You are on page 1of 19

Journal of Geophysical Research: Solid Earth

RESEARCH ARTICLE Making Cratonic Lithospheric Mantle


10.1029/2018JB016179
Bin Su1 and Yi Chen1,2,3
Key Points: 1
State Key Laboratory of Lithospheric Evolution, Institute of Geology and Geophysics, Chinese Academy of Sciences,
• The quantitative relationship
between garnet and cratonic Beijing, China, 2Chinese Academy of Sciences Center for Excellence in Tibetan Plateau Earth Sciences, Beijing, China,
3
peridotite has been determined College of Earth and Planetary Sciences, University of Chinese Academy of Sciences, Beijing, China
using thermodynamic modeling
approach
• High Cr# values of garnet inclusions Abstract The origin of cratonic lithospheric mantle has been attributed to either high-pressure (5–7 GPa)
in diamond imply a low-pressure
melting in hot mantle plumes or low-pressure (<5 GPa) melting in mid-ocean ridges or suprasubduction
(<4 GPa) origin for cratonic
lithospheric mantle zones. To resolve this long-standing debate, it is necessary to confirm under what depths the incipient
• Cratonic lithospheric mantle likely cratonic mantle melted. Compared with most cratonic mantle xenoliths and xenocrysts, which commonly
melted in the Archean oceanic ridges
experienced metasomatic modification after melt extraction, diamond inclusions with predominantly
and subsequently grew through the
stacking of oceanic lithosphere harzburgitic compositions can better track the compositional signature of pristine cratonic mantle. This paper
presents thermodynamic calculations performed with THERMOCALC software to establish a quantitative
Supporting Information:
relationship between garnet and cratonic peridotite compositions. Along the normal cratonic geotherm
• Supporting Information S1 (40 mW/m2), the XCa [atomic Ca/(Ca + Mg + Fe2+)] and Cr# [atomic Cr/(Cr + Al) × 100] values in garnet depend
mainly on the bulk CaO/Al2O3 and Cr# values, respectively. Furthermore, mantle melting modeling shows
Correspondence to: that the Cr# of the residue displays a negative correlation with melting pressure at melt fractions of greater
Y. Chen,
chenyi@mail.iggcas.ac.cn
than ~15%. Therefore, the high Cr# values (mostly 12–36) of global garnet inclusions in diamond support a
low-pressure (<4 GPa) origin for cratonic lithospheric mantle. More importantly, the incipient cratonic mantle
calculated from the garnet inclusions exhibits similar bulk CaO/Al2O3 and Cr# compositions to oceanic
Citation:
Su, B., & Chen, Y. (2018). Making cratonic
lithosphere but different values from arc lithosphere with higher CaO/Al2O3 and Cr# values. These results
lithospheric mantle. Journal of imply that cratonic mantle was initially formed by the extensive melting of hot ambient mantle in the
Geophysical Research: Solid Earth, 123, Archean ocean ridge environments. The shallow oceanic lithosphere was subsequently stacked to generate
7688–7706. https://doi.org/10.1029/
2018JB016179
the thick and stable cratonic lithospheric mantle.

Received 4 JUN 2018


Accepted 2 SEP 2018
1. Introduction
Accepted article online 5 SEP 2018 Cratons are underlain by thick, cold and chemically depleted lithospheric mantle. The preservation of
Published online 28 SEP 2018
low-density and high-viscosity lithospheric mantle has long been considered to be crucial for the stability
of cratons (e.g., Boyd, 1989; Foley, 2008; Griffin et al., 2003; Herzberg & Rudnick, 2012; Lee et al., 2011).
Cratonic lithospheric mantle primarily formed during the Archean (2.7–3.2 Ga), coupled with the emergence
of ancient continental crust (Carlson et al., 2005; Lee et al., 2011). The secular evolution of cratonic
lithospheric mantle was essential for changing Earth’s climate and supplying nutrients for biological diversity
(Froelich et al., 1982; Kasting et al., 1993). Hence, the cratonic mantle is likely to record information about
continental growth, ancient plate tectonics, and the evolution of the atmosphere, hydrosphere,
and biosphere.
The highly melt-depleted characteristic of cratonic peridotites indicates that Archean lithospheric mantle
formed after experiencing a high degree of partial melting (Herzberg, 2004; Walter, 2003), but its original
tectonic environment is still controversial. One hypothesis has suggested that cratonic mantle resulted from
high-pressure (5–7 GPa) melting in large mantle plumes (e.g., Arndt et al., 2009; Aulbach, 2012; Boyd, 1989;
Griffin et al., 2003; Herzberg, 1993), thus indicating that it was formed in situ (Figure 1a). Another hypothesis
prefers low-pressure (<5 GPa) melting in either mid-ocean ridges (MORs; Figure 1b) or suprasubduction
zones (SSZs; Figure 1c), followed by the stacking/accretion of oceanic lithosphere or arc lithosphere,
respectively (e.g., Canil, 2004; Carlson et al., 2005; Helmstaedt & Schulze, 1989; Herzberg, 2004; Herzberg
et al., 2010; Herzberg & Rudnick, 2012; Lee, 2006; Wittig et al., 2008). To resolve this puzzle, it is first necessary
to confirm under what depths the incipient cratonic mantle melted.
To date, the lithospheric mantle has mostly been studied based on peridotite xenoliths and xenocrysts in
volcanic rocks (Griffin et al., 2009; Kelemen et al., 1998; Pearson & Wittig, 2014; Rudnick et al., 1993;
©2018. American Geophysical Union. Ziberna et al., 2013). The whole-rock compositions of these peridotites have been used to constrain the melt
All Rights Reserved. fraction and pressure of melt extraction (e.g., Herzberg, 2004; Wittig et al., 2008). However, most cratonic

SU AND CHEN 7688


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Figure 1. Schematic illustration of the three geodynamic settings for formation of cratonic lithospheric mantle: High-
pressure partial melting in large and hot mantle plumes (a), low-pressure melt extraction in mid-oceanic ridges (b),
and suprasubduction zones (c). LAB, lithosphere-asthenosphere boundary.

peridotites were subjected to postmelting metasomatism and host-magma contamination, which strongly
affected their initial compositions (Foley, 2008; Herzberg, 2004; Lee et al., 2011). Moreover, seismic
tomography images of southern Africa have indicated that most xenoliths in kimberlites derived from
lower-velocity cratonic margins are not representative of true cratonic roots (Griffin et al., 2009). Therefore,
previous approaches based on whole-rock chemistry of cratonic mantle xenoliths may have drawn equivocal
conclusions. For example, the low FeO contents observed in many cratonic peridotites (mostly
orthopyroxene-rich) are considered to represent evidence for high-pressure melting (e.g., Aulbach, 2012;
Griffin et al., 2003; Pearson et al., 1995), whereas the low Al2O3 and mildly incompatible trace element con-
tents in the same samples support low-pressure melting (Canil, 2004; Wittig et al., 2008). Thus, it is challen-
ging to find reliable geochemical indicators that preserve the residual characteristics of cratonic mantle.
Diamonds erupted by kimberlites were formed during the early evolution of the Earth’s mantle and provide
an impermeable barrier that can insulate mineral inclusions from metasomatism (Luth & Stachel, 2014;
Richardson et al., 1984; Shirey & Richardson, 2011; Stachel & Harris, 2008). Thus, diamond inclusions are
expected to provide faithful information about the geochemical signatures of pristine cratonic mantle.
These inclusions are conventionally subdivided into peridotitic and eclogitic suites (Sobolev et al., 1973;
Stachel & Harris, 2008); this distinction is well defined by garnet inclusions, with peridotitic ones having
Cr2O3 contents of greater than ~1.5 wt %. Compared with garnet xenocrysts, as well as those in xenoliths,
the garnet inclusions of peridotitic suite contain lower CaO and higher Cr2O3 contents, defining host rocks
mostly harzburgitic compositions (Figure 2a). This difference suggests that the unique nature of highly
depleted cratonic mantle can be better inferred from these mineral inclusions. Geothermobarometric calcu-
lations for nontouching peridotitic inclusions in diamond shows that these diamond associations predomi-
nantly formed along representative cratonic mantle geotherms corresponding to 38–42 mW/m2 surface
heat flow, which is far below the dry peridotite solidus (Boyd et al., 1985; Phillips et al., 2004; Stachel &
Harris, 2008; Stachel & Luth, 2015). Moreover, mantle melting conditions must be calculated on the basis
of whole-rock compositions (e.g., Herzberg, 2004; Lee & Chin, 2014). Therefore, it is necessary to determine
the link between diamond inclusions and whole-rock compositions.
An effective approach is phase equilibrium forward modeling. Using thermodynamic models, it is practical to
calculate mineral/melt compositions and proportions as a function of temperature, pressure, and bulk com-
position, as well as to simulate the composition of melting residue. In contrast, whole-rock compositions can
also be constrained on the basis of selected mineral compositions. Here we establish a set of phase equili-
brium relationships for mantle peridotites in both closed and open systems and present a quantitative rela-
tionship between garnet and whole-rock compositions, which will provide new insight into the origin of
cratonic lithospheric mantle.

SU AND CHEN 7689


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

(a) Garnet inclusions in diamond (b) Garnet xenocrysts in kimberlite


20.0 20.0
Harzburgitic

Harzburgitic
15.0 15.0

Cr2O3 (wt%)

Cr2O3 (wt%)
Lherzolitic

Lherzolitic
10.0 10.0
Wehrlitic
Wehrlitic

5.0 5.0

Eclogitic N=655 Eclogitic N=530

0.0 0.0
0.0 3.0 6.0 9.0 12.0 15.0 0.0 3.0 6.0 9.0 12.0 15.0
CaO (wt%) CaO (wt%)

Figure 2. Cr2O3 versus CaO (wt %) discrimination diagrams (Sobolev et al., 1973) for garnet inclusions in diamond (a) and
garnet xenocrysts in kimberlite (b). The data sources for garnet inclusions in diamond are given in Stachel and Harris (2008),
and the compositions of garnet xenocrysts are from the GEOROC database (http://georoc.mpch-mainz.gwdg.de/georoc/).

2. Methods
2.1. Thermodynamic Models
Phase equilibrium modelings were performed using the THERMOCALC version 3.40 software (Powell et al.,
1998) with the internally consistent thermodynamic data set of Holland and Powell (2011; update
tc-ds622). The modeling is done in a comprehensive system, NCFMASOCr (Na2O-CaO-FeO-MgO-Al2O3-
SiO2-Fe2O3-Cr2O3), using recent activity–composition models for peridotite-forming minerals (garnet, spinel,
clinopyroxene, orthopyroxene, and olivine) and melt (Jennings & Holland, 2015). The a–x models were
updated in the expanded Cr-bearing system, which provides better consistency with natural rocks and
experimental results. Considering the thick and dry signatures of cratonic lithospheric mantle (Carlson
et al., 2005; Lee et al., 2011), low-pressure and hydrous phases (e.g., feldspar, chlorite, and amphibole) have
been ignored. Sodium is excluded in clinopyroxene-free peridotites.

2.2. Oxygen Fugacity (fO2) Calculations


Oxygen fugacity plays a vital role in the stability of diamond; therefore, it should be considered during mod-
eling. In this study, a recent calibration of the olivine-orthopyroxene-garnet oxybarometer (Stagno et al.,
2013) is used to establish the oxygen fugacity values of cratonic mantle peridotites. Compared with previous
calibration performed at pressures of 3 and 3.5 GPa (Gudmundsson & Wood, 1995), the new oxybarometer
was constructed at pressures from 3 to 7 GPa (Stagno et al., 2013) and was able to reproduce experimental
data more precisely. A key parameter in this oxybarometer is the Fe3+/∑Fe ratio in garnet (i.e., the andradite
end-member), which can be obtained by phase equilibrium modeling in the Fe3+-bearing system. Therefore,
by combining calculations of mineral compositions and the oxybarometer expression established by Stagno
et al. (2013), the oxygen fugacity of the cratonic mantle under various P-T conditions can be calculated.

2.3. Melting Residue Calculations


Here we take the bulk composition of KLB-1 as representative of fertile mantle to investigate mantle melting,
as the melt model was originally calibrated using the experimental data of KLB-1 (Jennings & Holland, 2015).
Moreover, KLB-1 has been used as a starting material in many high-pressure experiments (Davis et al., 2009,
and references therein), which can thus provide comparisons for this modeling. To better constrain the pres-
sure of melt extraction, both isobaric (fractional vs. batch) and polybaric cases were calculated from 0 to
7 GPa. Batch melting modeling assumes a closed-system environment where the residue is in equilibrium
with the entire melt. In contrast, fractional melting is performed under open-system conditions in which
the residue only equilibrates with the final drop of melt. For isobaric fractional melting, melt extraction is con-
sidered to be a cyclic process, with melt loss events occurring periodically at a given melt proportion thresh-
old of ~0.5% (Herzberg & O’Hara, 2002). For polybaric fractional melting, melt is assumed to be removed from
the melting residue each time the pressure decreases by 1 kbar along the adiabatic melting path (Herzberg &

SU AND CHEN 7690


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

O’Hara, 2002). Fractional melting processes change the composition of each successive residue, which is
melted (calculated) again after each instance of melt extraction. Detailed methods of fractional melting simu-
lations are provided in the supporting information (Text S1).
2.4. Whole-Rock and Mineral Compositions for Phase Equilibrium Modeling
The thermodynamic modeling of cratonic lithospheric mantle uses an average composition of Archean cra-
tonic mantle xenoliths from South Africa, Slave, Siberia, and Tanzania (Table S1 in the supporting informa-
tion). All xenoliths are characterized by low Al2O3, CaO, and Na2O contents and high MgO content due to
their high degree of melt extraction (Figure S1 in the supporting information). The Fe2O3 content of the
depleted cratonic peridotite is assumed to be 0.1 wt %, as normal mantle peridotites record Fe2O3 contents
of 0.1–0.4 wt % (Canil et al., 1994). The effects of Fe3+ on phase equilibrium, mineral composition, and oxygen
fugacity conditions will be discussed later. The bulk composition of KLB-1 is taken from Davis et al. (2009),
with an assumption of 0.3 wt % Fe2O3 (Jennings & Holland, 2015). Moreover, the effects of compositional var-
iations in the early cratonic mantle on phase relations are evaluated by adding pyroxene or garnet to cratonic
peridotites. The whole-rock and mineral compositions used for modeling are given in the supporting infor-
mation (Tables S1 and S2).

3. P-T Phase Relations and Compositional Isopleths of Garnet


Phase equilibrium modeling was first performed in an isochemical system to find one or more useful mineral
end-members insensitive to P-T changes along the normal cratonic geotherm. Then, open-system processes
were simulated to determine the quantitative relationship between mineral and whole-rock compositions.
The P-T phase relations for the average composition of cratonic peridotites at 600–1500 °C and 10–70 kbar
are shown in Figure 3, in which the supersolidus region has been simplified. The stability field of spinel peri-
dotite is limited to low-pressure regions (< 30 kbar) and diminishes with decreasing temperature. Garnet
occurs at higher pressure and equilibrates with spinel toward larger pressure ranges at lower temperatures,
which is consistent with the experimental results obtained in the Cr-bearing system (Girnis & Brey, 1999;
Webb & Wood, 1986). Spinel almost reacts out in the diamond stability field, where the modeled peridotite
has a low Cr# [atomic Cr/(Cr + Al) × 100] of 19.9. However, spinel can stabilize to greater depths in samples
with higher Cr# values (Klemme, 2004). Based on the above stabilities of garnet and spinel, the modeled peri-
dotite can be divided into spinel peridotite, garnet-spinel peridotite, and garnet peridotite.
Compared to olivine, orthopyroxene, and clinopyroxene, garnet generally shows a broader compositional
range (Griffin et al., 1999; Grütter et al., 2004) and consequently can be used to constrain the composition
of its host rock. In Figures 3a and 3b, the isopleths of 100*XCa [atomic Ca/(Ca + Mg + Fe2+) × 100] and Cr#
values of garnet are shown. In the stability field of garnet-spinel peridotite, the XCa values decrease with
increasing pressure, whereas the Cr# values display the opposite trend. In the garnet peridotite field, the
XCa isopleths are nearly parallel to the normal cratonic geotherm (e.g., 40 mW/m2; Hasterok & Chapman,
2011), with smaller values under lower-T and higher-P conditions; variations in the Cr# values in garnet are
limited, except at extremely high-T and low-P conditions. Although both XCa and Cr# in garnet exhibit varia-
tions in the calculated P-T range, they do not change significantly along the normal cratonic geotherm in the
spinel-free field (Figures 3c and 3d). Previous studies demonstrated that the minerals included in diamond
have reached subsolidus reequilibration along the normal cratonic geotherm before or during diamond for-
mation (Phillips et al., 2004; Stachel & Luth, 2015). Therefore, the XCa and Cr# values of garnet inclusions in
diamond have no relation with their derived depths but instead mainly depend on the bulk compositions
of their host peridotites. In the following sections, we will evaluate the effects of bulk compositions on these
two parameters.

4. Effects of Bulk Compositions on Garnet Compositions


4.1. Fe3+ Content
To evaluate the effect of bulk Fe3+ contents on garnet XCa and Cr#, the P-Fe3+/∑Fe pseudosection was mod-
eled at 1100 °C (Figure 4). In the stability field of garnet peridotite, oxygen fugacity shows a positive correla-
tion with bulk Fe3+/∑Fe, but it shows a decelerated increase with increasing bulk Fe3+/∑Fe. To ensure the
stability of diamond, the oxygen fugacity of the host rock should plot between the IW (iron-wüstite) and

SU AND CHEN 7691


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

(a) Garnet: 100*XCa (b) Garnet: Cr#


70 70

60 60

Pressure (kbar)
50 50

Pressure (kbar)
40 40

30 30

20 20

10 10
600 800 1000 1200 1400 600 800 1000 1200 1400
Temperature (°C) Temperature (°C)

20 20
(c) (d)
30 30
Pressure (kbar)

Pressure (kbar)
40 40

50 50
Dunite
Harzburgite
Primitive mantle (KLB-1)
Wehrlite
60 60

70 70
5.0 10.0 15.0 20.0 0 10 20 30 40
Garnet: 100*XCa Garnet: Cr#

Figure 3. P-T pseudosection in the NCFMASOCr system (a, b) based on the average composition (in mol %) of cratonic peri-
dotites (SiO2 = 37.20, Al2O3 = 0.58, CaO = 0.76, MgO = 56.28, FeO = 5.03, Na2O = 0.05, Fe2O3 = 0.03, and Cr2O3 = 0.14),
which are taken from Lee et al. (2011; Table S1 in the supporting information). Cratonic geotherms (red dashed lines) are
taken from Hasterok and Chapman (2011). The graphite-diamond transition (red solid line) was calculated using
THERMOCALC software. The colorful compositional isopleths were contoured using Surfer software. Mineral assemblage in
garnet peridotite field: Garnet-olivine-orthopyroxene-clinopyroxene; in Grt-Spl peridotite field: Garnet-spinel-olivine-
orthopyroxene-clinopyroxene; in spinel peridotite field: Spinel-olivine-orthopyroxene-clinopyroxene. In (c) and (d), XCa and
2
Cr# values in garnets from dunite, harzburgite, lherzolite (KLB-1), and wehrlite were computed along the 40 mW/m
geotherm.

EMOD (enstatite + magnesite = olivine + diamond + O2) buffer reactions (stippled field in Figure 4a; Luth &
Stachel, 2014). Therefore, the Fe3+/∑Fe ratios of the diamond-bearing cratonic peridotites are estimated to
range from ~0.3% to 3%, which are equivalent to bulk Fe2O3 values of ~0.03–0.24 wt %. As shown in
Figure 4b, the XCa and Cr# values of garnet in the diamond stability field are only weakly dependent on
bulk Fe3+/∑Fe ratios.

4.2. Orthopyroxene (SiO2) Enrichment


Cratonic peridotites metasomatized by silica-rich melts commonly result in orthopyroxene enrichment and a
decrease in the bulk Mg/Si ratio for a given Mg# (Herzberg, 1993; Kelemen et al., 1998). We model this sce-
nario by adding 35% orthopyroxene whilst removing 35% olivine relative to the average composition of cra-
tonic peridotites. The Fe3+/∑Fe values of orthopyroxene and olivine are assumed to be 0.05 and 0,
respectively (Canil & O’Neill, 1996); the bulk Fe3+/∑Fe value increases from 1.3% to 2.5% accordingly with
orthopyroxene enrichment. However, the oxygen fugacity remains almost unchanged due to the strong

SU AND CHEN 7692


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

3+ 3+
Figure 4. Calculated P-Fe /∑Fe pseudosection showing the effect of Fe contents on oxygen fugacity (a) and garnet
compositions (b) at 1100 °C. Fe2O3 content is 0 at x = 0 and 0.5 wt % at x = 1. Stippled field represents the diamond sta-
bility field. The upper oxygen fugacity limit for diamond/graphite is defined by EMOD/G reaction (Eggler & Baker, 1982),
and the reducing side is limited by IW reaction (Ballhaus et al., 1991).

redox buffering capacity of orthopyroxene-rich peridotites (Figure 5a), which is supported by the similar
oxygen fugacities observed in orthopyroxene-rich and normal cratonic xenoliths (Frost & McCammon,
2008; Luth & Stachel, 2014). Since the Al2O3 and CaO contents in orthopyroxene are much lower than
those in cratonic peridotites, orthopyroxene enrichment will not significantly change the bulk CaO/Al2O3
value. The Cr# in orthopyroxene (average: 28.5; Stachel & Harris, 2008) is higher than that in cratonic
peridotites (average: 19.9; Table S1 in the supporting information); thus, the bulk Cr# will increase slightly

Figure 5. P-X pseudosections at 1100 °C showing orthopyroxene enrichment (a), clinopyroxene enrichment (b), and garnet enrichment (c). In the diagram below,
changes of phase proportions with compositional variations are calculated at 1100 °C and 50 kbar. In (c), densities of garnet-rich peridotites (orange solid line)
were calculated following the method of Chen et al. (2013). The density derived from the preliminary reference Earth model (Dziewonski & Anderson, 1981) is
3
3.367 g/cm at 50 kbar (orange dashed line).

SU AND CHEN 7693


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

after orthopyroxene enrichment. In Figure 5a, the Cr# in garnet increases from ~21 to 24 in the diamond
stability field, seemingly affected by the bulk Cr#. In contrast, the garnet modal and XCa values are not signifi-
cantly affected by orthopyroxene enrichment.

4.3. Clinopyroxene (CaO) Enrichment


Clinopyroxene enrichment is commonly observed in the subcontinental lithospheric mantle, especially when
it has been infiltrated by carbonatite melts (Rudnick et al., 1993; Yaxley et al., 1998). Carbonatite melt
infiltration could increase the bulk CaO/Al2O3 and Mg# values and produce wehrlite with high melt/rock
ratios. To simplify the calculation, this process was simulated by the addition of abundant clinopyroxene to
a clinopyroxene-free harzburgite from South Africa (sample FRB1402; Boyd et al., 1993). The Fe3+/∑Fe ratio
of clinopyroxene is set as 0.27 according to Woodland (2009); then, the bulk Fe3+/∑Fe ratio will increase
accordingly (from 1.3% to 8.3%) with increasing proportions of clinopyroxene. Consequently, diamond will
transform into carbonate when the oxygen fugacity rises over the EMOD reaction (Figure 5b). Increasing
the bulk CaO/Al2O3 value increases the XCa values in garnet, with the magnitude of the effect decreasing
as a result of clinopyroxene buffering. Likewise, the Cr# in garnet increases with the enrichment of clinopyr-
oxene, which has a high Cr# (40.9; Stachel & Harris, 2008). Therefore, a positive correlation will exist between
XCa and Cr# in garnet (i.e., the lherzolitic trend; Sobolev et al., 1973) when cratonic harzburgites experience
clinopyroxene enrichment. Furthermore, our simulation shows that adjusting the bulk Mg# (from 89 to 94)
has a limited effect on the oxygen fugacity and garnet composition (XCa and Cr#; Figure S2 in the
supporting information).

4.4. Garnet (Al2O3) Enrichment


Depleted cratonic peridotites typically contain ~5–15% garnet, which is higher than the garnet contents of
experimental high-degree melting residues (Herzberg, 2004; Walter, 1998). The enrichment of garnet in cra-
tonic peridotites remains enigmatic (Gibson, 2017; Malkovets et al., 2007). As garnet has high Al2O3 (~15–
22 wt %) but low CaO (<7 wt %) contents, its enrichment will reduce the bulk CaO/Al2O3 value. Figure 5c
shows that oxygen fugacity will not change with variations in the garnet mode (Fe3+/∑Fe = 0.08;
Woodland & Koch, 2003). The XCa in garnet decreases with increasing proportions of garnet; conversely,
the Cr# in garnet shows a slight increase. The opposite trend between XCa and Cr# is likely induced by the
garnet composition used in this modeling (Stachel & Harris, 2008), as garnet yields higher Cr# (23.8 vs.
19.8) and lower CaO/Al2O3 (0.13 vs. 0.73) values than the representative whole-rock composition of
cratonic peridotites.

4.5. Cr# Variations


The above calculations show that the Cr# in garnet is expected to correlate with the bulk Cr#. We test
this hypothesis in Figure 6, which shows an isothermal (1100 °C) P-Cr# pseudosection with changes in
the bulk Cr2O3 and Al2O3 contents (their sum is constant) in cratonic peridotite. In the spinel-free garnet
peridotite field, the Cr# values of garnet sharply increase with bulk Cr# values (Figure 6a). However, in the
garnet-spinel peridotite field, the Cr# in garnet is pressure sensitive and composition insensitive. The XCa
in garnet also increases toward the right edge of Figure 6a due to the increase in the bulk
CaO/Al2O3 ratio.
To further investigate the effect of the bulk Cr#, we calculate the garnet Cr# as a function of the bulk Cr# along
the normal cratonic geotherm (Hasterok & Chapman, 2011). As shown in Figure 6b, when spinel appears, cra-
tonic peridotites with different bulk Cr# values have similar garnet Cr# values under the same pressure. This
result is applicable to various peridotites (Figure 3d), demonstrating the potential use of the garnet Cr# as a
geobarometer for garnet-spinel peridotites. Once spinel is exhausted, the Cr# in garnet is only related to the
bulk Cr# and changes very little with depth.
The above open-system simulations indicate that the XCa and Cr# values in garnet are mainly controlled
by the bulk CaO/Al2O3 and Cr# values, respectively, in the spinel-free field. In the following section, we
establish a quantitative relationship between whole-rock compositions and these two parameters
in garnet.

SU AND CHEN 7694


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Figure 6. (a) P-Cr# pseudosection at 1100 °C showing the effect of Cr# variations on garnet compositions. (b) Cr# values in
2
garnet for the average cratonic peridotite composition were calculated along the 40 mW/m geotherm for different bulk
Cr# values (5–45).

5. The Quantitative Relationship Between Garnet and Whole-Rock Compositions


5.1. Quantitative Relationship
Since garnet is the dominant Al2O3- and Cr2O3-bearing phase in garnet peridotite, the Cr# value in garnet is
sensitive to the bulk Cr# (Figure 6), and the modal proportion of garnet is controlled by the bulk Al2O3 con-
tent. Bulk CaO is mainly stored in clinopyroxene, with minor amounts stored in garnet; yet, garnet accounts
for a higher percentage of bulk CaO in garnet harzburgite. Therefore, according to our modeling results,
decreasing the bulk CaO/Al2O3 will lower the XCa value in garnet (Figure 5). To quantify their correlations
along the normal cratonic geotherm, we plot data at 1100 °C and 50 kbar (close to the average P-T condition
of diamond inclusions of the peridotitic suite; Stachel & Luth, 2015) from the open-system simulations
(Figures 5 and 6) and then rearrange them based on their garnet XCa and Cr# values (gray lines in
Figures 7a and 7b). The best fit is shown by the red lines, which can be expressed by the following formulas.
In Figure 7a, we fit two polynomials, where x is the bulk CaO/Al2O3 value (mass ratio) and y is the 100*XCa
value in garnet:

y ¼ 24:64x 2 þ 33:4x ðif x≤0:3Þ (1a)

y ¼ 0:1ð lnx Þ4 þ 0:06ð lnx Þ3  0:95ð lnx Þ2 þ 3lnx þ 13 ðif x > 0:3Þ (1b)

Furthermore, we also present the inverse formulas, with which the bulk CaO/Al2O3 value can be calculated
using the 100*XCa value in garnet:

x ¼ 0:07y 4  1:2y 3 þ 8:5y 2 þ 15y =1; 000 ðif y≤13Þ (2a)

x ¼ 0:08y 3  3:134y 2 þ 41:26y  181:5 ðif y > 13Þ (2b)

In Figure 7b, the Cr# in garnet clearly shows a linear relationship with bulk Cr#:

Cr#Grt ¼ 1:04Cr#bulk (3)

It should be emphasized that equation (3) is only valid in spinel-free peridotites, which are representative of
most cratonic mantle roots (Pearson & Wittig, 2014). As the XCa and Cr# values in garnet remain nearly con-
stant along the 40 mW/m2 geotherm (Figure 3), the compositions of incipient cratonic peridotites can be

SU AND CHEN 7695


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

2
Figure 7. (a, b) The quantitative relations between whole-rock and garnet compositions along the 40 mW/m geotherm.
The data of gray lines are collected from the open-system simulations at 1100 °C and 50 kbar (Figures 5 and 6). The best
fits to the data are represented by the red curves. (c, d) Garnet XCa and Cr# of cratonic peridotites calculated using bulk rock
compositions versus the compositions measured by electron probe (Alubach et al., 2007; Bizzarro & Stevenson, 2003; Boyd
et al., 1993; Boyd et al., 1997; Gibson et al., 2008, 2013; Ionov et al., 2010; Schulze, 1995; van Achterbergh et al., 2001).

computed using these formulas on the basis of garnet inclusions in diamond, since diamond inclusions had
reached subsolidus re-equilibration along the normal cratonic geotherm (38–42 mW/m2) before being
trapped in diamond (Stachel & Harris, 2008). Moreover, the XCa and Cr# values in garnet along the 35 and
45 mW/m2 geotherms are also nearly constant (Figure S3 in the supporting information), and they
similarly exhibit limited differences (100*XCa ± 1 and Cr# ± 3) to those along the 40 mW/m2 geotherm.
5.2. Reliability Evaluation
Reliability is evaluated by how well natural samples are reproduced by equations (1a), (1b), and (3) using the
reported whole-rock and garnet compositions. As the garnet compositions in equations (1a), (1b), and (3) are
simulated under P-T conditions along the 40 mW/m2 geotherm, we collect data for spinel-free garnet perido-
tites from the Siberian, South African, Tanzanian, North Atlantic, Slave, and Botswanan cratons, which have
continental lithosphere with similar thermal states (35–45 mW/m2). All data are given in the supporting infor-
mation (Table S3). Garnet compositions are calculated using bulk compositions and then compared with the
mineral compositions measured by electron probe. As shown in Figures 7c and 7d, most of the cratonic peri-
dotites plot well along the 1:1 correspondence line, further demonstrating that equations (1a), (1b), and (3)
are reliable for natural samples. Therefore, the initial compositional signatures of cratonic mantle roots can
be inferred based on the compositions of garnet inclusions in diamond.

6. Phase Equilibrium Modeling of Mantle Melting


Previous studies have proposed a negative correlation between bulk Cr2O3/Al2O3 and the pressure of mantle
partial melting (Aulbach, 2012; Stachel et al., 1998). Hence, the bulk Cr# in cratonic peridotites can be used as

SU AND CHEN 7696


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Figure 8. Cr#-CaO/Al2O3 diagrams for model mantle residues formed by isobaric (a) and polybaric (b) fractional melting of
fertile peridotite KLB-1 (Davis et al., 2009). See text for discussion.

a potential indicator of melting pressure. The Al2O3 contents of residues generated by different melting
mechanisms have been well established (e.g., Herzberg, 2004; Niu, 1997; Walter, 1998); however, the
situation for Cr2O3 is not clear. Thus, the behavior of the mantle Cr# during melting under various
pressures is still poorly constrained.
The melt model of Jennings and Holland (2015) was constructed based on experimental studies of the partial
melting of fertile peridotite (KLB-1). The modeled solidus and mineral modes along the solidus are consistent
with experimental results and pMELTS calculations (Jennings & Holland, 2015). Thermodynamic modeling
using such a melt model can also reproduce experimental melt compositions and fractions over a wide range
of pressures (Jennings et al., 2016; Jennings & Holland, 2015), thus complementing the pMELTS model, which
has an upper pressure limit of 3 GPa (Ghiorso et al., 2002). In particular, the melt model of Jennings and
Holland (2015) is very effective at low melt fractions, making it available for calculating the fractional melting
of the mantle.
Mantle melting tends to be fractional, as melt extraction is efficient in the mantle (Iwamori, 1993; Kelemen
et al., 1997). Therefore, the thermodynamic simulations in this study focus on isobaric and polybaric fractional
melting. As seen in Figure 8, the Cr# values in the residues are very sensitive to the melting pressure when the
melt fractions are greater than ~15%. Since Cr is a moderately compatible element in the bulk mantle, the
variations in Cr# are mainly controlled by bulk Al2O3 content. At low pressures (e.g., <2 GPa), Al2O3 is primar-
ily hosted in spinel (or plagioclase at <~1 GPa) and clinopyroxene, both of which preferentially melt at the
onset of melting (Falloon et al., 1999; Takahashi, 1986); thus, the Al2O3 content in the residue will decrease
rapidly. In contrast, Al2O3 is dominated by garnet at high pressures (e.g., >4 GPa). The retention of garnet dur-
ing the deep melting (e.g., F < ~30%) of peridotite (Takahashi et al., 1993; Walter, 1998) will result in only a
slight decrease in bulk Al2O3 content. Therefore, Cr# of the residue is a powerful indicator of mantle melting
pressure. The melting modeling also shows that CaO/Al2O3 generally decreases with melting, except during
low-pressure initial melting (Figure 8). The increase in CaO/Al2O3 during low-pressure initial melting is due to
the preferential dissolution of spinel (or plagioclase) relative to clinopyroxene. The modeling of isobaric batch
melting yields similar trends to the above polybaric modeling (Figure S4 in the supporting information).
Therefore, the combination of Cr# and CaO/Al2O3 values provides a new approach for determining the melt-
ing conditions of lithospheric mantle (Figure 8).

7. Making Cratonic Lithospheric Mantle


7.1. Low-Pressure Melting of Incipient Cratonic Mantle
The formation of cratonic lithospheric mantle at high or low melting pressures has been debated for over 30
years (Figure 1; e.g., Arndt et al., 2009; Aulbach, 2012; Boyd, 1989; Carlson et al., 2005; Griffin et al., 2003;
Herzberg, 2004; Herzberg & Rudnick, 2012; Lee, 2006; Pearson & Wittig, 2008). Most previous approaches

SU AND CHEN 7697


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Figure 9. Bulk Cr# versus CaO/Al2O3 for the incipient cratonic mantle calculated from garnet inclusions in diamond from
South African, Slave, Siberian, Tanzanian and other cratons (data sources are same to Figure 2a). The histogram on right
hand axis is the distribution of whole-rock Cr# values inferred from the garnet inclusions. The colorful curves represent the
compositional ranges of isobaric fractional melting residues as shown in Figure 8a.

have been based on the chemical compositions of cratonic mantle xenoliths that commonly underwent
refertilization after melt extraction (e.g., Foley, 2008; Rudnick et al., 1993; Yaxley et al., 1998).
Garnet inclusions in diamond enable us to sample the roots of cratonic mantle; these inclusions have been
protected from ubiquitous metasomatism and subsolidus reequilibration with the surrounding matrix since
diamond formation (Boyd et al., 1985; Shirey & Richardson, 2011). However, experimental studies show that
garnet is initially exhausted in depleted harzburgite residues after high-degree melting (Herzberg, 2004;
Walter, 1998). Thus, the origin of garnet in diamond is an essential problem that needs to be addressed.
One model is that garnet formed by reaction of infiltrating liquids with host peridotites (Bell et al., 2005;
Malkovets et al., 2007). However, chemical data of garnet inclusions in diamond indicate that metasomatic
processes only enriched most incompatible trace elements and left major element composition largely unaf-
fected (Stachel et al., 2004; Stachel & Harris, 2008; Taylor et al., 2003). Alternatively, garnet might grow from
isochemical processes, including exsolution from primary Al-rich orthopyroxene (Gibson, 2017; Simon et al.,
2003) or high-pressure transformation of spinel during lithospheric thickening (Stachel et al., 1998; Tainton &
McKenzie, 1994). For all above-mentioned origins, garnet compositions from diamond inclusions most likely
reflect true residue compositions (at least major elements) of pristine cratonic mantle and thus can be used to
infer melting pressure.
As simulated above, the garnet Cr# is positively correlated with the bulk Cr# in spinel-free assemblages
(Figure 7b), and the latter is mainly controlled by the melting pressure (Figure 8). Accordingly, the relation-
ship between garnet and melting pressure can be constructed. As shown in Figure 9, the bulk compositions
of pristine cratonic mantle calculated from garnet inclusions in diamond using equations (2a), (2b), and (3)
mostly (~85%) plot between the 5- and 40-kbar melting curves. The garnet data from different cratons
(e.g., South Africa, Slave, Siberia, and Tanzania) do not show clear distinctions. The overall high Cr# values
of global garnet inclusions (mostly 12–36) suggest that cratonic mantle roots were originally formed by shal-
low melt extraction (<~4.0 GPa) prior to migration to the diamond stability field. In addition, few garnet inclu-
sions with low Cr# values (<10) plot below the 40-kbar melting curve (Figure 9), probably because the Al2O3
content (3.5 wt %) of fertile upper mantle (KLB-1) used in this study is slightly lower than the other estimated
Al2O3 content of the primitive mantle (3.5–4.5 wt %; Walter, 2003).
Compared with garnet inclusions in diamond, garnet xenocrysts in kimberlites have relatively lower Cr2O3
contents (Figure 2b) and thus lower Cr# values (mostly 5–27). Moreover, most garnet xenocrysts plot in the

SU AND CHEN 7698


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Figure 10. Probability histograms for bulk Cr# and CaO/Al2O3 in cratonic mantle calculated from garnet inclusions in dia-
mond (a, e), cratonic peridotite xenoliths (b, f), SSZ mantle peridotites (c, g), and MOR mantle peridotites (d, h). In (a) and (e),
cratonic mantle compositions were calculated using equations (2a), (2b), and (3). SSZ mantle peridotites include SSZ
ophiolitic peridotites and forearc peridotites. MOR mantle peridotites consist of abyssal peridotites and MOR ophiolitic
peridotites. Data sources: Cratonic peridotite xenoliths (Carlson et al., 2005; Rudnick et al., 1993; Simon et al., 2003; Wittig
et al., 2008; Lee et al., 2011, and reference therein) and forearc peridotites (Bénard et al., 2017; Ohara & Ishii, 1998; Parkinson
& Pearce, 1998; Pearce et al., 2000; Savov et al., 2005; Savov et al., 2007). SSZ ophiolitic peridotites (Batanova et al., 2011;
Büchl et al., 2002; Habtoor et al., 2017; Khalil & Azer, 2007; Khedr & Arai, 2017; Liu et al., 2016; Mahéo et al., 2004; Marchesi
et al., 2009; Moghadam et al., 2015; O’Driscoll et al., 2012; Pagé et al., 2009; Rajabzadeh & Dehkordi, 2013; Saka et al.,
2014; Uysal et al., 2012; Uysal et al., 2015; Uysal et al., 2016; Zhang et al., 2016; Zhou et al., 1996), abyssal peridotites (Aumento
& Loubat, 1971; Brandon et al., 2000; Cannat et al., 1995; Casey, 1997; Coogan et al., 2004; Godard et al., 2008; Harvey et al.,
2006; Miyashiro et al., 1969; Niu, 2004; Niu & Hékinian, 1997; Paulick et al., 2006; Snow & Dick, 1995; Stephens, 1997), and
MOR ophiolitic peridotites (Austrheim & Prestvik, 2008; Beccaluva et al., 1984; Godard et al., 2000; Hanghøj et al., 2010; Iyer
et al., 2008; li & lee, 2006; Liu et al., 2014; Maaløe, 2005; Marchesi et al., 2006; Ottonello et al., 1984; Uysal et al., 2016; Yumul Jr
et al., 2008). SSZ = suprasubduction zone; MOR = mid-ocean ridge.

field of lherzolitic garnets with high XCa (Figure 2b). If these xenocrysts originated from refractory cratonic
mantle, their host rocks appear to have suffered extensive metasomatism by Ca-rich and Al-rich melts
(Ziberna et al., 2013). This would reduce the Cr# values of host rocks, thus causing the melting pressures
inferred from garnet xenocrysts to be overestimated. Likewise, the majority of cratonic mantle xenoliths
have higher CaO/Al2O3 and lower Cr# values than the values calculated from garnet inclusions in diamond
(Figure 10). This further implies the widespread refertilization of cratonic mantle xenoliths (e.g., Foley,
2008; Rudnick et al., 1993; Yaxley et al., 1998). The melting pressures estimated from the Cr# or Cr2O3/
Al2O3 values of metasomatic peridotite xenoliths would thus be meaningless (e.g., Aulbach, 2012). By
contrast, some cratonic mantle xenoliths free from metasomatism, such as orthopyroxene-poor
harzburgites from the Udachnaya kimberlite (Doucet et al., 2012), are close in composition to pristine
melting residues and have high Cr# values (14–65), which are very similar to those estimated from the
garnet inclusions in this study. The low FeO contents (<7 wt %) of many cratonic peridotites are
commonly cited as evidence for high-pressure melting (e.g., Griffin et al., 2003; Pearson et al., 1995). The

SU AND CHEN 7699


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Figure 11. Cr# versus 100*XCa diagrams for garnets growing from SSZ peridotites (a) and MOR peridotites (b). The garnet
compositions are calculated using bulk compositions with equations (1a), (1b), and (3). The gray region represents the
compositional range of garnet inclusions in diamond. Data sources are same to Figure 10. SSZ = suprasubduction zone;
MOR = mid-ocean ridge.

problem, however, is that low FeO peridotites generally contain excess orthopyroxene owing to melt-rock
reaction (e.g., Kelemen et al., 1998), as orthopyroxene enrichment plays a role in diluting bulk FeO content
(Lee, 2006; Lee et al., 2011; Pearson & Wittig, 2008). In addition to orthopyroxene-rich peridotites, other
cratonic peridotites have higher FeO contents in the 7 to 8 wt % range, corresponding to initial melting
pressures at 3 to 5 GPa and finial melting pressure as low as 1 GPa (Herzberg, 2004; Herzberg et al., 2010),
very similar to our results.
In conclusion, garnet inclusions in diamond can adequately preserve the compositional characteristics of inci-
pient cratonic mantle; their high Cr# values clearly reveal that the origin of cratonic mantle involves low-
pressure (<~4 GPa) melt extraction. Below, we attempt to constrain the possible tectonic environment in
which this melting took place.

7.2. Geodynamic Environment of Cratonic Mantle Formation


The formation of cratonic mantle at low pressures is ascribed to melting at either mid-ocean ridges or supra-
subduction zones (Canil et al., 2004; Kelemen et al., 1998; Lee et al., 2006; Stachel et al., 1998; Wittig et al.,
2008), followed by tectonic stacking/accretion with high-pressure garnet growth (Figures 1b and 1c). Either
tectonic scenario should produce garnet compositions similar to those of garnet inclusions in diamond.
The high degree of melting of cratonic peridotites at relatively low pressures has led some researchers to pro-
pose that cratonic mantle was formed by hydrous melting in subduction zones (e.g., Carlson et al., 2005;
Parman et al., 2004; Pearson & Wittig, 2008). This model may explain the excess orthopyroxene observed
in some Kaapvaal peridotites and the enrichment of incompatible trace elements in diamond inclusions
due to the infiltration of slab-derived liquids (Kelemen et al., 1998; Stachel et al., 1998). However, suprasub-
duction zone peridotites (hereafter referred to as SSZ peridotites), such as SSZ ophiolitic peridotites and fore-
arc peridotites, generally have higher CaO/Al2O3 (mostly >0.5; Figure 10g) and Cr# (more than 50% have Cr#
>30; Figure 10c) values than the cratonic mantle values calculated from garnet inclusions in diamond (mostly
16–30 Cr# and < 0.5 CaO/Al2O3; Figures 10a and 10e). If SSZ peridotites (i.e., arc lithosphere) represent the
initial lithospheric mantle and were then thickened to form cratonic lithospheric mantle, the garnets forming
in such peridotites would have higher Cr# and XCa values than garnet inclusions in diamond (Figure 11a). One
hypothesis is that SSZ peridotites were metasomatized by Al-rich melts during the accretion process, thus
reducing their Cr# and CaO/Al2O3 values. Nevertheless, as shown in Figure 5c, Al2O3 enrichment would
rapidly increase the garnet mode and rock density, thus making the cratonic mantle gravitationally unstable.
Moreover, although hydrous melting could promote more extensive melt extraction, the anhydrous nature of
cratonic mantle required for its long-term stability certainly contradicts the SSZ model (Herzberg &
Rudnick, 2012).

SU AND CHEN 7700


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

In contrast, peridotites from modern mid-oceanic ridges (hereafter referred to as MOR peridotites) exhibit
a similar range of Cr# values as the calculated cratonic peridotites, except that the peak of the former
extends to lower values (Figure 10d). These peridotites also exhibit a consistent distribution of bulk
CaO/Al2O3 ratios (Figures 10h and 10e). Nisbet et al. (1993) suggested that mantle potential temperatures
(TP) in the Archean (~1500–1600 °C) were higher than that of the modern mantle (1300–1400 °C); thus, the
average degree of partial melting of ambient mantle would increase from 8–10% at modern oceanic
ridges to ~30% in the ancient Earth. Therefore, the Cr# and CaO/Al2O3 values of hot MOR peridotites
would increase and decrease, respectively, becoming more similar to those of the incipient cratonic
mantle. In Figure 11b, the garnets growing from MOR peridotites display similar compositions to those
in diamond, implying a potentially genetic relationship between oceanic lithospheric mantle and cratonic
lithospheric mantle. Previous studies have noted that a major problem with the MOR model is the
deviation of the lithologies and compositions of oceanic lithosphere from those of cratonic lithosphere
(Griffin et al., 2003; Pearson & Wittig, 2008). The modern oceanic lithosphere comprises fertile, low-Mg#
lherzolite at the bases of ridges, with harzburgite at the top, whereas cratonic mantle mainly consists of
high Mg# harzburgite. Our modeling shows that there is a positive linear relationship between the residual
Mg# and the degree of melting (F), and melting pressures have little effect on the Mg# values of residues
(Figure S5 in the supporting information). As computed by Herzberg and Rudnick (2012), the hotter
Archean TP would result in greater melt extraction and generate the craton-wide distribution of
harzburgitic residue (Mg#: 92–94) beneath oceanic ridges. It thus seems plausible that cratonic lithospheric
mantle initially melted at Archean hot oceanic ridges and subsequently grew through the stacking of
oceanic lithosphere; garnet might reequilibrate or grow in the deeper lithospheric mantle and was then
partially trapped in diamond.
The MOR model not only interprets the origin of cratonic mantle during shallow-level melting processes but
also implies a genetic relationship between ancient oceanic lithospheric mantle and cratonic lithospheric
mantle, which may play a vital role in the formation and evolution of continental and oceanic plates at early
stage (Ben-Avraham et al., 1981; Carlson et al., 2005). Our phase equilibria calculations link the compositions
of garnet and mantle peridotite, allowing us to obtain the bulk composition of cratonic mantle using a single
garnet grain. Therefore, thermodynamic modeling approach in this study opens up a new avenue for our
understanding of the evolution of the Earth’s mantle.

8. Conclusions
The quantitative relationship between garnet and cratonic peridotite was estimated using thermodynamic
modeling with THERMOCALC software. The incipient cratonic mantle calculated using garnet inclusions in
diamond has more depleted compositions (higher Cr# and lower CaO/Al2O3) than cratonic mantle xenoliths.
The high Cr# values of garnet inclusions in diamond support a low-pressure (<4 GPa) origin of cratonic litho-
spheric mantle. The similarities between the whole-rock compositions (Cr# and CaO/Al2O3) of peridotites
from mid-oceanic ridges and the calculated cratonic peridotites support the transition from oceanic litho-
sphere to cratonic lithosphere in the early Earth. We propose that cratonic lithospheric mantle was initially
formed by the extensive melting of hot ambient mantle in ocean ridge settings and its subsequent thicken-
Acknowledgments
ing caused by the stacking of oceanic lithosphere.
This work was funded by the National
Science Foundation of China (41490614
and 41822202), the National Basic
References
Research Program of China (973 Alubach, S., Griffin, W. L., Pearson, N. J., O’Reilly, S. Y., & Doyle, B. J. (2007). Lithosphere formation in the central Slave craton (Canada): Plume
Program 2015CB856103), the National subcretion or lithosphere accretion? Contributions to Mineralogy and Petrology, 154(4), 409–427. https://doi.org/10.1007/s00410-007-
Postdoctoral Program for Innovative 0200-1
Talents (BX201700239), and China Arndt, N. T., Coltice, N., Helmstaedt, H., & Gregoire, M. (2009). Origin of Archean subcontinental lithospheric mantle: Some petrological
Postdoctoral Science Foundation constraints. Lithos, 109(1-2), 61–71. https://doi.org/10.1016/j.lithos.2008.10.019
(2017M620065). We thank Claude Aulbach, S. (2012). Craton nucleation and formation of thick lithospheric roots. Lithos, 149, 16–30. https://doi.org/10.1016/j.
Herzberg and an anonymous reviewer lithos.2012.02.011
for their constructive and expeditious Aumento, F., & Loubat, H. (1971). The Mid-Atlantic ridge near 45°N. XVI. Serpentinized ultramafic intrusions. Canadian Journal of Earth
reviews, and Michael Walter for his Sciences, 8(6), 631–663. https://doi.org/10.1139/e71-062
editorial handling. This paper is dedi- Austrheim, H., & Prestvik, T. (2008). Rodingitization and hydration of the oceanic lithosphere as developed in the Leka ophiolite, north–
cated in memoriam to my respectable central Norway. Lithos, 104(1-4), 177–198. https://doi.org/10.1016/j.lithos.2007.12.006
supervisor Kai Ye. The data used in this Ballhaus, C., Berry, R. F., & Green, D. H. (1991). High pressure experimental calibration of the olivine–orthopyroxene–spinel oxygen geoba-
paper are available via the supporting rometer: Implications for the oxidation state of the upper mantle. Contributions to Mineralogy and Petrology, 107(1), 27–40. https://doi.org/
information and cited references. 10.1007/BF00311183

SU AND CHEN 7701


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Batanova, V., Belousov, I., Savelieva, G., & Sobolev, A. (2011). Consequences of channelized and diffuse melt transport in supra-subduction
zone mantle: Evidence from the Voykar ophiolite (Polar Urals). Journal of Petrology, 52(12), 2483–2521. https://doi.org/10.1093/petrology/
egr053
Beccaluva, L., Macciotta, G., Piccardo, G., & Zeda, O. (1984). Petrology of lherzolitic rocks from the northern Apennine ophiolites. Lithos, 17,
299–316. https://doi.org/10.1016/0024-4937(84)90027-6
Bell, D. R., Grégoire, M., Grove, T., Chatterjee, N., Carlson, R., & Buseck, P. (2005). Silica and volatile-element metasomatism of Archean mantle:
A xenolith-scale example from the Kaapvaal craton. Contributions to Mineralogy and Petrology, 150(3), 251–267. https://doi.org/10.1007/
s00410-005-0673-8
Bénard, A., Arculus, R. J., Nebel, O., Ionov, D., & McAlpine, S. (2017). Silica-enriched mantle sources of subalkaline picrite–boninite–andesite
island arc magmas. Geochimica et Cosmochimica Acta, 199, 287–303. https://doi.org/10.1016/j.gca.2016.09.030
Ben-Avraham, Z., Nur, A., Jones, D., & Cox, A. (1981). Continental accretion: From oceanic plateaus to allochthonous terranes. Science,
213(4503), 47–54. https://doi.org/10.1126/science.213.4503.47
Bizzarro, M., & Stevenson, R. (2003). Major element composition of the lithospheric mantle under the North Atlantic craton: Evidence from
peridotite xenoliths of the Sarfartoq area, southwestern Greenland. Contributions to Mineralogy and Petrology, 146(2), 223–240. https://doi.
org/10.1007/s00410-003-0499-1
Boyd, F., Pokhilenko, N., Pearson, D., Mertzman, S., Sobolev, N., & Finger, L. (1997). Composition of the Siberian cratonic mantle: Evidence
from Udachnaya peridotite xenoliths. Contributions to Mineralogy and Petrology, 128(2-3), 228–246. https://doi.org/10.1007/
s004100050305
Boyd, F. R. (1989). Compositional distinction between oceanic and cratonic lithosphere. Earth and Planetary Science Letters, 96(1-2), 15–26.
https://doi.org/10.1016/0012-821X(89)90120-9
Boyd, F. R., Gurney, J. J., & Richardson, S. H. (1985). Evidence for a 150–200-km thick Archean lithosphere from diamond inclusion thermo-
barometry. Nature, 315(6018), 387–389. https://doi.org/10.1038/315387a0
Boyd, F. R., Pearson, D. G., Nixon, P. H., & Mertzman, S. A. (1993). Low-calcium garnet harzburgites from southern Africa: Their relations to
craton structure and diamond crystallization. Contributions to Mineralogy and Petrology, 113(3), 352–366. https://doi.org/10.1007/
BF00286927
190 186 187 187
Brandon, A., Snow, J., Walker, R., Morgan, J., & Mock, T. (2000). Pt– Os and Re– Os systematics of abyssal peridotites. Earth and
Planetary Science Letters, 177(3-4), 319–335. https://doi.org/10.1016/S0012-821X(00)00044-3
Büchl, A., Brügmann, G., Batanova, V. G., Münker, C., & Hofmann, A. W. (2002). Melt percolation monitored by Os isotopes and HSE abun-
dances: A case study from the mantle section of the Troodos Ophiolite. Earth and Planetary Science Letters, 204(3-4), 385–402. https://doi.
org/10.1016/S0012-821X(02)00977-9
Canil, D. (2004). Mildly incompatible elements in peridotites and the origins of mantle lithosphere. Lithos, 77(1-4), 375–393. https://doi.org/
10.1016/j.lithos. 2004.04.014
Canil, D., & O’Neill, H. S. C. (1996). Distribution of ferric iron in some upper-mantle assemblages. Journal of Petrology, 37(3), 609–635. https://
doi.org/10.1093/petrology/37.3.609
Canil, D., O’Neill, H. S. C., Pearson, D. G., Rudnick, R. L., McDonough, W. F., & Carswell, D. A. (1994). Ferric iron in peridotites and mantle oxi-
dation states. Earth and Planetary Science Letters, 123(1-3), 205–220. https://doi.org/10.1016/0012-821X(94)90268-2
Cannat, M., Mevel, C., Maia, M., Deplus, C., Durand, C., Gente, P., et al. (1995). Thin crust, ultramafic exposures, and rugged faulting patterns at
the Mid-Atlantic Ridge (22–24°N). Geology, 23(1), 49–52. https://doi.org/10.1130/0091-7613
Carlson, R. W., Pearson, D. G., & James, D. E. (2005). Physical, chemical, and chronological characteristics of continental mantle. Reviews of
Geophysics, 43, RG1001. https://doi.org/10.1029/2004RG000156
Casey, J. F. (1997). Comparison of major and trace-element geochemistry of abyssal peridotites and mafic plutonics with basalts from the
MARK region of the Mid-Atlantic Ridge. Proceeding of the Ocean Drilling Program, Scientific Results, 153, 181–241.
Chen, Y., Ye, K., Wu, T. F., & Guo, S. (2013). Exhumation of oceanic eclogites: Thermodynamic constraints on pressure, temperature, bulk
composition and density. Journal of Metamorphic Geology, 31(5), 549–570. https://doi.org/10.1111/jmg.12033
Coogan, L. A., Thompson, G., MacLeod, C. J., Dick, H., Edwards, S., Scheirer, A. H., & Barry, T. L. (2004). A combined basalt and peridotite
perspective on 14 million years of melt generation at the Atlantis Bank segment of the Southwest Indian Ridge: Evidence for temporal
changes in mantle dynamics? Chemical Geology, 207(1-2), 13–30. https://doi.org/10.1016/j.chemgeo.2004.01.016
Davis, F. A., Tangeman, J. A., Tenner, T. J., & Hirschmann, M. M. (2009). The composition of KLB-1 peridotite. American Mineralogist, 94(1),
176–180. https://doi.org/10.2138/am.2009.2984
Doucet, L. S., Ionov, D. A., Golovin, A. V., & Pokhilenko, N. (2012). Depth, degrees and tectonic settings of mantle melting during craton
formation: Inferences from major and trace element compositions of spinel harzburgite xenoliths from the Udachnaya kimberlite, central
Siberia. Earth and Planetary Science Letters, 359-360, 206–218. https://doi.org/10.1016/j.epsl.2012.10.001
Dziewonski, A. M., & Anderson, D. L. (1981). Preliminary reference Earth model. Physics of the Earth and Planetary Interiors, 25(4), 297–356.
https://doi.org/10.1016/0031-9201(81)90046-7
Eggler, D. H., & Baker, D. R. (1982). Reduced volatiles in the system C–H–O: Implications to mantle melting, fluid formation and diamond
genesis. In S. Akimoto, & M. H. Manghnani (Eds.), High pressure research in geophysics, (pp. 237–250). Tokyo: Center for Academic
Publishing.
Falloon, T. J., Green, D. H., Danyushevsky, L. V., & Faul, U. H. (1999). Peridotite melting at 1.0 and 1.5 GPa: An experimental evaluation of
techniques using diamond aggregates and mineral mixes for determination of near-solidus melts. Journal of Petrology, 40(9), 1343–1375.
https://doi.org/10.1093/petroj/40.9.1343
Foley, S. F. (2008). Rejuvenation and erosion of the cratonic lithosphere. Nature Geoscience, 1(8), 503–510. https://doi.org/10.1038/ngeo261
Froelich, P. N., Bender, M. L., Luedtke, N. A., Heath, G. R., & DeVries, T. (1982). The marine phosphorus cycle. American Journal of Science, 282(4),
474–511. https://doi.org/10.2475/ajs.282.4.474
Frost, D. J., & McCammon, C. A. (2008). The redox state of Earth’s mantle. Annual Review of Earth and Planetary Sciences, 36(1), 389–420.
https://doi.org/10.1146/annurev.earth.36.031207.124322
Ghiorso, M. S., Hirschmann, M. M., Reiners, P. W., & Kress, V. C. (2002). The pMELTS: A revision of MELTS for improved calculation of phase
relations and major element partitioning related to partial melting of the mantle to 3 GPa. Geochemistry, Geophysics, Geosystems, 3(5), 304.
https://doi.org/10.1029/2001GC000217
Gibson, S. A. (2017). On the nature and origin of garnet in highly-refractory Archean lithospheric mantle: Constraints from garnet exsolved in
Kaapvaal craton orthopyroxenes. Mineralogical Magazine, 81(04), 781–809. https://doi.org/10.1180/minmag.2016.080.158
Gibson, S. A., Malarkey, J., & Day, J. (2008). Melt depletion and enrichment beneath the western Kaapvaal Craton: Evidence from Finsch
peridotite xenoliths. Journal of Petrology, 49(10), 1817–1852. https://doi.org/10.1093/petrology/egn048

SU AND CHEN 7702


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Gibson, S. A., McMahon, S., Day, J., & Dawson, J. (2013). Highly refractory lithospheric mantle beneath the Tanzanian craton: Evidence from
Lashaine pre-metasomatic garnet-bearing peridotites. Journal of Petrology, 54(8), 1503–1546. https://doi.org/10.1093/petrology/egt020
Girnis, A. V., & Brey, G. P. (1999). Garnet–spinel–olivine–orthopyroxene equilibria in the FeO–MgO–Al2O3–SiO2–Cr2O3 systems: II.
Thermodynamic analysis. European Journal of Mineralogy, 11(4), 619–636. https://doi.org/10.1127/ejm/11/4/0619
Godard, M., Jousselin, D., & Bodinier, J. L. (2000). Relationships between geochemistry and structure beneath a palaeo-spreading centre: A
study of the mantle section in the Oman ophiolite. Earth and Planetary Science Letters, 180(1-2), 133–148. https://doi.org/10.1016/S0012-
821X(00)00149-7
Godard, M., Lagabrielle, Y., Alard, O., & Harvey, J. (2008). Geochemistry of the highly depleted peridotites drilled at ODP sites 1272 and 1274
(fifteen-twenty fracture zone, mid-Atlantic ridge): Implications for mantle dynamics beneath a slow spreading ridge. Earth and Planetary
Science Letters, 267(3-4), 410–425. https://doi.org/10.1016/j.epsl.2007.11.058
Griffin, W. L., Fisher, N. I., Friedman, J., Ryan, C. G., & O’Reilly, S. Y. (1999). Cr-pyrope garnets in the lithospheric mantle. I. Compositional
systematics and relations to tectonic setting. Journal of Petrology, 40(5), 679–704. https://doi.org/10.1093/petroj/40.5.679
Griffin, W. L., O’Reilly, S. Y., Abe, N., Aulbach, S., Davies, R., Pearson, N., et al. (2003). The origin and evolution of Archean lithospheric mantle.
Precambrian Research, 127(1-3), 19–41. https://doi.org/10.1016/S0301-9268(03)00180-3
Griffin, W. L., O’Reilly, S. Y., Afonso, J. C., & Begg, G. C. (2009). The composition and evolution of lithospheric mantle: A re-evaluation and its
tectonic implications. Journal of Petrology, 50(7), 1185–1204. https://doi.org/10.1093/petrology/egn033
Grütter, H. S., Gurney, J. J., Menzies, A. H., & Winter, F. (2004). An updated classification scheme for mantle-derived garnet, for use by diamond
explorers. Lithos, 77(1-4), 841–857. https://doi.org/10.1016/j.lithos.2004.04.012
Gudmundsson, G., & Wood, B. J. (1995). Experimental tests of garnet peridotite oxygen barometry. Contributions to Mineralogy and Petrology,
119(1), 56–67. https://doi.org/10.1007/BF00310717
Habtoor, A. M., Ahmed, A. H., Akizawa, N., Harbi, H., & Arai, S. (2017). Chemical homogeneity of high-Cr chromitites as indicator for wide-
spread invasion of boninitic melt in mantle peridotite of BirTuluha ophiolite, Northern Arabian Shield, Saudi Arabia. Ore Geology Reviews,
90, 243–259. https://doi.org/10.1016/j.oregeorev.2017.03.010
Hanghøj, K., Kelemen, P. B., Hassler, D., & Godard, M. (2010). Composition and genesis of depleted mantle peridotites from the WadiTayin
Massif, Oman Ophiolite; major and trace element geochemistry, and Os isotope and PGE systematics. Journal of Petrology, 51(1-2),
201–227. https://doi.org/10.1093/petrology/egp077
Harvey, J., Gannoun, A., Burton, K., Rogers, N., Alard, O., & Parkinson, I. (2006). Ancient melt extraction from the oceanic upper mantle
revealed by Re–Os isotopes in abyssal peridotites from the Mid-Atlantic ridge. Earth and Planetary Science Letters, 244(3-4), 606–621.
https://doi.org/10.1016/j.epsl.2006.02.031
Hasterok, D., & Chapman, D. S. (2011). Heat production and geotherms for the continental lithosphere. Earth and Planetary Science Letters,
307(1-2), 59–70. https://doi.org/10.1016/j.epsl.2011.04.034
Helmstaedt, H., & Schulze, D. J. (1989). Southern African kimberlites and their mantle sample: Implications for Archean tectonics and litho-
sphere evolution. Proceedings of the Fourth International Kimberlite Conference, Perth. Geological Society of Australia Special Publication,
14, 358–368.
Herzberg, C. (1993). Lithosphere peridotites of the Kaapvaal craton. Earth and Planetary Science Letters, 120(1-2), 13–29. https://doi.org/
10.1016/0012-821X(93) 90020-A
Herzberg, C. (2004). Geodynamic information in peridotite petrology. Journal of Petrology, 45(12), 2507–2530. https://doi.org/10.1093/pet-
rology/egh039
Herzberg, C., Condie, K., & Korenaga, J. (2010). Thermal history of the Earth and its petrological expression. Earth and Planetary Science Letters,
292(1-2), 79–88. https://doi.org/10.1016/j.epsl.2010.01.022
Herzberg, C., & O’Hara, M. J. (2002). Plume-associated ultramafic magmas of Phanerozoic age. Journal of Petrology, 43(10), 1857–1883. https://
doi.org/10.1093/petrology/43.10.1857
Herzberg, C., & Rudnick, R. (2012). Formation of cratonic lithosphere: An integrated thermal and petrological model. Lithos, 149, 4–15. https://
doi.org/10.1016/j.lithos.2012.01.010
Holland, T. J. B., & Powell, R. (2011). An improved and extended internally consistent thermodynamic dataset for phases of petrological
interest, involving a new equation of state for solids. Journal of Metamorphic Geology, 29(3), 333–383. https://doi.org/10.1111/j.1525-
1314.2010.00923.x
Ionov, D. A., Doucet, L. S., & Ashchepkov, I. V. (2010). Composition of the lithospheric mantle in the Siberian craton: New constraints from
fresh peridotites in the Udachnaya-East kimberlite. Journal of Petrology, 51(11), 2177–2210. https://doi.org/10.1093/petrology/egq053
Iwamori, H. (1993). A model for disequilibrium mantle melting incorporating melt transport by porous and channel flows. Nature, 366(6457),
734–737. https://doi.org/10.1038/366734a0
Iyer, K., Austrheim, H., John, T., & Jamtveit, B. (2008). Serpentinization of the oceanic lithosphere and some geochemical consequences:
Constraints from the Leka Ophiolite Complex, Norway. Chemical Geology, 249(1-2), 66–90. https://doi.org/10.1016/j.chemgeo.2007.12.005
Jennings, E. S., & Holland, T. J. B. (2015). A simple thermodynamic model for melting of peridotite in the system NCFMASOCr. Journal of
Petrology, 56(5), 869–892. https://doi.org/10.1093/petrology/egv020
Jennings, E. S., Holland, T. J. B., Shorttle, O., Maclennan, J., & Gibson, S. A. (2016). The composition of melts from a heterogeneous mantle and
the origin of ferropicrite: Application of a thermodynamic model. Journal of Petrology, 57, egw065–egw2310. https://doi.org/10.1093/
petrology/egw065
Kasting, J. F., Eggler, D. H., & Raeburn, S. P. (1993). Mantle redox evolution and the oxidation state of the Archean atmosphere. The Journal of
Geology, 101(2), 245–257. https://doi.org/10.1086/648219
Kelemen, P. B., Hart, S. R., & Bernstein, S. (1998). Silica enrichment in the continental upper mantle via melt/rock reaction. Earth and Planetary
Science Letters, 164(1-2), 387–406. https://doi.org/10.1016/S0012-821X(98)00233-7
Kelemen, P. B., Hirth, G., Shimizu, N., Spiegelman, M., & Dick, H. J. B. (1997). A review of melt migration processes in the adiabatically
upwelling mantle beneath oceanic spreading ridges. Philosophical Transactions of the Royal Society A, 355(1723), 283–318. https://doi.org/
10.1098/rsta.1997.0010
Khalil, A., & Azer, M. (2007). Supra-subduction affinity in the Neoproterozoic serpentinites in the Eastern Desert, Egypt: Evidence from mineral
composition. Journal of African Earth Sciences, 49(4-5), 136–152. https://doi.org/10.1016/j.jafrearsci.2007.08.002
Khedr, M. Z., & Arai, S. (2017). Peridotite-chromitite complexes in the Eastern Desert of Egypt: Insight into Neoproterozoic sub-arc mantle
processes. Gondwana Research, 52, 59–79. https://doi.org/10.1016/j.gr.2017.09.001
Klemme, S. (2004). The influence of Cr on the garnet–spinel transition in the Earth’s mantle: Experiments in the system MgO–Cr2O3–SiO2 and
thermodynamic modelling. Lithos, 77(1-4), 639–646. https://doi.org/10.1016/j.lithos.2004.03.017
Lee, C. T. A. (2006). Geochemical/petrologic constraints on the origin of cratonic mantle. In K. Benn, J. Mareschal, & K. Condie (Eds.), Archean
geodynamics and environments, (pp. 89–114). Washington, D.C.: American Geophysical Union Monograph.

SU AND CHEN 7703


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Lee, C. T. A., & Chin, E. J. (2014). Calculating melting temperatures and pressures of peridotite protoliths: Implications for the origin of cratonic
mantle. Earth and Planetary Science Letters, 403, 273–286. https://doi.org/10.1016/j.epsl.2014.06.048
Lee, C. T. A., Luffi, P., & Chin, E. J. (2011). Building and destroying continental mantle. Annual Review of Earth and Planetary Sciences, 39(1),
59–90. https://doi.org/10.1146/annurev-earth-040610-133505
Li, Z. X. A., & Lee, C. T. A. (2006). Geochemical investigation of serpentinized oceanic lithospheric mantle in the Feather river ophiolite,
California: Implications for the recycling rate of water by subduction. Chemical Geology, 235(1-2), 161–185. https://doi.org/10.1016/j.
chemgeo.2006.06.011
Liu, C. Z., Zhang, C., Xu, Y., Wang, J. G., Chen, Y., Guo, S., et al. (2016). Petrology and geochemistry of mantle peridotites from the Kalaymyo
and Myitkyina ophiolites (Myanmar): Implications for tectonic settings. Lithos, 264, 495–508. https://doi.org/10.1016/j.lithos.2016.09.013
Liu, C. Z., Zhang, C., Yang, L. Y., Zhang, L. L., Ji, W. Q., & Wu, F. Y. (2014). Formation of gabbronorites in the Purang ophiolite (SW Tibet) through
melting of hydrothermally altered mantle along a detachment fault. Lithos, 205, 127–141. https://doi.org/10.1016/j.lithos.2014.06.019
Luth, R. W., & Stachel, T. (2014). The buffering capacity of lithospheric mantle: Implications for diamond formation. Contributions to
Mineralogy and Petrology, 168(5), 1083. https://doi.org/10.1007/s00410-014-1083-6
Maaløe, S. (2005). The dunite bodies, websterite and orthopyroxenite dikes of the Leka ophiolite complex, Norway. Mineralogy and Petrology,
85(3-4), 163–204. https://doi.org/10.1007/s00710-005-0085-5
Mahéo, G., Bertrand, H., Guillot, S., Villa, I. M., Keller, F., & Capiez, P. (2004). The South Ladakh ophiolites (NW Himalaya, India): An intra-oceanic
tholeiitic arc origin with implication for the closure of the Neo-Tethys. Chemical Geology, 203(3-4), 273–303. https://doi.org/10.1016/j.
chemgeo.2003.10.007
Malkovets, V. G., Griffin, W. L., O’Reilly, S. Y., & Wood, B. J. (2007). Diamond, subcalcic garnet, and mantle metasomatism: Kimberlite sampling
patterns define the link. Geology, 35(4), 339–342. https://doi.org/10.1130/G23092A.1
Marchesi, C., Garrido, C. J., Godard, M., Belley, F., & Ferré, E. (2009). Migration and accumulation of ultra-depleted subduction-related melts in
the Massif du Sud ophiolite (New Caledonia). Chemical Geology, 266(3-4), 171–186. https://doi.org/10.1016/j.chemgeo.2009.06.004
Marchesi, C., Garrido, C. J., Godard, M., Proenza, J. A., Gervilla, F., & Blanco-Moreno, J. (2006). Petrogenesis of highly depleted peridotites and
gabbroic rocks from the Mayarí-Baracoa Ophiolitic Belt (eastern Cuba). Contributions to Mineralogy and Petrology, 151(6), 717–736. https://
doi.org/10.1007/s00410-006-0089-0
Miyashiro, A., Shido, F., & Ewing, M. (1969). Composition and origin of serpentinites from the Mid-Atlantic Ridge near 24 and 30 north lati-
tude. Contributions to Mineralogy and Petrology, 23(2), 117–127. https://doi.org/10.1007/BF00375173
Moghadam, H. S., Khedr, M. Z., Arai, S., Stern, R. J., Ghorbani, G., Tamura, A., & Ottley, C. J. (2015). Arc-related harzburgite–dunite–chromitite
complexes in the mantle section of the Sabzevar ophiolite, Iran: A model for formation of podiform chromitites. Gondwana Research,
27(2), 575–593. https://doi.org/10.1016/j.gr.2013.09.007
Nisbet, E. G., Cheadle, M. J., & Bickle, M. J. (1993). Constraining the potential temperature of the Archean mantle: A review of the evidence
from komatiites. Lithos, 30(3-4), 291–307. https://doi.org/10.1016/0024-4937(93)90042-B
Niu, Y. L. (1997). Mantle melting and melt extraction processes beneath ocean ridges: Evidence from abyssal peridotites. Journal of Petrology,
38(8), 1047–1074. https://doi.org/10.1093/petroj/38.8.1047
Niu, Y. L. (2004). Bulk-rock major and trace element compositions of abyssal peridotites: Implications for mantle melting, melt extraction and
post-melting processes beneath mid-ocean ridges. Journal of Petrology, 45(12), 2423–2458. https://doi.org/10.1093/petrology/egh068
Niu, Y. L., & Hékinian, R. (1997). Basaltic liquids and harzburgitic residues in the Garrett Transform: A case study at fast-spreading ridges. Earth
and Planetary Science Letters, 146(1-2), 243–258. https://doi.org/10.1016/S0012-821X(96)00218-X
O’Driscoll, B., Day, J. M., Walker, R. J., Daly, J. S., McDonough, W. F., & Piccoli, P. M. (2012). Chemical heterogeneity in the upper mantle
recorded by peridotites and chromitites from the Shetland Ophiolite Complex, Scotland. Earth and Planetary Science Letters, 333-334,
226–237. https://doi.org/10.1016/j.epsl.2012.03.035
Ohara, Y., & Ishii, T. (1998). Peridotites from the southern Mariana forearc: Heterogeneous fluid supply in mantle wedge. Island Arc, 7(3),
541–558. https://doi.org/10.1111/j.1440-1738.1998.00209.x
Ottonello, G., Joron, J., & Piccardo, G. (1984). Rare earth and 3D transition element geochemistry of peridotitic rocks: II. Ligurian peridotites
and associated basalts. Journal of Petrology, 25(2), 373–393. https://doi.org/10.1093/petrology/25.2.373
Pagé, P., Bédard, J. H., & Tremblay, A. (2009). Geochemical variations in a depleted fore-arc mantle: The Ordovician Thetford Mines ophiolite.
Lithos, 113(1-2), 21–47. https://doi.org/10.1016/j.lithos.2009.03.030
Parkinson, I. J., & Pearce, J. A. (1998). Peridotites from the Izu–Bonin–Mariana forearc (ODP Leg 125): Evidence for mantle melting and melt–
mantle interaction in a supra-subduction zone setting. Journal of Petrology, 39(9), 1577–1618. https://doi.org/10.1093/petroj/39.9.1577
Parman, S. W., Grove, T. L., Dann, J. C., & de Wit, M. J. (2004). A subduction origin for komatiites and cratonic lithospheric mantle. South African
Journal of Geology, 107(1-2) 107–118. https://doi.org/10.2113/107.1-2.107
Paulick, H., Bach, W., Godard, M., De Hoog, J., Suhr, G., & Harvey, J. (2006). Geochemistry of abyssal peridotites (Mid-Atlantic Ridge, 15°200 N,
ODP Leg 209): Implications for fluid/rock interaction in slow spreading environments. Chemical Geology, 234(3-4), 179–210. https://doi.
org/10.1016/j.chemgeo.2006.04.011
Pearce, J. A., Barker, P., Edwards, S., Parkinson, I., & Leat, P. (2000). Geochemistry and tectonic significance of peridotites from the South
Sandwich arc–basin system, South Atlantic. Contributions to Mineralogy and Petrology, 139(1), 36–53. https://doi.org/10.1007/
s004100050572
Pearson, D. G., Carlson, R. W., Shirey, S. B., Boyd, F. R., & Nixon, P. H. (1995). Stabilisation of Archean lithospheric mantle: A Re–Os isotope study
of peridotite xenoliths from the Kaapvaal craton. Earth and Planetary Science Letters, 134(3-4), 341–357. https://doi.org/10.1016/0012-
821X(95)00125-V
Pearson, D. G., & Wittig, N. (2008). Formation of Archean continental lithosphere and its diamonds: The root of the problem. Journal of the
Geological Society, 165(5), 895–914. https://doi.org/10.1144/0016-76492008-003
Pearson, D. G., & Wittig, N. (2014). The formation and evolution of cratonic mantle lithosphere–evidence from mantle xenoliths. In
R. W. Carlson (Ed.), Treatise on geochemistry, (Vol. 3, pp. 255–292). Amsterdam: Elsevier.
Phillips, D., Harris, J., & Viljoen, K. (2004). Mineral chemistry and thermobarometry of inclusions from De Beers Pool diamonds, Kimberley,
South Africa. Lithos, 77(1-4), 155–179. https://doi.org/10.1016/j.lithos.2004.04.005
Powell, R., Holland, T., & Worley, B. (1998). Calculating phase diagrams involving solid solutions via non-linear equations, with examples using
THERMOCALC. Journal of Metamorphic Geology, 16(4), 577–588. https://doi.org/10.1111/j.1525-1314.1998.00157.x
Rajabzadeh, M. A., & Dehkordi, T. N. (2013). Investigation on mantle peridotites from Neyriz ophiolite, south of Iran: Geodynamic signals.
Arabian Journal of Geosciences, 6(11), 4445–4461. https://doi.org/10.1007/s12517-012-0687-2
Richardson, S. H., Gurney, J. J., Erlank, A. J., & Harris, J. W. (1984). Origin of diamonds in old enriched mantle. Nature, 310(5974), 198–202.
https://doi.org/10.1038/310198a0

SU AND CHEN 7704


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Rudnick, R. L., McDonough, W. F., & Chappell, B. W. (1993). Carbonatite metasomatism in the northern Tanzanian mantle: Petrographic and
geochemical characteristics. Earth and Planetary Science Letters, 114(4), 463–475. https://doi.org/10.1016/0012-821X(93)90076-L
Saka, S., Uysal, I., Akmaz, R. M., Kaliwoda, M., & Hochleitner, R. (2014). The effects of partial melting, melt–mantle interaction and fractionation
on ophiolite generation: Constraints from the late cretaceous Pozantı-Karsantı ophiolite, southern Turkey. Lithos, 202-203, 300–316.
https://doi.org/10.1016/j.lithos.2014.05.027
Savov, I. P., Ryan, J. G., D’Antonio, M., & Fryer, P. (2007). Shallow slab fluid release across and along the Mariana arc–basin system: Insights
from geochemistry of serpentinized peridotites from the Mariana fore arc. Journal of Geophysical Research, 112, B09205. https://doi.org/
10.1029/2006JB004749
Savov, I. P., Ryan, J. G., D’Antonio, M., Kelley, K., & Mattie, P. (2005). Geochemistry of serpentinized peridotites from the Mariana Forearc
Conical Seamount, ODP Leg 125: Implications for the elemental recycling at subduction zones. Geochemistry, Geophysics, Geosystems, 6,
Q04J15. https://doi.org/10.1029/2004GC000777
Schulze, D. J. (1995). Low-Ca garnet harzburgites from Kimberley, South Africa: Abundance and bearing on the structure and evolution of the
lithosphere. Journal of Geophysical Research, 100(B7), 12,513–12,526. https://doi.org/10.1029/95JB00029
Shirey, S. B., & Richardson, S. H. (2011). Start of the Wilson cycle at 3 Ga shown by diamonds from subcontinental mantle. Science, 333(6041),
434–436. https://doi.org/10.1126/science.1206275
Simon, N. S. C., Irvine, G. J., Davies, G. R., Pearson, D. G., & Carlson, R. W. (2003). The origin of garnet and clinopyroxene in “depleted” Kaapvaal
peridotites. Lithos, 71(2-4), 289–322. https://doi.org/10.1016/S0024-4937(03)00118-X
Snow, J. E., & Dick, H. J. (1995). Pervasive magnesium loss by marine weathering of peridotite. Geochimica et Cosmochimica Acta, 59(20),
4219–4235. https://doi.org/10.1016/0016-7037(95)00239-V
Sobolev, N. V., Lavrent’ev, Y., Pokhilenko, N. P., & Usova, L. V. (1973). Chrome-rich garnets from the kimberlites of Yakutia and their para-
geneses. Contributions to Mineralogy and Petrology, 40(1), 39–52. https://doi.org/10.1007/BF00371762
Stachel, T., Aulbach, S., Brey, G. P., Harris, J. W., Leost, I., Tappert, R., & Viljoen, K. S. (2004). The trace element composition of silicate inclusions
on diamond: A review. Lithos, 77(1-4), 1–19. https://doi.org/10.1016/j.lithos.2004.03.027
Stachel, T., & Harris, J. W. (2008). The origin of cratonic diamonds—Constraints from mineral inclusions. Ore Geology Reviews, 34(1-2), 5–32.
https://doi.org/10.1016/j.oregeorev.2007.05.002
Stachel, T., & Luth, R. W. (2015). Diamond formation—Where, when and how? Lithos, 220-223, 200–220. https://doi.org/10.1016/j.
lithos.2015.01.028
Stachel, T., Viljoen, K. S., Brey, G., & Harris, J. W. (1998). Metasomatic processes in lherzolitic and harzburgitic domains of diamondiferous
lithospheric mantle: REE in garnets from xenoliths and inclusions in diamonds. Earth and Planetary Science Letters, 159(1-2), 1–12. https://
doi.org/10.1016/S0012-821X(98)00064-8
Stagno, V., Ojwang, D. O., McCammon, C. A., & Frost, D. J. (2013). The oxidation state of the mantle and the extraction of carbon from Earth’s
interior. Nature, 493(7430), 84–88. https://doi.org/10.1038/nature11679
Stephens, C. (1997). Heterogeneity of oceanic peridotite from the western canyon wall at MARK: Results from site 920. Proceeding of the
Ocean Drilling Program, Scientific Results, 153, 285–303.
Tainton, K., & McKenzie, D. (1994). The generation of kimberlites, lamproites, and their sources rocks. Journal of Petrology, 35(3), 787–817.
https://doi.org/10.1093/petrology/35.3.787
Takahashi, E. (1986). Melting of a dry peridotite KLB-1 up to 14 GPa: Implications on the origin of peridotitic upper mantle. Journal of
Geophysical Research, 91(B9), 9367–9382. https://doi.org/10.1029/JB091iB09p09367
Takahashi, E., Shimazaki, T., Tsuzaki, Y., & Yoshida, H. (1993). Melting study of a peridotite KLB-1 to 6.5 GPa, and the origin of basaltic magmas.
Philosophical Transactions of the Royal Society A, 342(1663), 105–120. https://doi.org/10.1098/rsta.1993.0008
Taylor, L. A., Anand, M., Promprated, P., Floss, C., & Sobolev, N. V. (2003). The significance of mineral inclusions in large diamonds from
Yakutia, Russia. American Mineralogist, 88(5-6), 912–920. https://doi.org/10.2138/am-2003-5-621
Uysal, İ., Ersoy, E. Y., Dilek, Y., Escayola, M., Sarıfakıoğlu, E., Saka, S., & Hirata, T. (2015). Depletion and refertilization of the Tethyan oceanic
upper mantle as revealed by the early Jurassic Refahiye ophiolite, NE Anatolia—Turkey. Gondwana Research, 27(2), 594–611. https://doi.
org/10.1016/j.gr.2013.09.008
Uysal, İ., Ersoy, E. Y., Dilek, Y., Kapsiotis, A., & Sarıfakıoğlu, E. (2016). Multiple episodes of partial melting, depletion, metasomatism and
enrichment processes recorded in the heterogeneous upper mantle sequence of the NeotethyanEldivan ophiolite, Turkey. Lithos, 246-
247, 228–245. https://doi.org/10.1016/j.lithos.2016.01.004
Uysal, İ., Ersoy, E. Y., Karslı, O., Dilek, Y., Sadıklar, M. B., Ottley, C. J., et al. (2012). Coexistence of abyssal and ultra-depleted SSZ type mantle
peridotites in a neo-Tethyan ophiolite in SW Turkey: Constraints from mineral composition, whole-rock geochemistry (major–trace–REE–
PGE), and Re–Os isotope systematics. Lithos, 132-133, 50–69. https://doi.org/10.1016/j.lithos.2011.11.009
van Achterbergh, E., Griffin, W., & Stiefenhofer, J. (2001). Metasomatism in mantle xenoliths from the Letlhakane kimberlites: Estimation of
element fluxes. Contributions to Mineralogy and Petrology, 141(4), 397–414. https://doi.org/10.1007/s004100000236
Walter, M. J. (1998). Melting of garnet peridotite and the origin of komatiite and depleted lithosphere. Journal of Petrology, 39(1), 29–60.
https://doi.org/10.1093/petroj/39.1.29
Walter, M. J. (2003). Melt extraction and compositional variability in mantle lithosphere. In R. W. Carlson (Ed.), The mantle and core,
(pp. 363–394). Oxford: Elseriver–Pergamon.
Webb, S. A. C., & Wood, B. J. (1986). Spinel–pyroxene–garnet relationships and their dependence on Cr/Al ratio. Contributions to Mineralogy
and Petrology, 92(4), 471–480. https://doi.org/10.1007/BF00374429
Wittig, N., Pearson, D. G., Webb, M., Ottley, C. J., Irvine, G. J., Kopylova, M., et al. (2008). Origin of cratonic lithospheric mantle roots: A geo-
chemical study of peridotites from the North Atlantic Craton, West Greenland. Earth and Planetary Science Letters, 274(1-2), 24–33. https://
doi.org/10.1016/j.epsl.2008.06.034
Woodland, A. B. (2009). Ferric iron contents of clinopyroxene from cratonic mantle and partitioning behaviour with garnet. Lithos, 112,
1143–1149. https://doi.org/10.1016/j.lithos.2009.04.009
Woodland, A. B., & Koch, M. (2003). Variation in oxygen fugacity with depth in the upper mantle beneath the Kaapvaal craton, southern
Africa. Earth and Planetary Science Letters, 214(1-2), 295–310. https://doi.org/10.1016/S0012-821X(03)00379-0
Yaxley, G. M., Green, D. H., & Kamenetsky, V. (1998). Carbonatite metasomatism in the southeastern Australian lithosphere. Journal of
Petrology, 39(11-12), 1917–1930. https://doi.org/10.1093/petroj/39.11-12.1917
Yumul, G. P. Jr., Zhou, M. F., Wang, C. Y., Zhao, T. P., & Dimalanta, C. B. (2008). Geology and geochemistry of the Shuanggou ophiolite (Ailao
Shan ophiolitic belt), Yunnan Province, SW China: Evidence for a slow-spreading oceanic basin origin. Journal of Asian Earth Sciences,
32(5-6), 385–395. https://doi.org/10.1016/j.jseaes.2007.11.007

SU AND CHEN 7705


21699356, 2018, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018JB016179 by Health Research Board, Wiley Online Library on [15/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2018JB016179

Zhang, P. F., Uysal, I., Zhou, M. F., Su, B. X., & Avcı, E. (2016). Subduction initiation for the formation of high-Cr chromitites in the Kop ophiolite,
NE Turkey. Lithos, 260, 345–355. https://doi.org/10.1016/j.lithos.2016.05.025
Zhou, M. F., Robinson, P., Malpas, J., & Li, Z. (1996). Podiform chromitites in the Luobusa ophiolite (southern Tibet): Implications for melt–rock
interaction and chromite segregation in the upper mantle. Journal of Petrology, 37(1), 3–21. https://doi.org/10.1093/petrology/37.1.3
Ziberna, L., Nimis, P., Zanetti, A., Marzoli, A., & Sobolev, N. V. (2013). Metasomatic processes in the central Siberian cratonic mantle: Evidence
from garnet xenocrysts from the Zagadochnaya kimberlite. Journal of Petrology, 54(11), 2379–2409. https://doi.org/10.1093/petrology/
egt051

SU AND CHEN 7706

You might also like