chung, S. L. & Zhao, Z. D. Fragments of hot and metasomatized 265 (1989).
mantle lithosphere in Middle Miocene ultrapotassic lavas, southern Tibet. Geology 39, 217. Solari, L. A., Aranda-Gómez, J. J., Moreno-Arredondo, A. & Maldonado, R. U–Pb age of
923–926 (2011). a late Cenozoic ultra-high temperature metamorphic event under Central Mexico, as
186. Ji, W. Q., Wu, F. Y., Chung, S. L. & Liu, C. Z. The Gangdese magmatic constraints on a inferred from granulite xenoliths from Cerro El Toro, Mexico. Int. Geol. Rev. 65, 1–22
latest Cretaceous lithospheric delamination of the Lhasa terrane, southern Tibet. Lithos (2022).
210, 168–180 (2014). 218. Ghent, E. D., Edwards, B. R., Russell, J. K. & Mortensen, J. Granulite facies xenoliths from
187. Klemperer, S. L. et al. Limited underthrusting of India below Tibet: 3He/4He analysis of Prindle volcano, Alaska: implications for the northern Cordilleran crustal lithosphere.
thermal springs locates the mantle suture in continental collision. Proc. Natl Acad. Sci. USA Lithos 101, 344–358 (2008).
119, e2113877119 (2022). 219. Campbell, I. H., Griffiths, R. W. & Hill, R. I. Melting in an Archaean mantle plume: heads it’s
188. Ren, Y. & Shen, Y. Finite frequency tomography in southeastern Tibet: evidence for the basalts, tails it’s komatiites. Nature 339, 697–699 (1989).
causal relationship between mantle lithosphere delamination and the north–south 220. Brown, M. The contribution of metamorphic petrology to understanding lithosphere
trending rifts. J. Geophys. Res. Solid Earth https://doi.org/10.1029/2008jb005615 (2008). evolution and geodynamics. Geosci. Front. 5, 553–569 (2014).

Nature Reviews Earth & Environment


Review article

221. Patiño Douce, A. E. & Harris, N. Experimental constraints on Himalayan anatexis. J. Petrol. were funded by grant 41890832 from the National Natural Science Foundation of China (NSFC).
39, 689–710 (1998). S.J. was funded by grant 42122018 from NSFC.
222. Weller, O. M. et al. The metamorphic and magmatic record of collisional orogens.
Nat. Rev. Earth Environ. 2, 781–799 (2021). Author contributions
223. Wang, C., Mitchell, R. N., Murphy, J. B., Peng, P. & Spencer, C. J. The role of megacontinents S.J. and Y.C. conceived the idea. S.J. drafted the manuscript. S.J. drafted the figures with
in the supercontinent cycle. Geology 49, 402–406 (2021). contributions from C.C. and L.C. M.B. contributed substantially to the manuscript preparation,
224. Korhonen, F. J., Clark, C., Brown, M. & Taylor, R. J. M. Taking the temperature of Earth’s interpretation, discussion, writing and editing of the manuscript. R.N.M., P.C. and C.C.
hottest crust. Earth Planet. Sci. Lett. 408, 341–354 (2014). contributed to the discussion of contents, interpretation and editing. All authors reviewed
225. Kondou, N., Tsunogae, T., Santosh, M. & Shimizu, H. Sapphirine plus quartz assemblage the manuscript before submission.
from Ganguvarpatti: diagnostic evidence for ultrahigh-temperature metamorphism in
central Madurai Block, southern India. J. Mineral. Petrol. Sci. 104, 285–289 (2009).
Competing interests
226. Jiao, S. J. & Guo, J. H. Paleoproterozoic UHT metamorphism with isobaric cooling (IBC)
The authors declare no competing interests.
followed by decompression-heating in the Khondalite Belt (North China craton): new
evidence from two sapphirine formation processes. J. Metamorph. Geol. 38, 357–378
(2020). Additional information
227. Whitney, D. L. & Evans, B. W. Abbreviations for names of rock-forming minerals. Am. Mineral. Supplementary information The online version contains supplementary material available at
95, 185–187 (2010). https://doi.org/10.1038/s43017-023-00403-2.
228. Kohn, M. J., Engi, M. & Lanari, P. (eds) Petrochronology: Methods and Applications.
Reviews in Mineralogy and Geochemistry Vol. 83 (Mineralogical Society of America, Peer review information Nature Reviews Earth & Environment thanks S. Harley, F. Horton and
2017). the other, anonymous, reviewer(s) for their contribution to the peer review of this work.
229. Chakraborty, S. Diffusion in solid silicates: a tool to track timescales of processes comes
of age. Annu. Rev. Earth Planet. Sci. 36, 153–190 (2008). Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
230. Chowdhury, P. & Chakraborty, S. Slow cooling at higher temperatures recorded within published maps and institutional affiliations.
high-P mafic granulites from the Southern Granulite Terrain, India: implications for the
presence and style of plate tectonics near the Archean–Proterozoic boundary. J. Petrol. Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this
60, 441–486 (2019). article under a publishing agreement with the author(s) or other rightsholder(s); author self-
archiving of the accepted manuscript version of this article is solely governed by the terms
of such publishing agreement and applicable law.
Acknowledgements
The authors thank X. H. Li, Ling Chen and X. F. Liang for their comments on an initial draft,
Y. B. Liu for help with Fig. 2b–e and X. F. Liang for help with Fig. 6c and f. S.J., G.Y.H. and J.H.G. © Springer Nature Limited 2023

Nature Reviews Earth & Environment

View publication stats

You might also like