Letter

CaSiO3 perovskite in diamond indicates the recycling of oceanic crust into the lower mantle

  • Nature volume 555, pages 237241 (08 March 2018)
  • doi:10.1038/nature25972
  • Download Citation
Received:
Accepted:
Published:

Abstract

Laboratory experiments and seismology data have created a clear theoretical picture of the most abundant minerals that comprise the deeper parts of the Earth’s mantle. Discoveries of some of these minerals in ‘super-deep’ diamonds—formed between two hundred and about one thousand kilometres into the lower mantle—have confirmed part of this picture1,2,3,4,5. A notable exception is the high-pressure perovskite-structured polymorph of calcium silicate (CaSiO3). This mineral—expected to be the fourth most abundant in the Earth—has not previously been found in nature. Being the dominant host for calcium and, owing to its accommodating crystal structure, the major sink for heat-producing elements (potassium, uranium and thorium) in the transition zone and lower mantle, it is critical to establish its presence. Here we report the discovery of the perovskite-structured polymorph of CaSiO3 in a diamond from South African Cullinan kimberlite. The mineral is intergrown with about six per cent calcium titanate (CaTiO3). The titanium-rich composition of this inclusion indicates a bulk composition consistent with derivation from basaltic oceanic crust subducted to pressures equivalent to those present at the depths of the uppermost lower mantle. The relatively ‘heavy’ carbon isotopic composition of the surrounding diamond, together with the pristine high-pressure CaSiO3 structure, provides evidence for the recycling of oceanic crust and surficial carbon to lower-mantle depths.

Main

A key goal of solid-Earth geosciences is to establish the mineralogy of the Earth’s mantle throughout its depth, which acts as a primary control on mantle dynamics and chemistry. Diamonds are unique in this regard because they provide access to the deepest intact material from the Earth’s interior through the minerals contained within their volumes. Over the past three decades, a growing number of studies have used a class of diamonds known as super-deep diamonds to study mantle processes in the deep sublithospheric mantle, the transition zone and the lower mantle1,2,3,4,5,6. Early studies1,2,3 suggest that some of the assemblages included within super-deep diamonds represent samples of the lower mantle and the transition zone that variably retrogressed to lower pressures. Later studies indicate that some of these assemblages and minerals might originate from shallower depths7,8, although still beneath the lithosphere.

The most common minerals found within super-deep diamonds are ferropericlase [(Mg,Fe)O] and CaSiO3 (refs 1, 2, 3, 9). Ferropericlase is stable at most pressure and temperature conditions in the mantle; therefore, when found as a single inclusion within diamond, this mineral cannot be considered an unambiguous indicator of a super-deep origin7.

The CaSiO3 phase found within super-deep diamonds typically has the crystal structure of walstromite (BaCa2Si3O9)1,6,8,9. Perovskite-structured CaSiO3 (Ca-Pv) is considered one of the most important components in the Earth’s lower mantle, comprising approximately 7% of the peridotitic mantle and about 23% of the volume of a subducted mid-ocean ridge basalt slab9,10,11. As such, it is likely to be the fourth most abundant terrestrial mineral. Within the peridotitic lower mantle, Ca-Pv is the dominant sink for Ca and for incompatible elements, including the key heat-producing elements K, U and Th (ref. 12). However, Ca-Pv has so far never been found in nature and even high-pressure laboratory experiments have failed to quench it to a metastable phase at the conditions of the Earth’s surface. Although early studies1,2,3,4 of super-deep diamonds make a clear case for the presence of Ca-Pv in the transition zone and lower mantle, the structure of this phase was either undetermined or documented to be the lower-pressure polymorph—CaSiO3 walstromite—and interpreted as a back-transformation of perovskite-structured CaSiO3. The phase transformation from Ca-Pv to CaSiO3 walstromite would require a volume change13 of about 28%, which is impossible for diamond to accommodate owing to its extremely high bulk modulus14. The absence of healed fractures in the diamond host reported in ref. 13 implies that CaSiO3 walstromite is unlikely to represent inverted Ca-Pv. Plastic deformation of the diamond lattice could accommodate some of the volume change necessary for the phase transformation of the inclusion. However, although plastic deformation in super-deep diamonds has been well documented15 and is expected to be substantial, it has never been quantified. Some super-deep diamonds with documented phase assemblages that include ferropericlase, enstatite (inverted bridgmanite) or CaSiO3 walstromite probably originate from lower-mantle depths1,2,3,4,9, but ambiguity remains. Therefore, finding an un-retrogressed Ca-Pv would provide confirmation of lower-mantle sampling by super-deep diamonds.

In this study we investigated an inclusion within a diamond from the Cullinan mine in the Gauteng province of South Africa. The Cullinan kimberlite is a group I kimberlite; that is, its chemistry and Sr, Nd and Hf isotope signatures are thought to reflect a melt source beneath the lithospheric mantle, within the Earth’s convecting mantle16. The Cullinan mine is renowned for producing exceptionally large diamonds (such as the 3,107-carat Cullinan diamond6,17), most of which have been suggested to be super-deep diamonds6.

The diamond examined here has a 31 μm × 26 μm × 10 μm CaSiO3 inclusion, which was exposed by polishing. X-ray diffraction, Raman spectroscopy and electron backscatter diffraction (EBSD) reveal the CaSiO3 in this inclusion to have a perovskite structure. To our knowledge, this represents the only finding of non-reverted Ca-Pv in nature and the first Ca-Pv, including those synthesized in the laboratory, to preserve its high-pressure structure at the surface of the Earth.

Cathodoluminescence imaging of the host diamond surrounding the Ca-Pv inclusion (Fig. 1) reveals multiple growth zones and a complex internal structure, typical of super-deep diamonds4,18. Fourier transform infrared (FTIR) spectroscopy (Extended Data Fig. 1) of the diamond host indicates a nitrogen content of 34 p.p.m., with 97% in the B-aggregated form; that is, the diamond host is type IaB. The low nitrogen content and very high level of B aggregation are typical characteristics of super-deep diamonds4,19, indicating prolonged exposure to the high temperatures that are prevalent at transition-zone and lower-mantle depths.

Figure 1: Cathodoluminescence image and carbon isotopic composition of the diamond containing the Ca-Pv inclusion.
Figure 1

The Ca-Pv inclusion is shown in yellow. Carbon isotopic compositions measured at five locations (red circles) are expressed as δ13C values.

The chemical composition of the Ca-Pv inclusion, determined by electron microprobe analysis, is almost pure CaSiO3 (Ca0.98Si0.98O3), with minor impurities of Ti, Al, Fe and Mg totalling 0.04 atoms per formula unit (Extended Data Table 1).

Backscattered-electron imaging and energy-dispersive X-ray spectroscopy (EDS) element maps (Fig. 2) show that the Ca-Pv crystal includes 14 irregular areas of CaTiO3 perovskite with sizes between 1 μm and 7–8 μm and an approximate stoichiometry of Ca(Ti0.92Si0.07Al0.02)O3. The texture, size and abundance of these CaTiO3 intergrowths are very similar to those of inclusions reported in CaSiO3 walstromite phases in super-deep diamonds20 from Juina, Brazil. The exposed surface of our Ca-Pv inclusion makes accurate estimation of its bulk composition difficult, but image analysis indicates that the host crystal in bulk may contain up to 6% by volume CaTiO3. CaTiO3 perovskite is a common mineral in nature and remains stable well into the lower mantle21. By contrast, a Ca-Pv sample that retains its perovskite structure at room temperature and pressure has no experimentally synthesized analogues, unless a considerable amount of CaTiO3 (about 34 mol%; ref. 21) is dissolved within its structure, far more than the CaTiO3 component observed here. However, our discovery of natural Ca-Pv with less than 2 mol% CaTiO3 in the CaSiO3-rich portion of the inclusion indicates that, unlike experiments, nature must provide pressure–temperature–time pathways that are capable of preserving this metastable phase.

Figure 2: Backscattered-electron image of the Ca-Pv inclusion and energy-dispersive X-ray spectroscopy elemental maps.
Figure 2

a, Backscattered-electron image of the Ca-Pv inclusion (dark grey) surrounded by the diamond host (black), showing smaller inclusions of CaTiO3 perovskite (light grey). bd, Energy-dispersive X-ray spectroscopy elemental maps of Ca (b), Ti (c) and Si (d). The colour intensity (black within the grain outline through to saturation in a specific colour) is proportional to the element concentration.

As stated above, X-ray diffraction data show that the CaSiO3 inclusion has a perovskite structure. The small size of the inclusion (thickness ≤ 10 μm, as estimated by confocal Raman spectroscopy) and its entrapment within the diamond host resulted in only a limited number of measured diffraction reflections (n = 91), of which only nine were unique (Extended Data Table 2). All of the 91 reflections were used to refine the Ca-Pv unit-cell parameters: a = 5.397 ± 0.004 Å, b = 5.404 ± 0.004 Å, c = 7.646 ± 0.004 Å, volume 223.0 ± 0.03 Å3

However, alternative unit-cell refinements using other numerical approaches could provide considerably different (by more than 1%) unit-cell parameters owing to the relatively poor accuracy and precision with which the spacings between crystal planes (d spacings) were measured in this study. These relatively large uncertainties are typical when studying minerals of this size and arise not only from the limited number of reflections, but also because the measurements were performed using an area detector, which provides lower precision in d-spacing determination than a point detector. Such relatively large uncertainty on the cell parameters makes any comparison with the unit-cell volume of CaTiO3 perovskite unreliable, although we can establish that the structures of CaTiO3 perovskite and Ca-Pv are very similar. Ewald projections along the three crystallographic axes (Fig. 3a) indicate an orthorhombic unit cell. The unit cell and the chemical composition confirm that the mineral is Ca-Pv. Recent numerical simulations on ‘host–inclusion’ systems22 indicate that an inclusion partly exposed to atmospheric pressure loses only a portion of its residual pressure, depending on the elastic properties of both the host mineral and the inclusion. The Ca-Pv inclusion studied here is partly exposed at the diamond surface, but with two-thirds of its volume still buried in the diamond host. Thus, any measurements on this grain would be affected by some residual pressure still acting on the inclusion, which in turn affects the X-ray diffraction data and Raman spectra.

Figure 3: Ewald projections of X-ray diffraction data and Raman spectroscopy results.
Figure 3

a, Ewald projections along three different orientations for the Ca-Pv. b, Baseline-corrected Raman spectrum of the Ca-Pv inclusion compared with that of the CaTiO3-perovskite intergrowth (Fig. 2a).

Raman spectra (Fig. 3b) of the inclusion show that the spectrum of the CaTiO3 perovskite is in excellent agreement with Raman data for CaTiO3 perovskite from the RRUFF database23 (Extended Data Fig. 2). The CaSiO3 and CaTiO3 spectra are similar. Small differences are evident because of the presence of two Raman peaks for the CaSiO3 spectrum, which could belong to the lower-pressure CaSiO3 polymorph23 wollastonite-2M. This wollastonite polymorph is not stable at pressures higher than 3 GPa along a mantle geotherm24, well below the diamond stability field. Therefore, its presence is probably due to minor partial inversion of the Ca-Pv phase caused by the polishing of the sample to expose the inclusion, as reported previously25.

EBSD measurements conducted on several areas of the grain provide no evidence of amorphous portions (Fig. 4). The EBSD pattern of the CaSiO3 area (red circle), shown as the non-indexed EBSD pattern in Fig. 4b, is complex and could not be indexed by a single phase. The observed pattern can be indexed by using a combination of reference EBSD patterns for CaTiO3 perovskite (Fig. 4c) and CaSiO3 wollastonite-2M (Fig. 4d), again confirming that CaSiO3 is present in this diamond with a perovskite-type structure. Because EBSD measures surface responses (within tens of nanometres from the surface), this signal can only come from the CaSiO3 phase.

Figure 4: EBSD images of the Ca-Pv inclusion in diamond.
Figure 4

a, Location (indicated by the red filled circle) from which the EBSD images were obtained. b, Non-indexed EBSD pattern relative to the CaSiO3 inclusion. c, d, EBSD patterns indexed with the reference patterns of CaTiO3 (c) and CaSiO3 wollastonite-2M (d). The coloured lines in c represent the EBSD indexed pattern of CaTiO3, whereas the coloured lines in d represent the EBSD indexed pattern of wollastonite-2M. The numbers reported in both the figures at the intersections between the coloured lines represent the zone axes.

We suggest that the natural Ca-Pv found trapped within the super-deep diamond was a result of the unmixing of the high-pressure solid solution Ca(Ti,Si)O3. If the two phases exsolved from a homogeneous bulk composition, this phase would contain about 3.9% TiO2.

We estimate that the stoichiometry of the original phase composition was (Ca0.98Mg0.01Fe0.01)(Si0.93Ti0.06Al0.01)O3. This composition is consistent with that of CaSiO3 samples crystallized in experiments26 from a mid-ocean ridge basalt (MORB)-like bulk composition at about 26 GPa and is similar to that of CaSiO3 walstromite and CaTiO3 intergrowths found within Juina super-deep diamonds; these intergrowths have been suggested to originate from basalt-like compositions subducted to lower-mantle depths that later retrogressed during their ascent to the Earth’s surface20. The preservation of the high-pressure perovskite structure in the case of the Cullinan inclusion supports the derivation of such compositions from lower-mantle depths.

The possible subducted basaltic protolith origin of the Cullinan Ca-Pv inclusion suggests that we might expect to observe some evidence of a crustal parentage in the carbon isotopic composition of the host diamond (Fig. 1; Extended Data Table 3), which has δ13C values ranging from −2.3‰ to −4.6‰, where δ13C = (13C/12C)sample/(13C/12C)PDB − 1 (PDB, Pee Dee Belemnite reference material). The core region of the diamond, defined by cathodoluminescence imaging (Fig. 1), contains the Ca-Pv inclusion and has an average δ13C value of −2.3‰ ± 0.5‰, considerably lower than the typical upper-mantle value27 of −5.5‰. By contrast, the outer-rim region of the diamond has a composition (mean δ13C of −4.1‰ ± 0.5‰) that is closer to −5.5‰. Crustal carbon reservoirs have carbon isotopic compositions that are both ‘heavier’ and ‘lighter’ than the typical upper-mantle value. While ‘isotopically light’ carbon compositions (δ13C < −25‰) have been found in super-deep diamonds from Juina, which are thought to be derived from subducted basalt protoliths20,28, ‘isotopically heavy’ (−3‰ to −0.5‰) carbon compositions, such as those measured in the core of the studied diamond, have also been reported in super-deep diamonds from Brazil (Sao Luis and Juina) and Guinea (Kankan)18,19,20. If the δ13C value of −2.3‰ is compared with the median value (−4.91‰) of 1,473 published analyses of lithospheric diamonds containing peridotitic inclusions—a group of diamonds usually accepted to be minimally affected by subduction27—it is found to be an outlier, beyond three times the median absolute deviation. Such anomalously high carbon isotopic compositions are thought to reflect a greater influence of subducted carbonate in the fluid that formed these super-deep diamonds18,19. The carbon isotope compositions of the rim of the Cullinan diamond (Fig. 1) may represent an overgrowth that developed under upper-mantle conditions or from a distinct source of carbon in the lower mantle. Regardless, the high δ13C values of the portion of the diamond that contains the Ca-Pv inclusion supports the premise that it originates from a subducted basaltic protolith.

Our discovery of Ca-Pv in a super-deep diamond firmly establishes this phase as a component of the Earth’s deep mantle, confirming previous suggestions1,2,3,4,9 that lower-pressure CaSiO3 polymorphs included in these diamonds may represent retrogressed Ca-Pv. The estimated original bulk composition of the Cullinan Ca-Pv inclusion is consistent with compositions that are stable in subducted oceanic basalt protoliths at about 26 GPa, in the uppermost lower mantle26. Our finding thus confirms the expectation from calculations10 and high-pressure experiments21,25, that Ca-Pv is the main Ca-bearing phase in the lower mantle in both basic and ultrabasic compositions, reaching up to 23 vol% in MORB-like compositions26. The combined bulk composition of the Ca-Pv phase found here provides overwhelming evidence of the return of recycled oceanic crust into the Earth’s lower mantle20, whereas the relatively high carbon isotopic composition of the diamond in contact with the inclusion indicates the subduction of crustal carbon to lower mantle depths.

Methods

Micro-Raman spectroscopy

The Ca-Pv sample was analysed using an InVia Renishaw micro-Raman spectrometer installed at the Department of Chemical Sciences of the University of Padova. The spectra were baseline-corrected. A 632.8-nm-wavelength excitation laser was used at a power of 7 mW. The Raman spectrum of the Ca-Pv crystal was collected for 40 s using a 50× objective with a spatial resolution of 1.1 μm and a spectral resolution estimated to be about 3 cm−1. The most intense Raman peaks observed for the Ca-Pv inclusion are (in order of decreasing intensity): 774, 247, 470, 337, 181 and 226 cm−1.

Direct comparison between the Raman spectrum of natural Ca-Pv with that of CaTiO3—both from the CaTiO3 inclusions (Fig. 3) and from the RRUFF database23—indicates that the two spectra are very similar. A small but important difference is the presence of limited traces of wollastonite-2M in the Raman spectrum of the natural Ca-Pv (see peaks at 971 cm−1 and 637 cm−1), which, as expected, are not evident in the spectrum of the CaTiO3 perovskite inclusions.

On the basis of the results of ref. 29, the broad Raman bands in the 650–850 cm−1 region correspond to second-order Raman scattering and only the sharp peaks in the 200–500 cm−1 region are first-order Raman bands. However, we considered the entire Raman spectrum of natural Ca-Pv, regardless of first- or second-order scattering, for a direct comparison with CaTiO3 perovskite and wollastonite.

The strong similarity between the Raman spectra of the CaTiO3 and CaSiO3 perovskites in Fig. 3b could indicate that the spectra are dominated by emission from a larger, underlying, unexposed volume of crystalline CaTiO3 surrounded by a matrix of amorphous CaSiO3. This possibility can be discounted for a number of reasons. First, the partially exposed inclusion is under some stress and this will affect the Raman band shift, depending on the elastic properties of the two perovskites. More importantly, the Raman spectrum of such a hypothetical large amorphous area of CaSiO3 would be totally distinct from that measured here (Fig. 3b), and would be characterized by the presence of three very intense Raman bands at about 370 cm−1, 640 cm−1 and 970 cm−1 (depending on the pressure and temperature conditions30,31). These Raman bands are absent in the spectrum of the CaSiO3 portion of the perovskite-structured inclusion. Also, because of the confocal nature of the Raman measurements and their small spot size, they cannot be substantially influenced by the spatially associated CaTiO3 intergrowth. Last, the EBSD measurements rule out this possibility because EBSD is a surface technique (see Methods section ‘EBSD’).

Cathodoluminescence

The cathodoluminescence scanning electron microscopy image shown in Fig. 1 was obtained using a Philips XL 30 scanning electron microscope with a cathodoluminescence attachment consisting of a Hamamatsu R376 photomultiplier tube (EOAS UBC). The accelerating voltage was 20 keV and the electron beam current was 100 μA.

Infrared spectroscopy

Infrared spectra were collected on a Nicolet 6700 FTIR spectrometer. The absorbance spectra were measured at maximum light transmission for 40 s at a spectral resolution of 0.5 cm−1. Background spectra were collected for 120 s before the analysis and were subtracted from each measured absorbance spectrum. The nitrogen concentration and aggregation were determined using the procedure described in ref. 32 using a spreadsheet (‘FTIR analyser 3d’) created by J. Chapman (Rio Tinto Diamonds Ltd). Preliminary processing and baseline determination were performed using the EssentialFTIR software. The analytical and processing error was ±10% (1σ, relative error). The FTIR spectrum of the Cullinan diamond studied here is shown in Extended Data Fig. 1.

Electron microprobe analysis

Quantitative chemical analyses were performed at the Department of Earth, Ocean and Atmospheric Sciences of the University of British Columbia, using a fully automated CAMECA SX-50 electron microprobe operating in the wavelength-dispersion mode with the following operating conditions: excitation voltage, 15 kV; beam current, 20 nA; peak counting time, 20 s; background counting time, 10 s; actual spot diameter, 5 μm. Data reduction was done using the PAP φ(ρZ) method33. The detection limits for most oxides were lower than 0.08 wt% and those for Cr2O3, MnO2 and NiO were lower than 0.12 wt%. Because of the small crystal size of the natural Ca-Pv and the presence of CaTiO3 perovskite inclusions, we were able to perform only three reliable analyses; the results are reported in Extended Data Table 1. The Na and K contents were not analysed.

Scanning electron microscopy and EDS

We studied our sample using scanning electron microscopy and EDS to investigate the distribution of Ca, Si and Ti over the grain. We used a CamScan MX3000 electron microscope equipped with a LaB6 source, a four-quadrant solid-state backscattered-electron detector and an EDAX EDS system for micro-analysis installed at the Department of Geosciences of the University of Padova. The measurement conditions were: accelerating voltage, 20 kV; filament emission, about 13 nA; working distance (the distance between the specimen and the lowest part of the electromagnetic lens in the column of the scanning electron microscope), 27 mm. The backscattered-electron image of the grain and its EDS maps for Ca, Si and Ti are shown in Fig. 2.

Single-crystal micro-X-ray diffraction

Single-crystal X-ray diffraction measurements were performed at the Department of Geosciences of the University of Padova, using a Rigaku Oxford Diffraction Supernova goniometer equipped with a Dectris Pilatus 200 K area detector and a Mova X-ray micro-source (Mo Kα radiation) operating at 50 kV and 0.8 mA. The sample-to-detector distance was 68 mm. Data reduction was performed using the CrysAlis software (Rigaku Oxford Diffraction) to obtain the Ewald projections shown in Fig. 3. The diffraction analysis results are reported in Extended Data Table 2 in comparison with those of a reference CaTiO3 single-crystal sample34 with the following unit-cell parameters: a = 5.388 ± 0.001 Å, b = 5.447 ± 0.001 Å, c = 7.654 ± 0.001 Å, volume, 224.63 ± 0.001 Å3.

Carbon isotope analyses

The carbon isotope compositions (δ13C) reported in Extended Data Table 3 were determined using a Cameca IMS 7f-GEO secondary ion mass spectrometer. The polished diamond was pressed into an indium mount with a 1-inch-diameter aluminium holder. Natural reference diamonds with δ13C values between −13.6‰ (2σ = 0.3‰) and 2.6‰ (2σ = 0.3‰) were used to determine the instrumental mass fractionation and drift before and after sample analyses. The sample and reference diamonds were coated with gold (20 nm thickness). The measurements were conducted using 133Cs+ at 10 keV impact energy and a beam current of about 4 nA. The 15-μm-diameter 133Cs+ primary-ion beam was used for pre-sputtering. During the measurements, the ion beam diameter was reduced to 5 μm. Secondary ions of 12C and 13C were extracted at −9 keV with an energy bandwidth of 90 eV. No electron-gun charge compensation was required. The 13C/12C ratios were measured using dual Faraday cups. The mass resolving power MM was 2,900. The 12C and 13C ions were counted for 1 s in each cycle of the 30 cycles and the total measurement time for each spot was 8 min. The standard deviation of the analysis is estimated to be about 0.4‰ to 0.5‰ at the 2σ (95% uncertainty) level.

EBSD

EBSD analyses were performed at CNR-ICMATE in Padova, using a Quanta 200F FEG-ESEM system operating in high-vacuum mode with an accelerating voltage of 30 kV, emission current of 174 μA and beam spot of 4.5 μm, without any conductive coating. EBSD patterns were collected at a working distance of 10 mm and a specimen tilt of 75° using an EDAX Digiview EBSD system. The instrument was controlled by the OIMTM 5.31 software, which contains a large EBSD pattern database.

Statistical analysis of carbon isotope composition

We used a compilation of 1,473 carbon-isotope analysis datasets for diamonds containing inclusions of lithospheric peridotite paragenesis from ref. 35. We calculated the median absolute deviation36 of the data using a b factor of 1.4826 and a very conservative threshold factor of 3.

Data availability

All relevant data are presented in Extended Data Tables 1, 2, 3 and Figs 1, 2, 3, 4. Original spectral data and electron microprobe data are available from the corresponding author.

References

  1. 1.

    , , , & New Ca-silicate inclusions in diamonds – tracers from the lower mantle. Earth Planet. Sci. Lett. 173, 1–6 (1999)

  2. 2.

    in Mantle Petrology: Field Observations and High-Pressure Experimentation Vol. 6 (eds et al.) 125–153 (Geochemical Society, 1999)

  3. 3.

    , , & Kankan diamonds (Guinea) II: lower mantle inclusion paragenesis. Contrib. Mineral. Petrol. 140, 16–27 (2000)

  4. 4.

    , & Lower mantle diamonds from Rio Soriso (Juina area, Mato Grosso, Brazil). Contrib. Mineral. Petrol. 149, 430–445 (2005)

  5. 5.

    et al. Hydrous mantle transition zone indicated by ringwoodite included within diamond. Nature 507, 221–224 (2014)

  6. 6.

    et al. Large gem diamonds from metallic liquid in Earth’s deep mantle. Science 354, 1403–1405 (2016)

  7. 7.

    , , , & Ferropericlase—a lower mantle phase in the upper mantle. Lithos 77, 655–663 (2004)

  8. 8.

    , , & Slab melting as a barrier to deep carbon subduction. Nature 529, 76–79 (2016)

  9. 9.

    & Mineral associations in diamonds from the lowermost upper 254 mantle and uppermost lower mantle. In Proc. of the 10th International Kimberlite Conference Vol. 1 (eds et al.) 235–253 (Springer, 2013)

  10. 10.

    & Geophysics of chemical heterogeneity in the mantle. Annu. Rev. Earth Planet. Sci. 40, 569–595 (2012)

  11. 11.

    Composition and Petrology of the Earth’s Mantle (McGraw-Hill, 1975)

  12. 12.

    & Trace element partitioning and substitution mechanisms in calcium perovskites. Contrib. Mineral. Petrol. 149, 85–97 (2005)

  13. 13.

    et al. Depth of formation of CaSiO3-walstromite included in super-deep diamonds. Lithos 265, 138–147 (2016)

  14. 14.

    , , & Diamond thermoelastic properties and implications for determining the pressure of formation of diamond-inclusion systems. Russ. Geol. Geophys. 56, 211–220 (2015)

  15. 15.

    , , & Plastic deformation of lower mantle diamonds by inclusion phase transformations. Eur. J. Mineral. 20, 333–339 (2008)

  16. 16.

    et al. Hf isotope systematics of kimberlites and their megacrysts: new constraints on their source region. J. Petrol. 45, 1583–1612 (2004)

  17. 17.

    The origin of large irregular gem-quality type II diamonds and the rarity of blue type IIb varieties. S. Afr. J. Geol. 117, 219–236 (2014)

  18. 18.

    , , , & Isotopic constraints on the nature and circulation of deep mantle C-H-O-N fluids: carbon and nitrogen systematics within ultra-deep diamonds from Kankan (Guinea). Geochim. Cosmochim. Acta 139, 26–46 (2014)

  19. 19.

    , , & Kankan diamonds (Guinea) III: δ13C and nitrogen characteristics of deep diamonds. Contrib. Mineral. Petrol. 142, 465–475 (2002)

  20. 20.

    et al. Deep mantle cycling of oceanic crust: evidence from diamonds and their mineral inclusions. Science 334, 54–57 (2011)

  21. 21.

    , & High pressure phase equilibria in the system CaTiO3-CaSiO3: stability of perovskite solid solutions. Phys. Chem. Miner. 24, 488–494 (1997)

  22. 22.

    et al. Elastic geothermobarometry: corrections for the geometry of the host-inclusion system. Geology (2018)

  23. 23.

    , , & in Highlights in Mineralogical Crystallography (eds & ) 1–30 (De Gruyter, 2016)

  24. 24.

    , & Experimental determination of phase relations in the CaSiO3 system from 8 to 15 GPa. Am. Mineral. 79, 1219–1222 (1994)

  25. 25.

    & Synthesis of majorite and other high pressure garnets and perovskites. Earth Planet. Sci. Lett. 12, 411–418 (1971)

  26. 26.

    & Subsolidus and melting phase relations of basaltic composition in the uppermost lower mantle. Geochim. Cosmochim. Acta 66, 2099–2108 (2002)

  27. 27.

    , , & Diamond formation: a stable isotope perspective. Annu. Rev. Earth Planet. Sci. 42, 699–732 (2014)

  28. 28.

    et al. Stable isotope evidence for crustal recycling as recorded by superdeep diamonds. Earth Planet. Sci. Lett. 432, 374–380 (2015)

  29. 29.

    & The Raman spectra of several orthorhombic calcium oxide perovskites. Phys. Chem. Miner. 16, 21–28 (1988)

  30. 30.

    , , , & Structural analysis of CaSiO3 glass by X-Ray diffraction and Raman spectroscopy. J. Non-Cryst. Solids 80, 167–174 (1986)

  31. 31.

    , & Melting of CaSiO3 perovskite to 430 kbar and first in-situ measurements of lower mantle eutectic temperatures. Geophys. Res. Lett. 24, 909–912 (1997)

  32. 32.

    & Geologically significant information from routine analysis of the mid-infrared spectra of diamonds. Int. Geol. Rev. 37, 95–110 (1995)

  33. 33.

    & in Microbeam Analysis (ed. ) 104–106 (San Francisco Press, 1985)

  34. 34.

    & Electron difference density and structural parameters in CaTiO3. Acta Crystallogr. C 48, 644–649 (1992)

  35. 35.

    , & Sources of carbon in inclusion bearing diamonds. Lithos 112, 625–637 (2009)

  36. 36.

    Robust Statistics 107–108 (Wiley, 1981)

Download references

Acknowledgements

We thank M. Regier for proofreading the paper. F.N. is supported by the European Research Council (ERC) Starting Grant number 307322. M.K.’s work and sample collection was possible thanks to an NSERC Discovery grant. N.K. acknowledges funding from the Dr. Eduard Gübelin Association through a 2015 research scholarship. D.G.P. was funded by an NSERC CERC award. M.A. was supported by the ERC under the European Union’s Horizon 2020 research and innovation programme (grant 714936) ‘TRUE DEPTHS’ and by the SIR-MIUR grant (RBSI140351) ‘MILE DEEp’. We thank L. Litti and M. Meneghetti of the Laboratory of Nanostructures and Optics of the Department of Chemical Sciences, University of Padova for their help in acquiring and interpreting the Raman data. F.N. and D.G.P. were supported by the Deep Carbon Observatory. M.G.P. was supported by NERC grant NE/M015181/1.

Author information

Affiliations

  1. Dipartimento di Geoscienze, Università degli Studi di Padova, Via Giovanni Gradenigo 6, I-35131 Padova, Italy

    • F. Nestola
  2. Department of Earth, Ocean and Atmospheric Sciences, University of British Columbia, Vancouver, British Columbia V6T 1Z4, Canada

    • N. Korolev
    •  & M. Kopylova
  3. Institute of Precambrian Geology and Geochronology RAS, 199034 St Petersburg, Russia

    • N. Korolev
  4. Dipartimento di Scienze della Terra, Università degli Studi di Milano, Via Botticelli 23, I-20133 Milano, Italy

    • N. Rotiroti
  5. Department of Earth and Atmospheric Sciences, University of Alberta, Edmonton, Alberta T6G 2E9, Canada

    • D. G. Pearson
  6. Department of Earth Sciences, University College London, Gower Street, London WC1E 6BT, UK

    • M. G. Pamato
  7. Department of Earth and Environmental Sciences, University of Pavia, Via Ferrata 1, I-27100 Pavia, Italy

    • M. Alvaro
  8. CNR-Istituto di Geoscienze e Georisorse, Sezione di Padova, Via Giovanni Gradenigo 6, I-35131 Padova, Italy

    • L. Peruzzo
  9. University of Cape Town, Cape Town, South Africa

    • J. J. Gurney
  10. Rhodes University, Grahamstown, South Africa

    • A. E. Moore
  11. Petra Diamonds, Bryanston, South Africa

    • J. Davidson

Authors

  1. Search for F. Nestola in:

  2. Search for N. Korolev in:

  3. Search for M. Kopylova in:

  4. Search for N. Rotiroti in:

  5. Search for D. G. Pearson in:

  6. Search for M. G. Pamato in:

  7. Search for M. Alvaro in:

  8. Search for L. Peruzzo in:

  9. Search for J. J. Gurney in:

  10. Search for A. E. Moore in:

  11. Search for J. Davidson in:

Contributions

F.N. conceived the study, wrote the initial manuscript and performed X-ray diffraction and micro-Raman measurements. N.K. found the mineral, made original mineral identifications on a confocal Raman spectrometer, performed microprobe and cathodoluminescence measurements, prepared samples for secondary ion mass spectrometry measurements and assisted with the manuscript preparation. M.K. supervised the study of the Cullinan diamond collection, which was acquired by J.J.G., A.E.M. and J.D., and assisted with the manuscript preparation. D.G.P. made the geochemical interpretations and led the manuscript revisions. M.G.P. assisted with the manuscript preparation and crystallographic interpretations. N.R., M.G.P. and M.A. assisted with the X-ray data interpretation. L.P. collected and interpreted the EBSD data. J.J.G., A.E.M. and J.D. designed the sampling programme.

Competing interests

The authors declare no competing financial interests.

Corresponding author

Correspondence to F. Nestola.

Reviewer Information Nature thanks B. Harte and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.