Next Article in Journal
Crystals from the Powellite-Scheelite Series at the Nanoscale: A Case Study from the Zhibula Cu Skarn, Gangdese Belt, Tibet
Next Article in Special Issue
Columbite U-Pb Geochronology of Kalu’an Lithium Pegmatites in Northern Xinjiang, China: Implications for Genesis and Emplacement History of Rare-Element Pegmatites
Previous Article in Journal
Geology and Genesis of Xianglushan Fe-Cu Orefield in Sichuan (SW China): Evidence from C-O-S-Pb Isotopes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Subduction-Induced Fractionated Highly Siderophile Element Patterns in Forearc Mantle

1
State Key Laboratory of Lithospheric Evolution, Institute of Geology and Geophysics, Chinese Academy of Sciences, Beijing 100029, China
2
Laboratory for Marine Geology, Qingdao National Laboratory for Marine Science and Technology, Qingdao 266061, China
3
University of Chinese Academy of Sciences, Beijing 100049, China
4
CAS Center for Excellence in Tibetan Plateau Earth Sciences, Beijing 100101, China
*
Author to whom correspondence should be addressed.
Minerals 2019, 9(6), 339; https://doi.org/10.3390/min9060339
Submission received: 17 April 2019 / Revised: 16 May 2019 / Accepted: 30 May 2019 / Published: 1 June 2019
(This article belongs to the Special Issue Accessory Mineral Petrogenesis and Isotopic Robustness)

Abstract

:
Compositions of highly siderophile elements (HSEs) in forearc mantle have been little studied and effects of slab dehydration on their abundances in forearc mantle remains unclear. This study reports two different kinds of HSE patterns for peridotites from a New Caledonia forearc ophiolite. The Group-I samples show relatively flat patterns of Ir-group-platinum-group elements (IPGEs) and enrichment of Pt over Pd. Such patterns imply that interstitial sulfides were significantly removed through melt extraction, whereas sulfides enclosed within silicates were mostly unaffected. Meanwhile, Pt-Fe alloys were generated, resulting in suprachondritic Pt/Pd ratios. In contrast, the Group-II samples display convex HSE patterns and are depleted in all HSEs except for Ru, yielding strongly positive Ru anomalies. This indicates that both enclosed and interstitial sulfides were substantially consumed, whereas chromite was generated to stabilize Ru. Compared to abyssal peridotites, subduction-related peridotites commonly have stronger fractionation in the HSEs. Therefore, the HSE data of mantle peridotites are potentially able to discriminate the tectonic settings of ophiolites.

1. Introduction

Highly siderophile elements (HSEs), also known as iron-loving elements, include platinum-group elements (PGEs; Os, Ir, Ru, Rh, Pt, Pd), Re, and Au. Core separation is expected to strip almost all HSEs away from the mantle, due to their extremely high metal/silicate partition coefficients (Dmetal/silicate > 104) [1]. Nevertheless, they are present in the Earth’s mantle in chondritic relative abundances and at higher levels than expected, which has been explained by the late accretion hypothesis [2,3,4]. Such a model has been challenged by the discovery of non-chondritic HSE patterns of mantle samples [5,6,7], which have been explained by variable processes, e.g., addition of non-chondritic extraterrestrial material [8] or differentiated outer-core material to the primitive mantle [7]. Moreover, abundant evidences suggest that non-chondritic HSE patterns can result from partial melting and metasomatism within lithospheric mantle [6,9,10,11].
Subduction zones are likely to be important locations for generating non-chondritic HSE compositions in the subarc mantle (including mantle wedge and forearc mantle), which is evidenced by the highly fractionated HSE patterns of mantle wedge xenoliths, e.g., high Pt/Pd ratios [12,13]. The fractionation might be caused by the mobility of HSEs in slab-released fluids/melts [14], and/or the formation of PGE alloys during hydrous melting [12]. In comparison, whether subduction also produces fractionated HSE patterns of the forearc mantle remains unclear, as HSE compositions of forearc mantle have been rarely reported due to the lack of xenolith samples. This study presents HSE data of mantle peridotites from a New Caledonia ophiolite that originated in forearc settings [15,16,17]. The data support the existence of two different types of fractionated HSE patterns in forearc mantle.

2. Geological Setting and Sample Description

2.1. Geological Setting

The main island of New Caledonia is located in the Southwest Pacific and represents the emerged part of the Norfolk Ridge (Figure 1a), which is considered to be a continental crust fragment rifted from the east margin of Gondwana in the late Cretaceous [18]. The islands of Quvéa, Lifou, and Maré are the original island arc. The main island of New Caledonia has been traditionally divided into four major tectonic units: The continental basement (including “Montagnes Blanches” Nappe, late Cretaceous Paleogene cover, Teremba Terrane, Koh-Central Terrane, and Boghen Terrane), the magmatic Poya terrane, a high-pressure metamorphic belt, and ophiolitic massifs [19].
The New Caledonia ophiolites are exposed in a large massif (the Massif du Sud) occurring in the south of the main island and smaller isolated massifs along the west coast, such as Me Maoya, Kopeto, Koniambo, and Tiebaghi (Figure 1a). They are connected with the oceanic lithosphere of the Loyalty Basin [20,21] and have been regarded as the archetype of forearc ophiolites [22]. The New Caledonia ophiolites are allochthonous and were thrust from NNE to SSW over the continental basement of the main island. The New Caledonia ophiolites consist almost exclusively of mantle rocks (predominantly refractory harzburgites) and crustal rocks are commonly missing. Only a lower crustal cumulate sequence has been observed in the Massif du Sud. It has been suggested that the New Caledonia ophiolites were generated during the Paleocene, and were emplaced onto the main island during the early Eocene [23,24]. Mantle peridotites selected in this study were collected from the Me Maoya klippe, which overlies the mafic Poya terrane (Figure 1b).

2.2. Sample Description

Twenty spinel harzburgites from the Me Maoya massif were selected in this study. They were collected from big boulders in the creek, and no field relation between different samples exists. These harzburgites consist mainly of olivine (67–83% in modal contents), orthopyroxene (16–32%), and spinel (1–2%). Primary clinopyroxene is completely absent in all hazburgites. More than half of the studied samples are fresh, with loss on ignition (LOI) less than 1 wt% (Table 1). In comparison, the remaining eight samples have experienced variable degrees of serpentinization with higher LOI values of 1.36–8.15 wt%. They display porphyroclastic to slightly deformed textures (Figure 2), in which orthopyroxenes commonly occur as porphyroclasts and olivines occur as neoblasts (Figure 2a). Orthopyroxene generally displays irregular forms and shows abundant exsolution lamellae of clinopyroxene (Figure 2a). Olivine neoblasts are relatively small (~0.1–1 mm) and locally display kink bands (Figure 2a). Large spinel grains usually display a lobate texture, whereas small subhedral spinels are interstitial among silicate minerals (Figure 2c,d). Quite small (<0.06 mm) clinopyroxene can be observed to be interstitial along the boundary of olivine, othopyroxene, and spinel (Figure 2c,d). In the serpentinized harzburgites, olivine commonly develops a mesh texture and has been variably replaced by serpentine (Figure 2b). Interstitial sulfide is almost absent in studied samples; only tiny sulfide droplets and ferro-alloys with sizes <15 μm locally occur as inclusions within olivine, orthopyroxene, and spinel (Figure 2e,f). Some sulfides have been partially replaced by Fe hydroxide phases in the serpentinized samples.

3. Analytical Methods

Whole-rock HSE contents were analyzed using isotope dilution. Detailed analytical procedures have been previously described [27]. About 2 g of sample powders were mixed with appropriate amounts of a mixed 187Re-190Os spike and a mixed 191Ir-99Ru-194Pt-105Pd spike into Carius tubes, which were digested by a mixed acid of 3 mL 12 N HCl and 6 mL 16 N HNO3 at ~240 °C for 48–72 h in an oven. Osmium was extracted from the aqua regia solution by solvent extraction into CCl4 and further purified by micro-distillation [28]. Rhenium and remaining PGE (Ir, Ru, Pt, Pd) were first separated from the solution into subgroups (Re–Ru, Ir–Pt, and Pd) by a 2 mL anion exchange resin (AG-1×8, 100–200 mesh). Subsequently, the Re–Ru was further purified by a 0.25 mL anion exchange resin; the Pd and Ir–Pt were further purified by an Eichrom LN spec resin to completely remove Zr and Hf.
Osmium concentrations and isotopes were measured by negative thermal ionization mass spectrometer (N-TIMS) on a GV Isoprobe-T instrument. Details are given in Liu et al. [25]. The concentrations of other HSEs were measured on a Thermo Fisher Scientific Neptune multi-collector ICP-MS (MC-ICPMS) with an electron multiplier in peak-jumping mode or using Faraday cups in static mode, according to the measured signal intensity. In-run precisions for 185Re/187Re, 191Ir/193Ir, 99Ru/101Ru, 194Pt/196Pt, and 105Pd/106Pd were 0.1%–0.3% (2 s). The Re, Ir, Ru, Pt, and Pd standards were used to correct mass fractionations. The correction rate is less than 0.3%. The total procedural blanks were 3 pg for Os, 4 pg for Re, 3 pg for Ir, 9 pg for Ru, 10 pg for Pt, and 6 pg for Pd. Three standards, i.e., WPR-1, UB-N, and BHVO-2, were used to monitor the accuracy of the analytical procedures. The 187Os/188Os ratios for these standards were previously reported [25] and their HSEs contents are given in Table 1. These values are consistent with the published values within errors [29,30,31].

4. Results

The HSE contents are listed in Table 1. The reproducibility of the concentrations, based on three duplicate analyses of samples (i.e., 15NC24, 15NC26, and 15NC31), are more than 10% for some HSEs, indicating that the “nuggest effect” (inhomogeneous distribution of PGE-bearing phases) is significant for the Me Maoya harzburgites. However, the HSE patterns themselves are generally reproducible. Compared to the primitive mantle (PM; Reference [32]), the Me Maoya harzburgites have variably lower contents of HSE (Figure 3), i.e., ~0.002–0.98 times the PM. The studied samples can be divided into two groups according to their HSE patterns. The Group-I harzburgites display relatively flat patterns from Os to Pt and are depleted in Pd, whereas Re is variably enriched relative to Pd (Figure 3a). They have high Pt/Pd ratios, giving maximum (Pt/Pd) N (N: Chondrite-normalized; Reference [33]) ratios of 53.1 (cf. 1.03 of the PM). Compared to the Group-I harzburgites, the Group-II samples are strongly depleted in all HSEs except for Ru (Figure 3b). They display convex patterns, with strong fractionation in IPGEs (Os, Ir, and Ru). They have low (Os/Ir)N ratios of 0.09–0.78, which are much lower than the inferred values (~1.01) of the primitive mantle. They also display strongly positive Ru anomalies (Figure 3b) and have supra-chondritic (Ru/Ir) N ratios of 1.29–17.95 (cf. 1.46 of the PM). Compared to the Group-I samples, the Group-II samples have higher Ru/Os but relatively lower Pt/Pd ratios (Figure 4), i.e., 0.4–7.8 and 5.0–81.5 versus 16.8–71.5 and 0.8–6.0, respectively. There is no significant difference in the Os isotope ratios between the two groups of the Me Maoya harzburgites. The 187Os/188Os ratios do not show any correlation with Ru/Ir or Pt/Pd ratios (Figure 5).

5. Discussion

Both major elements and Os isotope data of the studied samples have been previously reported [25]. They have refractory compositions, with high MgO (42.37–46.53 wt%) but low Al2O3 (0.1–0.78 wt%) and CaO (0.22–0.56 wt%). They have high olivine fosterite contents (Fo) of 0.91–0.93 and spinel Cr# values of 0.55–0.72. The refractory features of the Me Maoya peridotites are also supported by their low HSE contents (i.e., ~0.002–0.98 times the PM). The refractory compositions imply that the Me Maoya peridotites represent residues after high degrees of partial melting and have affinities to forearc peridotites, which is consistent with previous inference that the New Caledonia ophiolite are relics of ocean lithosphere formed in a forearc setting [15,16,17].
Mantle peridotites of the New Caledonia ophiolites have commonly experienced strong serpentinization and supergene weathering, which produced the world-famous nickel laterite ore deposits [34]. The studied Me Maoya harzburgites have also been subjected to variable degrees of alteration, as evidenced by their variable LOI values of 0–8.15 wt%. Therefore, the effects of alteration processes on their HSEs should be evaluated before scrutinizing the HSE data. Serpentinization might have little effects on HSEs of mantle peridotites, as HSEs are unreactive in a reducing environment [7,8,35,36]. In contrast, supergene weathering occurred under an oxidizing environment, and thus, could potentially affect HSEs of the studied samples [7,37,38]. Nevertheless, no correlation between the LOI values and HSE contents (Figure 6) suggests that alteration (neither serpentinization nor supergene weathering) is not the culprit for non-chondritic HSE patterns of the Me Maoya harzburgites.
The Group-I harzburgites have HSE contents lower overall than the values of the PM (Figure 3a). They display relatively flat IPGEs (i.e., Os, Ir, Ru) patterns and strong depletion in Pd (Figure 3a). Such fractionated HSE patterns can be produced by partial melting, as melt extraction can result in depletion of PPGEs (i.e., Rh, Pt, Pd) and Re relative to IPGEs in mantle peridotites [6]. The fractionation has been commonly attributed to incongruent melting of base metal sulfides (BMS) during partial melting, which produces Cu-Ni-rich sulfide melts and leaves refractory monosulfide solid solutions (Mss) in the residues [39]. The refractory Mss concentrates IPGEs, whereas the Cu-Ni-rich sulfides are enriched in PPGEs and Re [9,39]. The Cu-Ni-rich sulfide melts can be physically removed along with magmas, resulting in the depletion of PPGE and Re relative to IPGEs in the residual mantle [40].
Although two samples are depleted in Pt over Pd, most Group-I harzburgites do not show any Pt depletion at all or are even slightly enriched in Pt, with (Pt/Pd) N ratios of 3.27–53.10. Enrichment of Pt over Pd in mantle peridotites cannot be simply explained by incongruent melting of sulfides, as they have similar partition coefficients in sulfides [41]. Thus, other mechanisms should be sought to account for their high Pt/Pd ratios. As shown in Figure 7a, similarly high Pt/Pd ratios have been also previously reported for mantle wedge xenoliths [12,13], which have been explained by formation of Pt–Fe alloys during multi-stage melt depletion [12]. Awaruite, i.e., Pt-rich Fe-Ni alloy (Ni3Fe-Ni2Fe), has been reported for depleted harzburgites from the New Caledonia ophiolites [42], although it has not been discovered in the studied samples. Formation of such Pt-rich alloys during hydrous melting can effectively fractionate Pt from Pd in mantle residues. It should also be noted that Re is variably enriched over Pd in Group-I samples. This might be due to subduction-related metasomatic processes, during which Re was added by the metasomatic agents.
Compared to the Group-I harzburgites, the Group-II samples are extremely depleted in all HSEs except for Ru (Figure 7b). Their HSE display highly fractionated convex patterns, with remarkably positive Ru anomalies, and also show strong fractionation in IPGEs, with sub-chondritic Os/Ir ratios (i.e., 0.09–0.78). Such features indicate that BMS in their protoliths, not only interstitial sulfides, but also sulfide inclusions within silicates, should have been substantially removed during hydrous melting.
Although their sub-chondritic Os/Ir can be explained by sulfide breakdown in the presence of hydrous fluids, as has been proposed for mantle xenoliths [46], removal of sulfides alone is hard to explain their remarkably positive Ru anomalies (i.e., supra-chondritic Ru/Ir ratios). Supra-chondritic Ru/Ir ratios have been reported for peridotites from different tectonic settings and might be an indigenous feature of the Earth’s mantle [8]. It has led to speculation that fractionated meteoritic materials were added as late-veneer after core segregation [8], or a portion of Ru was not stripped to the core and preserved in sulfides within the mantle [47]. However, elevated Ru/Ir ratios >10 are not common for mantle samples, and have been rarely reported for some mantle xenoliths [35]. Therefore, extremely high Ru/Ir ratios up to 26 of the Group-II samples, coupled with their convex HSE patterns, are unlikely indigenous features of Earth’s mantle.
Besides, it has been suggested that strongly supra-chondritic Ru/Ir ratios of mantle peridotites might reflect addition of Ru from the percolating melts [35]. However, the New Caledonia peridotites with Os concentrations of <1 ppb have 187Os/188Os ratios comparable to those of the ambient oceanic mantle (i.e., 187Os/188Os < 0.13). This implies that the HSE-rich phase, e.g., sulfide, in the subducting slab can remain stable during dehydration at shallow depths [25]. Thus, minimal HSEs were transported to the forearc mantle via slab dehydration. The lack of interstitial sulfide in the studied samples also supports that the effect of the percolating melt on Me Maoya peridotites is extremely limited.
A positive Ru anomaly reflects the stability of Ru-bearing phases during melt extraction. Both laurite (RuS2) and Os-Ir-Ru alloys are two major platinum-group metals (PGM) that stabilize Ru in the mantle, both of which have been discovered in mantle peridotites [48,49]. Hydrous melting at high fugacity of oxygen (fO2) in forearc mantle is conducive to PGE alloy nucleation, as it can efficiently reduce the fugacity of sulfur (fS2). During partial melting, Mss in mantle peridotites are totally substituted by laurite, and then by Os–Ir–Ru alloys with a further decrease in fS2, whereas Pt-rich alloys start to form at higher fS2 than do Ru-rich alloys/sulfides [50]. Although both laurite and Os-Ir-Ru alloys can stabilize Ru in the mantle, the enrichment of Ru in Group-II samples requires a phase that does so preferentially over Os, Ir, and the other HSEs. Chromite is, plausibly, such a candidate. Based on a positive relationship between whole-rock Cr and Ru contents, it has been inferred that chromite is an important host mineral of Ru in komatiites [51], which is supported by high Ru contents in chromites obtained by laser ablation [52,53]. Strong compatibility of Ru in chromites has been explained by two different mechanisms. The first possibility is that Ru-bearing minerals (e.g., laurite) occur as micro-scale inclusions within chromites [54]. Alternatively, Ru could substitute into the crystal lattice of chromites at high fO2 and low fS2, as the ionic radii of Ru3+ is theoretically similar to Fe3+ and Cr3+ within the spinel structure [55]. Both mechanisms can stabilize Ru in forearc mantle, whereas other HSEs are scavenged. This explains the strong positive Ru anomalies observed for the Group-II harzburgites.
Two different types of HSE patterns observed for New Caledonia harzburgites indicate that HSE in forearc mantle have been variably fractionated. This implies that the forearc mantle has been heterogeneously affected by hydrous melting (Figure 8). In some portions of the forearc mantle, the interstitial sulfides were significantly removed, whereas sulfides enclosed within silicates and spinels remained mainly untouched. Meanwhile, Pt-rich alloys were produced in these regions during high degrees of hydrous melting. In contrast, some other portions of the forearc mantle might have higher fO2 or contain less sulfur, in which almost all sulfides have been stripped along with melt extraction. In this case, breakdown of sulfides will release all HSEs to the hydrous magmas, whereas Ru was partitioned to chromites.
Combined with HSE compositions of mantle wedge xenoliths [12,13], our data support that subduction can result in strong fractionation in HSEs of the subarc mantle at different depths. Subduction continuously fluxes fluids from the slab to the overlying mantle. The fluids transport fluid-mobile elements (such as Cs, Rb, and Sr) and potential oxidants (e.g., H2O, S, and C), which oxidize the subarc mantle [56]. Sulfides in the subarc mantle are quickly consumed during hydrous melting under oxidizing conditions, as water can increase the sulfur solubility in melts [57]. Hydrous melting should result in a quick decrease of sulfur fugacity in the subarc mantle, driving the formation of PGM [50]. Thus, subduction-modified mantle is more likely to obtain fractionated HSE patterns compared to mid-ocean ridge mantle, where anhydrous melting is expected. Compared to abyssal peridotites, some mantle wedge xenoliths and forearc peridotites have remarkably higher Ru/Ir and Pt/Pd ratios (Figure 9). Thus, HSE compositions of mantle peridotites can be applied to discriminate tectonic settings of ophiolite formation (Figure 9). For example, high Ru/Ir ratios have also been reported for mantle peridotites from the Oman ophiolites [58], consistent with their affinity to supra-subduction zone (SSZ) settings [59]. The HSE compositions of peridotites from worldwide ophiolites are compiled in Figure 9. The Ru/Ir ratios in peridotites from Oman and Scotland ophiolites display increasing tendencies at low Al2O3 contents, which are similar to those in New Caledonia ophiolites and some mantle wedge xenoliths. This implies that Oman ophiolite and Scotland ophiolite may have undergone evolution in the subduction zone setting, which are consistent with previous studies on those ophiolites [58,60].

6. Conclusions

Refractory mantle peridotites from the New Caledonia ophiolites provide a unique opportunity to study the HSE signatures of forearc mantle and evaluate effect of slab dehydration on their abundances. The peridotites from the Me Maoya klippe of the New Caledonia ophiolites show non-chondritic, strongly fractionated HSE patterns. There are no correlations between HSE concentrations and loss on ignition among the samples, which suggests that HSEs are not affected by secondary alteration processes. Based on the characteristics of HSE, the studied peridotites can be classified into two distinct types. The Group-I samples have relatively flat patterns of IPGEs, coupled with supra-chondrite Pt/Pd ratios, resembling the features of some mantle wedge xenoliths. This reflects that interstitial sulfides were substantially consumed but sulfides enclosed were retained during partial melting of peridotites, and Pt-Fe alloys were generated to produce Pt enrichment over Pd. The Group-II samples are characterized by convex HSE patterns with a strong positive Ru anomaly, which reflects that both interstitial and enclosed sulfides were significantly removed, and Ru was partitioned to chromites. The HSE data of the New Caledonia ophiolite peridotites and wedge xenoliths support that subduction-related peridotites commonly display much stronger fractionation in HSE patterns than abyssal peridotites. Therefore, tectonic settings of ophiolites can be effectively discriminated by using the HSE compositions of the constrained mantle peridotites.

Author Contributions

C.-Z.L. designed the project and collected the samples. Y.X. carried out the analysis of HSE compositions. Both authors contributed to the data interpretation and writing.

Funding

This research was financially supported by National Natural Science Foundation of China (41673038), Key Research Program of Frontier Sciences (QYZDB- SSW-DQC032) from Chinese Academy of Sciences, and National Key R&D Program of China (2016YFE0203000).

Acknowledgments

We thank Dominique Cluzel, Fu-Yuan Wu, Wei Lin, Yang Chu, Wen-Bin Ji, and Qing-Ren Meng for help in the field and Zhu-Yin Chu for help in HSE analysis. Comments from two anonymous reviewers are constructive and significantly improved the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Day, J.M.D.; Brandon, A.D.; Walker, R.J. Highly siderophile elements in Earth, Mars, the Moon, and asteroids. Rev. Mineral. Geochem. 2016, 81, 161–238. [Google Scholar] [CrossRef] [PubMed]
  2. Chou, C.L. Fractionation of siderophile elements in the Earth’s upper mantle. Proc. Lunar Planet. Sci. 1978, 9, 219–230. [Google Scholar]
  3. Holzheid, A.; Sylvester, P.; O’Neill, H.S.C.; Rubie, D.C.; Palme, H. Evidence for a late chondritic veneer in the Earth’s mantle from high-pressure partitioning of palladium and platinum. Nature 2000, 406, 396. [Google Scholar] [CrossRef] [PubMed]
  4. Bottke, W.F.; Walker, R.J.; Day, J.M.D.; Nesvorny, D.; Elkins-Tanton, L. Stochastic late accretion to Earth, the Moon, and Mars. Science 2010, 330, 1527–1530. [Google Scholar] [CrossRef]
  5. Pattou, L.; Lorand, J.P.; Gros, M. Non-chondritic platinum-group element ratios in the Earth’s mantle. Nature 1996, 379, 712–715. [Google Scholar] [CrossRef]
  6. Rehkämper, M.; Halliday, A.N.; Alt, J.; Fitton, J.G.; Zipfel, J.; Takazawa, E. Non-chondritic platinum-group element ratios in oceanic mantle lithosphere: Petrogenetic signature of melt percolation? Earth Planet. Sci. Lett. 1999, 172, 65–81. [Google Scholar] [CrossRef]
  7. Snow, J.E.; Schmidt, G. Constraints on Earth accretion deduced from noble metals in the oceanic mantle. Nature 1998, 391, 166–169. [Google Scholar] [CrossRef]
  8. Becker, H.; Horan, M.F.; Walker, R.J.; Gao, S.; Lorand, J.P.; Rudnick, R.L. Highly siderophile element composition of the Earth’s primitive upper mantle: Constraints from new data on peridotite massifs and xenoliths. Geochim. Cosmochim. Acta 2006, 70, 4528–4550. [Google Scholar] [CrossRef]
  9. Alard, O.; Griffin, W.L.; Lorand, J.P.; Jackson, S.E. Non-chondritic distribution of the highly siderophile elements in mantle sulphides. Nature 2000, 407, 891–894. [Google Scholar] [CrossRef] [PubMed]
  10. Lorand, J.P.; Luguet, A.; Alard, O. Platinum group elements: A new set of key tracers for the Earth’s interior. Elements 2009, 4, 247. [Google Scholar] [CrossRef]
  11. Harvey, J.; König, S.; Luguet, A. The effects of melt depletion and metasomatism on highly siderophile and strongly chalcophile elements: S–Se–Te–Re–PGE systematics of peridotite xenoliths from Kilbourne Hole, New Mexico. Geochim. Cosmochim. Acta 2015, 166, 210–233. [Google Scholar] [CrossRef]
  12. Kepezhinskas, P.; Defant, M.J. Nonchondritic Pt/Pd ratios in arc mantle xenoliths: Evidence for platinum enrichment in depleted island-arc mantle sources. Geology 2001, 29, 851–854. [Google Scholar] [CrossRef]
  13. McInnes, B.I.; McBride, J.S.; Evans, N.J.; Lambert, D.D.; Andrew, A.S. Osmium isotope constraints on ore metal recycling in subduction zones. Science 1999, 286, 512–516. [Google Scholar] [CrossRef] [PubMed]
  14. Brandon, A.D.; Creaser, R.A.; Shirey, S.B.; Carlson, R.W. Osmium recycling in subduction zones. Science 1996, 272, 861–864. [Google Scholar] [CrossRef]
  15. Pirard, C.; Hermann, J.; O’Neill, H.S.C. Petrology and geochemistry of the crust-mantle boundary in a nascent arc, Massif du Sud Ophiolite, New Caledonia, SW Pacific. J. Petrol. 2013, 54, 1759–1792. [Google Scholar] [CrossRef]
  16. Marchesi, C.; Garrido, C.J.; Godard, M.; Belley, F.; Ferré, E. Migration and accumulation of ultra-depleted subduction-related melts in the Massif du Sud ophiolite (New Caledonia). J. Petrol. 2009, 266, 171–186. [Google Scholar] [CrossRef]
  17. Whattam, S.A. Arc-continent collisional orogenesis in the SW Pacific and the nature, source and correlation of emplaced ophiolitic nappe components. Lithos 2009, 113, 88–114. [Google Scholar] [CrossRef]
  18. Sdrolias, M.; Müller, R.D.; Gaina, C. Tectonic evolution of the southwest Pacific using constraints from backarc basins. In Evolution and Dynamics of the Australian Plate; Hillis, R.R., Müller, R.D., Eds.; Geological Society of America: Boulder, CO, USA, 2003; pp. 343–359. [Google Scholar]
  19. Cluzel, D.; Aitchison, J.C.; Picard, C. Tectonic accretion and underplating of mafic terranes in the Late Eocene intraoceanic fore-arc of New Caledonia (Southwest Pacific): Geodynamic implications. Tectonophysics 2001, 340, 23–59. [Google Scholar] [CrossRef]
  20. Collot, J.Y.; Malahoff, A.; Recy, J.; Latham, G.; Missegue, F. Overthrust Emplacement of New Caledonia Ophiolite: Geophysical Evidence. Tectonics 1987, 6, 215–232. [Google Scholar] [CrossRef]
  21. Patriat, M.; Collot, J.; Etienne, S.; Poli, S.; Clerc, C.; Mortimer, N.; Pattier, F.; Juan, C.; Roest, W.R. New Caledonia obducted Peridotite Nappe: offshore extent and implications for obduction and postobduction processes. Tectonics 2018, 37, 1077–1096. [Google Scholar] [CrossRef]
  22. Milsom, J. Forearc ophiolites: A view from the western Pacific. In Ophiolites in Earth History; Dilek, Y., Robinson, P.T., Eds.; Geological Society of London Special Publication: London, UK, 2003; Volume 218, pp. 507–515. [Google Scholar]
  23. Cluzel, D.; Jourdan, F.; Meffre, S.; Maurizot, P.; Lesimple, S. The metamorphic sole of New Caledonia ophiolite:40Ar/39Ar, U-Pb, and geochemical evidence for subduction inception at a spreading ridge. Tectonics 2012, 31. [Google Scholar] [CrossRef]
  24. Cluzel, D.; Meffre, S.; Maurizot, P.; Crawford, A.J. Earliest Eocene (53Ma) convergence in the Southwest Pacific: evidence from pre-obduction dikes in the ophiolite of New Caledonia. Terra Nova 2006, 18, 395–402. [Google Scholar] [CrossRef]
  25. Liu, C.Z.; Xu, Y.; Wu, F.Y. Limited recycling of crustal osmium in forearc mantle during slab dehydration. Geology 2018, 43, 239–242. [Google Scholar] [CrossRef]
  26. Cluzel, D.; Maurizot, P.; Collot, J.; Sevin, B.J.E. An outline of the Geology of New Caledonia; from Permian-Mesozoic Southeast-Gondwanaland active margin to Tertiary obduction and supergene evolution. Epis. -Newsmag. Int. Geol. Sci. 2012, 35, 72–86. [Google Scholar]
  27. Chu, Z.Y.; Wu, F.Y.; Walker, R.J.; Rudnick, R.L.; Pitcher, L.; Puchtel, I.S.; Yang, Y.H.; Wilde, S.A. Temporal Evolution of the Lithospheric Mantle beneath the Eastern North China Craton. J. Petrol. 2009, 50, 1857–1898. [Google Scholar] [CrossRef] [Green Version]
  28. Birck, J.L.; Roy-Barman, M.; Capmas, F. Re-Os isotopic measurements at the femtomole level in natural samples. Geostand. Geoanal. Res. 1997, 21, 19–27. [Google Scholar] [CrossRef]
  29. Day, J.M.D.; Waters, C.L.; Schaefer, B.F.; Walker, R.J.; Turner, S. Use of Hydrofluoric Acid Desilicification in the Determination of Highly Siderophile Element Abundances and Re-Pt-Os Isotope Systematics in Mafic-Ultramafic Rocks. Geostand. Geoanal. Res. 2016, 40, 49–65. [Google Scholar] [CrossRef]
  30. Ishikawa, A.; Senda, R.; Suzuki, K.; Dale, C.W.; Meisel, T. Re-evaluating digestion methods for highly siderophile element and 187Os isotope analysis: Evidence from geological reference materials. Chem. Geol. 2014, 384, 27–46. [Google Scholar] [CrossRef]
  31. Li, J.; Zhao, P.P.; Liu, J.G.; Wang, X.C.; Yang, A.Y.; Wang, G.Q.; Xu, J.F. Reassessment of Hydrofluoric acid desilicification in the Carius tube digestion technique for Re–Os isotopic determination in geological samples. Geostand. Geoanal. Res. 2015, 39, 17–30. [Google Scholar] [CrossRef]
  32. Day, J.M.D.; Walker, R.J.; Warren, J.M. 186Os–187Os and highly siderophile element abundance systematics of the mantle revealed by abyssal peridotites and Os-rich alloys. Geochim. Cosmochim. Acta 2017, 200, 232–254. [Google Scholar] [CrossRef]
  33. Fischer-Gödde, M.; Becker, H.; Wombacher, F. Rhodium, gold and other highly siderophile element abundances in chondritic meteorites. Geochim. Cosmochim. Acta 2010, 74, 356–379. [Google Scholar] [CrossRef]
  34. Butt, C.W.; Cluzel, D. Nickel Laterite Ore Deposits: Weathered Serpentinites. Elements 2013, 9, 123–128. [Google Scholar] [CrossRef]
  35. Fischer-Gödde, M.; Becker, H.; Wombacher, F. Rhodium, gold and other highly siderophile elements in orogenic peridotites and peridotite xenoliths. Chem. Geol. 2011, 280, 365–383. [Google Scholar] [CrossRef]
  36. Foustoukos, D.I.; Bizimis, M.; Frisby, C.; Shirey, S.B. Redox controls on Ni–Fe–PGE mineralization and Re/Os fractionation during serpentinization of abyssal peridotite. Geochim. Cosmochim. Acta 2015, 150, 11–25. [Google Scholar] [CrossRef]
  37. Luguet, A.; Lorand, J.P.; Seyler, M. Sulfide petrology and highly siderophile element geochemistry of abyssal peridotites: a coupled study of samples from the Kane Fracture Zone (45°W 23°20N, MARK area, Atlantic Ocean). Geochim. Cosmochim. Acta 2003, 67, 1553–1570. [Google Scholar] [CrossRef]
  38. Pearson, D.G.; Irvine, G.J.; Ionov, D.A.; Boyd, F.R.; Dreibus, G.E. Re–Os isotope systematics and platinum group element fractionation during mantle melt extraction: a study of massif and xenolith peridotite suites. Chem. Geol. 2004, 208, 29–59. [Google Scholar] [CrossRef]
  39. Bockrath, C.; Ballhaus, C.; Holzheid, A. Fractionation of the platinum-group elements during mantle melting. Science 2004, 305, 1951–1953. [Google Scholar] [CrossRef] [PubMed]
  40. Ballhaus, C.; Bockrath, C.; Wohlgemuth-Ueberwasser, C.; Laurenz, V.; Berndt, J. Fractionation of the noble metals by physical processes. Contrib. Mineral. Petrol. 2006, 152, 667–684. [Google Scholar] [CrossRef]
  41. Li, C.; Barnes, S.-J.; Makovicky, E.; Rose-Hansen, J.; Makovicky, M. Partitioning of nickel, copper, iridium, rhenium, platinum, and palladium between monosulfide solid solution and sulfide liquid: Effects of composition and temperature. Geochim. Cosmochim. Acta 1996, 60, 1231–1238. [Google Scholar] [CrossRef]
  42. Augé, T.Y.; Cabri, L.J.; Legendre, O.; McMahon, G.; Cocherie, A. PGE distribution in base-metal alloys and sulfides of the New Caledonia Ophiolite. Can. Mineral. 1999, 37, 1147–1161. [Google Scholar]
  43. Brandon, A.D.; Snow, J.E.; Walker, R.J.; Morgan, J.W.; Mock, T.D. 190Pt–186Os and 187Re–187Os systematics of abyssal peridotites. Earth Planet. Sci. Lett. 2000, 177, 319–335. [Google Scholar] [CrossRef]
  44. Liu, C.Z.; Snow, J.E.; Brügmann, G.; Hellebrand, E.; Hofmann, A.W. Non-chondritic HSE budget in Earth’s upper mantle evidenced by abyssal peridotites from Gakkel ridge (Arctic Ocean). Earth Planet. Sci. Lett. 2009, 283, 122–132. [Google Scholar] [CrossRef]
  45. Marchesi, C.; Garrido, C.J.; Harvey, J.; González-Jiménez, J.M.; Hidas, K.; Lorand, J.-P.; Gervilla, F. Platinum-group elements, S, Se and Cu in highly depleted abyssal peridotites from the Mid-Atlantic Ocean Ridge (ODP Hole 1274A): Influence of hydrothermal and magmatic processes. Contrib. Mineral. Petrol. 2013, 166, 1521–1538. [Google Scholar] [CrossRef] [Green Version]
  46. Handler, M.R.; Bennett, V.C.; Dreibus, G. Evidence from correlated Ir/Os and Cu/S for late-stage Os mobility in peridotite xenoliths: Implications for Re-Os systematics. Geology 1999, 27, 75–78. [Google Scholar] [CrossRef]
  47. Rubie, D.C.; Laurenz, V.; Jacobson, S.A.; Morbidelli, A.; Palme, H.; Vogel, A.K.; Frost, D.J. Highly siderophile elements were stripped from Earth’s mantle by iron sulfide segregation. Science 2016, 353, 1141–1144. [Google Scholar] [CrossRef] [PubMed]
  48. Luguet, A.; Shirey, S.B.; Lorand, J.-P.; Horan, M.F.; Carlson, R.W. Residual platinum-group minerals from highly depleted harzburgites of the Lherz massif (France) and their role in HSE fractionation of the mantle. Geochim. Cosmochim. Acta 2007, 71, 3082–3097. [Google Scholar] [CrossRef]
  49. O’Driscoll, B.; Garwood, R.; Day, J.M.D.; Wogelius, R. Platinum-group element remobilization and concentration in the Cliff chromitites of the Shetland Ophiolite Complex, Scotland. Mineral. Mag. 2018, 82, 471–490. [Google Scholar] [CrossRef] [Green Version]
  50. Lorand, J.P.; Luguet, A. Chalcophile and Siderophile Elements in Mantle Rocks: Trace Elements Controlled By Trace Minerals. Rev. Mineral. Geochem. 2016, 81, 441–488. [Google Scholar] [CrossRef]
  51. Fiorentini, M.L.; Beresford, S.W.; Barley, M.E. Ruthenium-chromium variation: A new ltihogeochemical tool in the exploration for komatiite-hosted Ni-Cu-(PGE) deposit. Econ. Geol. 2008, 103, 431–437. [Google Scholar] [CrossRef]
  52. Locmelis, M.; Pearson, N.L.; Barnes, S.J.; Fiorentini, M.L. Ruthenium in komatiitic chromite. Geochim. Cosmochim. Acta 2011, 75, 3645–3661. [Google Scholar] [CrossRef]
  53. O’Driscoll, B.; Walker, R.J.; Day, J.M.D.; Ash, R.D.; Daly, J.S. Generations of melt extraction, melt–rock interaction and high-temperature metasomatism preserved in peridotites of the∼ 497 Ma Leka Ophiolite Complex, Norway. J. Petrol. 2015, 56, 1797–1828. [Google Scholar] [CrossRef]
  54. Pagé, P.; Barnes, S.J.; Bédard, J.H.; Zientek, M.L. In situ determination of Os, Ir, and Ru in chromites formed from komatiite, tholeiite and boninite magmas: Implications for chromite control of Os, Ir and Ru during partial melting and crystal fractionation. Chem. Geol. 2012, 302–303, 3–15. [Google Scholar] [CrossRef]
  55. Capobianco, C.J.; Drake, M.J. Partitoning of ruthenium, rhodium, and palladium between spinel and silicate melt and implications for platinum group element fractionation trends. Geochim. Cosmochim. Acta 1990, 54, 869–874. [Google Scholar] [CrossRef]
  56. Brounce, M.; Kelley, K.A.; Cottrell, E.; Reagan, M.K. Temporal evolution of mantle wedge oxygen fugacity during subduction initiation. Geology 2015, 43, 775–778. [Google Scholar] [CrossRef]
  57. Fortin, M.A.; Riddle, J.; Desjardins-Langlais, Y.; Baker, D.R. The effect of water on the sulfur concentration at sulfide saturation (SCSS) in natural melts. Geochim. Cosmochim. Acta 2015, 160, 100–116. [Google Scholar] [CrossRef]
  58. Hanghøj, K.; Kelemen, P.B.; Hassler, D.; Godard, M. Composition and Genesis of Depleted Mantle Peridotites from the Wadi Tayin Massif, Oman Ophiolite; Major and Trace Element Geochemistry, and Os Isotope and PGE Systematics. J. Petrol. 2010, 51, 201–227. [Google Scholar] [CrossRef]
  59. Macleod, C.J.; Lissenberg, C.J.; Bibby, L.E. “Moist MORB” axial magmatism in the Oman ophiolite: The evidence against a mid-ocean ridge origin. Geology 2013, 41, 459–462. [Google Scholar] [CrossRef]
  60. O’Driscoll, B.; Day, J.M.D.; Walker, R.J.; Daly, J.S.; Mcdonough, W.F.; Piccoli, P.M. Chemical heterogeneity in the upper mantle recorded by peridotites and chromitites from the Shetland Ophiolite Complex, Scotland. Earth Planet. Sci. Lett. 2012, 333–334, 226–237. [Google Scholar] [CrossRef]
  61. O’Driscoll, B.; Walker, R.J.; Clay, P.L.; Day, J.M.D.; Ash, R.D.; Daly, J.S. Length-scales of chemical and isotopic heterogeneity in the mantle section of the Shetland Ophiolite Complex, Scotland. Earth Planet. Sci. Lett. 2018, 488, 144–154. [Google Scholar] [CrossRef] [Green Version]
  62. Aldanmaz, E.; Koprubasi, N. Platinum-Group-Element Systematics of Peridotites from Ophiolite Complexes of Northwest Anatolia, Turkey: Implications for Mantle Metasomatism by Melt Percolation in a Supra-subduction Zone Environment. Int. Geol. Rev. 2006, 48, 420–442. [Google Scholar] [CrossRef]
  63. Batanova, V.G.; Brügmann, G.E.; Bazylev, B.A.; Sobolev, A.V.; Kamenetsky, V.S.; Hofmann, A.W. Platinum-group element abundances and Os isotope composition of mantle peridotites from the Mamonia complex, Cyprus. Chem. Geol. 2008, 248, 195–212. [Google Scholar] [CrossRef]
  64. Büchl, A.; Brügmann, G.; Batanova, V.G.; Münker, C.; Hofmann, A.W. Melt percolation monitored by Os isotopes and HSE abundances: A case study from the mantle section of the Troodos Ophiolite. Earth Planet. Sci. Lett. 2002, 204, 385–402. [Google Scholar] [CrossRef]
  65. Liu, C.Z.; Wu, F.Y.; Wilde, S.A.; Yu, L.J.; Li, J.L. Anorthitic plagioclase and pargasitic amphibole in mantle peridotites from the Yungbwa ophiolite (southwestern Tibetan Plateau) formed by hydrous melt metasomatism. Lithos 2010, 114, 413–422. [Google Scholar] [CrossRef]
  66. Liu, T.; Wu, F.Y.; Liu, C.Z.; Zhang, C.; Ji, W.B.; Xu, Y. Reconsideration of Neo-Tethys evolution constrained from the nature of the Dazhuqu ophiolitic mantle, southern Tibet. Contrib. Mineral. Petrol. 2019, 174. [Google Scholar] [CrossRef]
  67. Snow, J.E.; Schmidt, G.; Rampone, E. Os isotopes and highly siderophile elements (HSE) in the Ligurian ophiolites, Italy. Earth Planet. Sci. Lett. 2000, 175, 119–132. [Google Scholar] [CrossRef]
  68. Uysal, I.; Ersoy, E.Y.; Karslı, O.; Dilek, Y.; Sadıklar, M.B.; Ottley, C.J.; Tiepolo, M.; Meisel, T. 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 2012, 132–133, 50–69. [Google Scholar] [CrossRef]
  69. Zhou, M.; Robinson, P.T.; Malpas, J.; Edwards, S.J.; Qi, L. REE and PGE Geochemical Constraints on the Formation of Dunites in the Luobusa Ophiolite, Southern Tibet. J. Petrol. 2005, 46, 615–639. [Google Scholar] [CrossRef]
  70. Schulte, R.F.; Schilling, M.; Anma, R.; Farquhar, J.; Horan, M.F.; Komiya, T.; Piccoli, P.M.; Pitcher, L.; Walker, R.J. Chemical and chronologic complexity in the convecting upper mantle: Evidence from the Taitao ophiolite, southern Chile. Geochim. Cosmochim. Acta 2009, 73, 5793–5819. [Google Scholar] [CrossRef]
Figure 1. (a) Sketch geological map of New Caledonia, modified from Reference [25]; (b) cross section of New Caledonia, modified from Reference [26].
Figure 1. (a) Sketch geological map of New Caledonia, modified from Reference [25]; (b) cross section of New Caledonia, modified from Reference [26].
Minerals 09 00339 g001
Figure 2. Microstructure of the Me Maoya harzburgites. (a) Harzburgite with a porphyroclastic texture, orthopyroxene porphyroclasts, and olivine porphyroclasts are surrounded by olivine neoblasts; note kink bands in some olivine (15NC21); (b) olivine partially altered to serpentine along cracks (15NC12); (c) large spinel grain with a lobate shape in a olivine matrix (15NC25); (d) orthopyroxene with clinopyroxene exsolution lamellae, small spinel, and clinopyroxene, disseminated in the olivine matrix (15NC23); (e) tiny sulfide enclosed within olivine (15NC22); (f) rounded sulfide bleb enclosed in spinel (15NC26). Ol: Olivine; Opx: Orthopyroxene; Cpx: Clinopyroxene; Sp: Spinel; Chr: Chromite; Sulf: Sulfide.
Figure 2. Microstructure of the Me Maoya harzburgites. (a) Harzburgite with a porphyroclastic texture, orthopyroxene porphyroclasts, and olivine porphyroclasts are surrounded by olivine neoblasts; note kink bands in some olivine (15NC21); (b) olivine partially altered to serpentine along cracks (15NC12); (c) large spinel grain with a lobate shape in a olivine matrix (15NC25); (d) orthopyroxene with clinopyroxene exsolution lamellae, small spinel, and clinopyroxene, disseminated in the olivine matrix (15NC23); (e) tiny sulfide enclosed within olivine (15NC22); (f) rounded sulfide bleb enclosed in spinel (15NC26). Ol: Olivine; Opx: Orthopyroxene; Cpx: Clinopyroxene; Sp: Spinel; Chr: Chromite; Sulf: Sulfide.
Minerals 09 00339 g002
Figure 3. HSE patterns of Me Maoya harzburgites. (a) Group-I harzburgites; (b) Group-II harzburgites. Data of the primitive mantle (PM) are from Reference [32]. The data were chondrite-normalized using values from Reference [33].
Figure 3. HSE patterns of Me Maoya harzburgites. (a) Group-I harzburgites; (b) Group-II harzburgites. Data of the primitive mantle (PM) are from Reference [32]. The data were chondrite-normalized using values from Reference [33].
Minerals 09 00339 g003
Figure 4. Ru/Os versus Pt/Pd of Me Maoya harzburgites. Increasing of Pt/Pd ratios result from formation of Pt-bearing phases along with hydrous melting.
Figure 4. Ru/Os versus Pt/Pd of Me Maoya harzburgites. Increasing of Pt/Pd ratios result from formation of Pt-bearing phases along with hydrous melting.
Minerals 09 00339 g004
Figure 5. 187Os/188Os versus Pt/Pd (a) and Ru/Ir (b) in Me Maoya harzburgites.
Figure 5. 187Os/188Os versus Pt/Pd (a) and Ru/Ir (b) in Me Maoya harzburgites.
Minerals 09 00339 g005
Figure 6. Loss on ignition (LOI) values versus Os (a), Ir (b), Ru (c), Pt (d), Pd (e) and Re (f) contents.
Figure 6. Loss on ignition (LOI) values versus Os (a), Ir (b), Ru (c), Pt (d), Pd (e) and Re (f) contents.
Minerals 09 00339 g006
Figure 7. Pt/Pd (a) and Ru/Ir (b) versus bulk Al2O3 contents in the Me Maoya harzburgites. Fields of mantle wedge xenoliths [12,13] and abyssal peridotites [32,37,43,44,45] are also shown.
Figure 7. Pt/Pd (a) and Ru/Ir (b) versus bulk Al2O3 contents in the Me Maoya harzburgites. Fields of mantle wedge xenoliths [12,13] and abyssal peridotites [32,37,43,44,45] are also shown.
Minerals 09 00339 g007
Figure 8. Cartoon for the genesis of two different HSE patterns in forearc mantle. The forearc mantle is both physically and chemically heterogeneous. For the Group-I patterns, interstitial sulfides in the mantle were significantly removed, whereas the sulfides included within silicates remained mainly untouched. Pt-bearing alloys were also generated to produce Pt enrichment over Pd. The Group-II patterns were produced in forearc mantle in which almost all sulfides were stripped. Breakdown of sulfides released all HSEs to the hydrous magmas, and Ru was trapped within the chromite.
Figure 8. Cartoon for the genesis of two different HSE patterns in forearc mantle. The forearc mantle is both physically and chemically heterogeneous. For the Group-I patterns, interstitial sulfides in the mantle were significantly removed, whereas the sulfides included within silicates remained mainly untouched. Pt-bearing alloys were also generated to produce Pt enrichment over Pd. The Group-II patterns were produced in forearc mantle in which almost all sulfides were stripped. Breakdown of sulfides released all HSEs to the hydrous magmas, and Ru was trapped within the chromite.
Minerals 09 00339 g008
Figure 9. Plots of Pt/Pd (a) and Ru/Ir (b) versus Al2O3 contents. Data of the Oman ophiolites are from [58], the Shetland ophiolites from [60,61], other peridotites are from [53,62,63,64,65,66,67,68,69,70]. Fields of mantle wedge xenoliths [12,13] and abyssal peridotites are shown for comparison [32,37,43,44,45].
Figure 9. Plots of Pt/Pd (a) and Ru/Ir (b) versus Al2O3 contents. Data of the Oman ophiolites are from [58], the Shetland ophiolites from [60,61], other peridotites are from [53,62,63,64,65,66,67,68,69,70]. Fields of mantle wedge xenoliths [12,13] and abyssal peridotites are shown for comparison [32,37,43,44,45].
Minerals 09 00339 g009
Table 1. Highly siderophile elements (HSE) data (in ppb) of the Me Maoya harzburgites.
Table 1. Highly siderophile elements (HSE) data (in ppb) of the Me Maoya harzburgites.
Al2O3MgOCaOLOIOl FoSp Cr#OsIrRuPtPdRe(Os/Ir)N(Ru/Ir)N(Ru/Os)N(Pt/Pd)N
Standards
WPR-1 16.916.723.730824011.2
UB-N 3.543.206.336.955.830.22
BHVO-2 0.080.070.115.522.730.57
Group-I harzburgites
15NC110.3943.410.366.170.920.551.760.621.097.460.600.0112.591.180.468.10
15NC120.3242.370.408.150.910.551.090.750.463.010.600.0291.320.410.313.27
15NC170.1945.530.450.640.910.683.452.324.334.070.050.0131.361.250.9253.03
15NC210.4345.680.480.740.920.560.390.321.330.560.060.0121.112.792.516.08
15NC230.3246.530.400.300.920.560.300.490.761.470.270.0190.561.041.863.55
15NC250.7445.090.510.240.930.571.621.232.783.190.090.1971.201.521.2623.09
15NC270.4743.400.563.580.910.551.070.922.403.230.160.0061.061.751.6513.15
15NC280.2845.090.421.360.910.661.730.791.874.900.130.0162.001.590.8024.56
15NC290.3645.290.560.280.910.590.840.580.723.290.120.0011.320.830.6317.86
15NC300.7844.450.500.100.920.590.240.431.863.260.040.0020.512.905.7053.10
15NC310.1045.240.222.460.910.720.700.522.061.330.040.0101.232.662.1721.66
15NC31-R 0.460.662.681.200.150.0200.642.724.295.21
15NC330.5044.840.420.820.920.560.530.743.400.280.040.0260.653.084.724.56
15NC340.3745.420.49−0.020.910.570.790.673.644.550.230.0081.073.643.3912.89
Group-II harzburgites
15NC130.4143.330.425.960.920.590.020.101.430.120.100.0050.189.5952.630.78
15NC150.2443.720.366.180.920.580.050.061.210.180.030.0020.7613.5317.813.91
15NC200.4245.990.400.040.920.600.020.200.720.080.060.0010.092.4226.500.87
15NC220.4145.500.37−0.060.920.630.030.280.540.030.040.0020.101.2913.250.49
15NC240.2945.750.360.060.920.640.030.110.540.040.010.0010.253.2913.252.61
15NC24-R 0.010.100.600.040.040.0020.094.0344.170.65
15NC260.3145.460.321.880.920.590.060.071.010.060.060.0010.789.6812.390.65
15NC26-R 0.020.041.070.050.060.0020.4617.9539.380.54
15NC320.3345.870.340.020.920.620.020.040.920.070.030.0030.4615.4333.861.52
1: Major element contents are in wt%; 2: HSE contents are in ppb; 3: Chondrite normalized values are from Fischer-Godde et al. [33]; 4: The data of major elements and the contents of Re and Os in the studied samples are from Liu et al. [25].

Share and Cite

MDPI and ACS Style

Xu, Y.; Liu, C.-Z. Subduction-Induced Fractionated Highly Siderophile Element Patterns in Forearc Mantle. Minerals 2019, 9, 339. https://doi.org/10.3390/min9060339

AMA Style

Xu Y, Liu C-Z. Subduction-Induced Fractionated Highly Siderophile Element Patterns in Forearc Mantle. Minerals. 2019; 9(6):339. https://doi.org/10.3390/min9060339

Chicago/Turabian Style

Xu, Yang, and Chuan-Zhou Liu. 2019. "Subduction-Induced Fractionated Highly Siderophile Element Patterns in Forearc Mantle" Minerals 9, no. 6: 339. https://doi.org/10.3390/min9060339

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop