Advertisement

Lower mantle “garbage can”

Calcium silicate perovskite has finally been identified in a natural sample and now has the mineral name davemaoite. Tschauner et al. discovered the type mineral trapped at high pressure and temperature as a diamond inclusion (see the Perspective by Fei). Structural and chemical analysis of the mineral showed that it is able to host a wide variety of elements, not unlike fitting bulky objects into garbage can. Specifically, it has a large amount of trapped potassium. Davemaoite can thus host three of the major heat-producing elements (uranium and thorium were previously shown experimentally) affecting heat generation in Earth’s lower mantle. —BG

Abstract

Calcium silicate perovskite, CaSiO3, is arguably the most geochemically important phase in the lower mantle, because it concentrates elements that are incompatible in the upper mantle, including the heat-generating elements thorium and uranium, which have half-lives longer than the geologic history of Earth. We report CaSiO3-perovskite as an approved mineral (IMA2020-012a) with the name davemaoite. The natural specimen of davemaoite proves the existence of compositional heterogeneity within the lower mantle. Our observations indicate that davemaoite also hosts potassium in addition to uranium and thorium in its structure. Hence, the regional and global abundances of davemaoite influence the heat budget of the deep mantle, where the mineral is thermodynamically stable.
Calcium silicate, CaSiO3, occurs in a variety of natural and synthetic polymorphs. The low-pressure polymorph wollastonite is a common metamorphic mineral. Breyite (1), an intermediate-pressure polymorph (2), has been found as inclusions in diamond. At the pressures and depth range of Earth’s transition zone (420 to 660 km) and lower mantle (LM; 660 to ~2700 km), CaSiO3 assumes a perovskite structure. Perovskite-type CaSiO3, first synthesized by Liu and Ringwood (3), is a liquidus phase for basaltic and peridotic bulk rock compositions at LM pressures and temperatures and has been experimentally shown to host many elements that are incompatible in upper-mantle minerals (47). These include rare-earth elements (REEs), large ion lithophile elements (LILEs; K, Sr, and Ba), Ti, U, and Th. In other words, these elements are compatible rather than incompatible in a LM mineral assemblage that contains a few vol % of CaSiO3-perovskite. The composition and abundance of this phase in the LM are therefore key in constraining the budget and distribution of REEs and LILEs and the elements with abundant radioactive isotopes (K, U, and Th) that make an important contribution to the heat of Earth’s mantle (8). Through these parameters, CaSiO3 perovskite provides essential constraints on the fate of recycled crust in deep Earth, thermochemical anomalies, and the existence of a magma ocean at the base of Earth’s mantle. The synthetic perovskite phase of pure CaSiO3 has been found to assume either cubic or tetragonal symmetry (9) and belongs to the tausonite (SrTiO3)–type perovskites that adhere to fundamentally different structural distortion mechanisms (10) and crystal-chemical constraints than the GdFeO3-type perovskites such as bridgmanite [MgSiO3-perovskite (11)] and the CaTiO3 mineral, actually named perovskite.
The difficulty of finding CaSiO3-perovskite in nature stems from its stability at pressures only above 20 GPa (3, 4) along with a low kinetic barrier for back-conversion into low-pressure polymorphs (2). This barrier is lower than for bridgmanite, which has been found as a rare occurrence in highly shocked meteorites despite its stability only above 23 GPa (11, 12). Nestola et al. (13) reported the presence of CaSiO3-perovskite as an inclusion in a diamond from the Cullinan mine, South Africa. The reported phase deviates from synthetic CaSiO3-perovskite in several ways: (i) its volume at ambient conditions is >20% larger (9); (ii) it sustains the beam of an electron microscope, whereas any synthetic CaSiO3-perovskite vitrifies rapidly at ambient conditions; (iii) its cell axis ratios and Raman spectrum are nearly equal to those of CaTiO3; and (iv) its space group indicates a structural distortion mechanism different from that of synthetic CaSiO3-perovskite but much closer to that of CaTiO3 (7, 8). Nestola et al. (13) proposed that this distinctive phase of CaSiO3 is the result of partial decomposition of a Ti-bearing CaSiO3-perovskite. The coexistence of CaTiO3 + CaSiO3 polymorphs in diamond inclusions may also point to retrograde transformation of stoichiometric Ca-Si-Ti-perovskites (1) that form in the deep upper mantle (5 to 10 GPa).
The findings by Nestola et al. (13) are notable by themselves but differ from the expected high-pressure CaSiO3-perovskite and have not resulted in the approval of CaSiO3-perovskite as a mineral. We report the discovery of CaSiO3-perovskite as a mineral approved by the Commission of New Minerals, Nomenclature and Classification (CNMNC) of the International Mineralogical Association (IMA). The new mineral [IMA2020-012a (14)] is named “davemaoite” in honor of Dave (Ho-kwang) Mao for his eminent contributions to the field of deep-mantle geophysics and petrology. The type material—inclusions in a diamond from Orapa, Botswana—is deposited in the Natural History Museum Los Angeles (catalog number NHMLA 74541, formerly GRR1507 of the Caltech mineral collection) (15). Davemaoite coexists with orthorhombic carbonaceous α-iron and wüstite (Fe0.8Mg0.2)O at 8 to 9 GPa remnant pressure (Fig. 1). Separate inclusions of ilmenite, iron, and ice-VII in the same diamond (16) have remnant pressures of 7 GPa and 8 to 9 GPa, respectively. The x-ray diffraction (XRD) pattern of davemaoite is that of a cubic perovskite (Fig. 1) (15), with, at most, contributions of <5 vol % of material with ±2.5% smaller or larger volume, whereas an overall distortion of the lattice can be excluded on the basis of the reflection intensities (Fig. 1). Cubic ABO3 perovskites have no internal structural degrees of freedom and comprise only one chemical formula unit. Thus, the identification of this phase is unambiguous even in diffraction patterns with contributions from more than one phase (Fig. 1).
Fig. 1. X-ray diffraction pattern and Rietveld refinement of davemaoite.
Davemaoite (dvm), red; (Fe,C), olive; wüstite (wüs), blue; residual of fit, green. Tick marks indicate allowed reflections (same color-coding). The weighted refinement factor, Rwp, was 0.046, and the structure-factor moduli-based refinement factor, RF, for dvm was 0.10. Phase proportions were 35(5), 45(5), and 15(5)% for dvm, (Fe,C), and wüs, respectively. In dvm, the intensity of reflections 100 and 201 indicates sublattice disorder with A-cations shifted from site 1b to the 1/8-occupied site 8g. This is consistent with accommodation of K and Na along with Al and Fe. The paragenesis was found in two inclusions. Volumes of analyzed davemaoite inclusions ranged from 45.1 to 46.3 Å3 [equal to 0 to 6 GPa (7) in pure CaSiO3], but the actual composition of davemaoite markedly changes the pressure inferred from the volume (9). Coexisting wüstite Fe0.8(1)Mg0.2(1)O has a volume of 74.34(1) Å3, corresponding to 8 to 9 GPa (22). Indications of weak diffraction around dvm peaks near noise level are assigned to minor contributions of material of slightly larger and smaller volume (probably reflecting chemical variation; fit not shown here) but are inconsistent with a tetragonal distortion. a.u., arbitrary units. (Inset) IR spectrum of davemaoite after subtraction of a diamond spectrum collected close to the inclusion. The unprocessed IR spectrum is shown in fig. S3. Bars indicate all calculated IR bands.
Davemaoite was identified through the XRD pattern of cubic perovskite at a location in the hosting diamond with a CaKα x-ray fluorescence (XRF) signal far above background. Both, XRF and XRD data were obtained at beamline 34-ID-E at the Advanced Photon Source (15). We superimposed the CaKα XRF map (Fig. 2) on a visible light image of the holotype material at the beamline where we subsequently collected the XRF and XRD data. We also made corresponding maps of Fe and Ti (fig. S1). Areas with XRF signal at noise level in Fig. 2 show no x-ray diffraction besides that of diamond. Right after acquisition of the XRF map, we examined by XRD the inclusions that were found by XRF. XRD and XRF were collected on the inclusions when they were fully entrapped in the doubly polished platelet of the hosting diamond. We focused the x-ray beam to an area of 0.5 μm by 0.5 μm in order to identify inclusions with high spatial resolution. The inclusions of 4 μm by 6 μm and 4 μm by 16 μm areas within the red circle of the Ca XRF map corresponded to the XRD patterns of the cubic perovskite (Fig. 1). We added frames with perovskite patterns to obtain better signal and powder statistics.
Fig. 2. Reflected light image of the diamond at the beamline.
The XRF map of CaKα is superimposed. The x and y axes give the sample coordinates in micrometers. The red circle indicates the area ablated during LA-ICP-MS analysis. The two Ca-rich areas in the circle correspond to diffraction patterns of davemaoite. The more intense signal around x,y = 72,−3074 corresponds to the shallow inclusion at 8 to 10 μm depth. Depth was assessed from the attenuation of the XRF signal of Ca and Fe and from the depth of the pit after ablation. (Inset) Time-resolved 44Ca signal of LA-ICP-MS measurement. Ablation started at time (t) = 48 s and finished at 108 s. Cps, counts per second; div, divisions of the grid.
We confirmed the identification of davemaoite by infrared spectroscopy. Cubic ABO3 perovskites have no Raman-active modes and three infrared (IR)–active modes (15). We observed two of the three IR-active modes (Fig. 1, inset), while the third one was below the diffraction limit for objects as small as these inclusions. As expected, we observed no Raman peaks (fig. S2). We calculated mode energies by fitting force constants to match ab initio calculated zone-center phonon energies of the tetragonal structure (17) and mapped the tetragonal Brillouin zone onto that of the cubic structure. Intensities are based on the calculated phonon density of state.
Subsequently, we used laser-ablation inductively coupled mass spectrometry (LA-ICP-MS) to excavate and analyze the chemical compositions of two of the inclusions with a 100-μm-diameter laser beam. We indicate the ablation area with a red circle (Fig. 2). We monitored all mass peaks under medium mass resolution (mm = 4000), which allowed us to resolve numerous carbon-related molecular interferences on low-mass isotopes. We hit two inclusions of davemaoite at 5 to 8 μm and 80 to 100 μm depth below the polished surface (Fig. 2, inset). The time-resolved 44Ca signal of the LA-ICP-MS measurement (Fig. 2, inset) shows signal clearly above background level. We also collected equivalent profiles for 56Fe, 39K, and 52Cr (fig. S1). In all cases, the signals rose above background at the same depth, consistent with their origin in the same two inclusions. We obtained an average davemaoite composition of (Ca0.43(1)K0.20(1)Na0.06Fe0.11(1)Al0.08Mg0.06Cr0.04(2))(Si1.0(2)Al0.00(1))O3 (15). We performed a Rietveld refinement (Fig. 1) and provide a crystallographic information file (15).
Next, we describe the hosting diamond and then discuss the pressure-temperature (P-T) conditions of formation of the type davemaoite and its composition. The hosting diamond is different from diamonds that contain potential retrograde products of high-pressure minerals at 0 to 1 GPa remnant pressures (13, 18). The diamond is of type IaAB with frosted octahedral faces and trigon features (15). We found davemaoite, iron, wüstite, ilmenite, and ice-VII in the center of the diamond. Our analysis of the N-defect IR bands (fig. S3) (12) indicates a low average residence temperature (~1500 K) or a short residence time in the mantle, similar to the holo- and cotype diamonds of ice-VII (18). Short residence time and low average residence temperature are common features of lithospheric diamonds, but in sublithospheric diamonds both parameters act in favor of conserving high remnant pressures and high-pressure minerals by reducing viscoelastic relaxation of the hosting diamond and by preventing retrograde transformations. The bulk modulus of davemaoite depends on its composition (7, 8) and is unknown for the given composition. However, coexisting wüstite is at a remnant pressure of 8 to 9 GPa (16). For a single inclusion of wüstite, this remnant pressure would correspond to an entrapment pressure of ~40 GPa if pressure were to evolve along a purely elastic path (15). However, diamond becomes viscoelastic between 1100 and 1200 K, even at laboratory time scales (19). To account for this nonelastic behavior of the hosting diamond, we use the method of Wang et al. (20), which does not rely on initial assumptions about the entrapment temperature and uses the P-T paths of separate inclusions in the same diamond: wüstite, iron, ice-VII, and ilmenite. Using this approach, we assessed entrapment conditions of 29 ± 5 GPa at 1400 to 1600 K (15). Because viscoelastic processes are path- and time-dependent, we cannot exclude a higher entrapment pressure or temperature.
We cannot entirely rule out that our chemical analysis is affected by minor contaminants, although we did not observe an XRD signal or a marked XRF signal of any phase other than davemaoite, wüstite, and iron in the excavated region. Furthermore, we note that (i) the low Ti is a result unaffected by potential contamination and (ii) the 39K signal occurs at the same depth as the 44Ca signal of davemaoite. We did not observe by XRD alternative hosts of K and Ca such as liebermannite and harmunite-type (Ca,K,Na)(Al,Si)2O4 anywhere in this diamond. Both of these phases, and any phase dominated by K and Ca, give diffraction patterns that are very different from the perovskite-type pattern we observed. Hence, we believe the presence of K and Al in davemaoite is not likely the result of a contaminated analysis but rather indicates coupled substitution of a large and a small cation K,Na + Al,Fe for Ca. Generally, a substitution of K for Ca and Al for Si shifts the material into the stability field of ABO3 perovskites with a trend toward high crystal symmetry (21).
Our result indicates that the postspinel phase (Ca,Na,K)(Al,Si)2O4 is not required as a host of Ca, alkalis, and Al in at least the upper region of the lower mantle. It is possible that type davemaoite formed retrograde out of postspinel through a reaction (Ca,Na,K)(Al,Fe3+,Si)2O4 + Fe0 → (Ca,Na,K)(Al,Si)O3 + FeO, but for this process one expects the presence of remnant postspinel in the paragenesis, which is not observed. In the deep mantle, davemaoite takes on a role similar to that of garnet in the upper mantle. Both minerals have a “garbage can” crystal chemistry that allows them to host many elements that are incompatible in upper-mantle minerals (6, 7). Our observations are fully consistent with the experimental results that this mineral dissolves LILEs, specifically K (6, 7). Experimental studies that were based on peridotite and MORB (mid-ocean ridge basalt)–like bulk compositions formed davemaoite with lower K content and higher Ti content than the type material, which is expected for these starting compositions. We argue that the low Ti and high K content of type davemaoite reflects a different, K-rich source composition, possibly resulting from deep-mantle metasomatism, which is also indicated by the presence of ice-VII and by the hosting diamond itself (16). This point emphasizes the importance of studying natural specimens of high-pressure minerals, because they record a petrologic complexity of deep Earth that may not be assessed in experiments. Depletion of Ti in type davemaoite is a possible result of the presence of phases that strongly partition Ti, such as liuite, FeTiO3-perovskite. Ilmenite has been observed in the same diamond at similar remnant pressure as the davemaoite-wüstite-iron inclusion, and its P-T path intersects the phase boundary of liuite (15). Hence, our findings indicate that the source rock composition of type davemaoite deviated from peridotite and, thus, that chemical segregation occurs in the lower mantle, possibly down to 900 km according to our estimate (fig. S4). This variation in rock composition affects heat generation through radioactive decay in the lower mantle where davemaoite scavenges K, as shown here, and U and Th, as experimentally shown (7).

Acknowledgments

We thank N. Tomioka and an anonymous reviewer for their helpful comments.
Funding: This work was supported by awards NSF-EAR-1838330, -EAR-1942042, and -EAR-1322082; the NSF Cooperative Agreement No. DMR-1644779; and the State of Florida. Use of the Advanced Photon Source and the Advanced Light Source were supported by the US Department of Energy, Basic Energy Sciences, contracts DE-AC02-06CH11357 and DE-AC02-05CH11231, respectively.
Author contributions: O.T., S.H., S.Y., and M.H. participated in design, interpretation, data collection, and analysis of the reported results and in drafting and revising the manuscript. W.L., S.N.G.C., H.A.B., J.T., and G.R.R. participated in data collection and revising the manuscript.
Competing interests: The authors have no competing interests.
Data and materials availability: Additional chemical and crystallographic information about davemaoite is provided in the supplementary materials. Raw data are deposited at Dryad (23). Crystallographic and chemical information on type davemaoite is deposited in the Inorganic Crystal Structure Database. The type material is deposited with the NHMLA under accession number 74541.

Supplementary Materials

This PDF file includes:

Materials and Methods
Figs. S1 to S4
Tables S1 and S2
References (2440)

Other Supplementary Material for this manuscript includes the following:

Data S1

References and Notes

1
F. E. Brenker, F. Nestola, L. Brenker, L. Peruzzo, J. W. Harris, Origin, properties and structure of breyite: The second most abundant mineral inclusion in super-deep diamonds. Am. Mineral. 106, 38–43 (2020).
2
O. Tschauner, High-pressure minerals. Am. Mineral. 104, 1701–1731 (2019).
3
L. Liu, A. E. Ringwood, Synthesis of a perovskite polymorph of CaSiO3. Earth Planet. Sci. Lett. 28, 209–211 (1975).
4
K. Hirose, R. Sinmyo, J. Hernlund, Perovskite in Earth’s deep interior. Science 358, 734–738 (2017).
5
S. Gréaux, T. Irifune, Y. Higo, Y. Tange, T. Arimoto, Z. Liu, A. Yamada, Sound velocity of CaSiO3 perovskite suggests the presence of basaltic crust in the Earth’s lower mantle. Nature 565, 218–221 (2019).
6
K. Hirose, N. Shimizu, W. van Westrenen, Y. W. Fei, Trace element partitioning in Earth’s lower mantle and implications for geochemical consequences of partial melting at the core-mantle boundary. Phys. Earth Planet. Inter. 146, 249–260 (2004).
7
S. Gréaux, L. Gautron, D. Andrault, N. Bolfan-Casanova, N. Guignot, M. A. Bouhifd, Experimental high pressure and high temperature study of the incorporation of uranium in Al-rich CaSiO3 perovskite. Phys. Earth Planet. Inter. 174, 254–263 (2009).
8
J. Korenaga, Urey ratio and the structure and evolution of Earth’s mantle. Rev. Geophys. 46, RG2007 (2008).
9
H. W. Chen, S. H. Shim, K. Leinenweber, V. B. Prakapenka, Y. Meng, C. Prescher, Crystal structure of CaSiO3 perovskite at 28–62 GPa and 300 K under quasi-hydrostatic stress conditions. Am. Mineral. 103, 462–468 (2018).
10
A. M. Glazer, Simple ways of determining perovskite structures. Acta Crystallogr. A 31, 756–762 (1975).
11
O. Tschauner, C. Ma, J. R. Beckett, C. Prescher, V. B. Prakapenka, G. R. Rossman, Discovery of bridgmanite, the most abundant mineral in Earth, in a shocked meteorite. Science 346, 1100–1102 (2014).
12
N. Tomioka, K. Fujino, Natural (Mg,Fe)SiO3-ilmenite and -perovskite in the Tenham meteorite. Science 277, 1084–1086 (1997).
13
F. Nestola, N. Korolev, M. Kopylova, N. Rotiroti, D. G. Pearson, M. G. Pamato, M. Alvaro, L. Peruzzo, J. J. Gurney, A. E. Moore, J. Davidson, CaSiO3 perovskite in diamond indicates the recycling of oceanic crust into the lower mantle. Nature 555, 237–241 (2018).
14
R. Miyawaki, F. Hatert, M. Pasero, S. J. Mills, IMA Commission on New Minerals, Nomenclature and Classification (CNMNC) – Newsletter 58. Eur. J. Mineral. 32, 645–651 (2020).
15
Materials and methods are available as supplementary materials.
16
O. Tschauner, S. Huang, E. Greenberg, V. B. Prakapenka, C. Ma, G. R. Rossman, A. H. Shen, D. Zhang, M. Newville, A. Lanzirotti, K. Tait, Ice-VII inclusions in diamonds: Evidence for aqueous fluid in Earth’s deep mantle. Science 359, 1136–1139 (2018).
17
R. Caracas, R. Wentzcovitch, G. D. Price, J. Brodholt, CaSiO3 perovskite at lower mantle pressures. Geophys. Res. Lett. 32, L06306 (2005).
18
T. Stachel, G. P. Brey, J. W. Harris, Inclusions in sublithospheric diamonds: Glimpses of deep Earth. Elements 1, 73–78 (2005).
19
D. J. Weidner, Y. Wang, M. T. Vaughan, Strength of diamond. Science 266, 419–422 (1994).
20
W. Wang, O. Tschauner, S. Huang, Z. Wu, Y. Meng, H. Bechtel, H. K. Mao, Coupled deep-mantle carbon-water cycle: Evidence from lower-mantle diamonds. Innovation 2, 100117 (2021).
21
V. M. Goldschmidt, Die Gesetze der Krystallochemie. Naturwissenschaften 14, 477–485 (1926).
22
S. D. Jacobsen, J. F. Lin, R. J. Angel, G. Shen, V. B. Prakapenka, P. Dera, H. K. Mao, R. J. Hemley, Single-crystal synchrotron X-ray diffraction study of wüstite and magnesiowüstite at lower-mantle pressures. J. Synchrotron Radiat. 12, 577–583 (2005).
23
O. Tschauner, S. Huang, S. Yang, M. Humayun, W. Liu, S. N. Gilbert Corder, H. A. Bechtel, J. Tischler, G. R. Rossman, Raw data for davemaoite Dryad (2021); https://doi.org/10.5061/dryad.jq2bvq89m.
24
A. Rohrbach, M. W. Schmidt, Redox freezing and melting in the Earth’s deep mantle resulting from carbon-iron redox coupling. Nature 472, 209–212 (2011).
25
E. Kroumova, M. I. Aroyo, J. M. Perez-Mato, A. Kirov, C. Capillas, S. Ivantchev, H. Wondratschek, Bilbao Crystallographic Server: Useful databases and tools for phase-transition studies. Phase Transit. 76, 155–170 (2003).
26
T. Sun, D. B. Zhang, R. M. Wentzcovitch, Dynamic stabilization of cubic CaSiO3 perovskite at high temperatures and pressures from ab initio molecular dynamics. Phys. Rev. B 89, 094109 (2014).
27
W. R. Taylor, A. L. Jaques, M. Ridd, Nitrogen-defect aggregation characteristics of some Australasian diamonds: Time-temperature constraints on the source regions of pipe and alluvial diamonds. Am. Mineral. 75, 1290–1310 (1990).
28
S. R. Boyd, I. Kiflawi, G. S. Woods, Infrared absorption by the B nitrogen aggregate in diamond. Philos. Mag. B 72, 351–361 (1995).
29
S. R. Boyd, I. Kiflawi, G. S. Woods, The relationship between infrared absorption and the A defect concentration in diamond. Philos. Mag. B 69, 1149–1153 (1994).
30
C. D. Clark, S. T. Davey, One-phonon infrared absorption in diamond. J. Phys. Chem. 17, 1127–1140 (1984).
31
K. P. Jochum, M. Willbold, I. Raczek, B. Stoll, K. Herwig, Chemical characterisation of the USGS reference glasses GSA-1G, GSC-1G, GSD-1G, GSE-1G, BCR-2G, BHVO-2G and BIR-1G using EPMA, ID-TIMS, ID-ICP-MS and LA-ICP-MS. Geostand. Newsl. 29, 285–302 (2005).
32
M. Humayun, F. A. Davis, M. M. Hirschmann, Major element analysis of natural silicates by laser ablation ICP- MS. J. Anal. At. Spectrom. 25, 998–1005 (2010).
33
D. Fliegel, C. Frei, G. Fontaine, Z. Hu, S. Gao, D. Günther, Sensitivity improvement in laser ablation inductively coupled plasma mass spectrometry achieved using a methane/argon and methanol/water/argon mixed gas plasma. Analyst 136, 4925–4934 (2011).
34
W. Kraus, G. Nolze, PowderCell – a program for the representation and manipulation of crystal structures and calculation of the resulting X-ray powder patterns. J. Appl. Cryst. 29, 301–303 (1996).
35
O. Narygina, L. S. Dubrovinsky, N. Miyajima, C. A. McCammon, I. Yu. Kantor, M. Mezouar, V. B. Prakapenka, N. A. Dubrovinskaia, V. Dmitriev, Phase relations in Fe-Ni-C system at high pressures and temperatures. Phys. Chem. Miner. 38, 203–214 (2011).
36
C. Prescher, V. B. Prakapenka, DIOPTAS: A program for reduction of two-dimensional X-ray diffraction data and data exploration. High Press. Res. 35, 223–230 (2015).
37
A. C. Larson, R. B. von Dreele, General Structure Analysis System (GSAS), Los Alamos National Laboratory Report LAUR 86-748 (2004).
38
H. Putz, J. C. Schön, M. Jansen, Combined method for ab initio structure solution from powder diffraction data. J. Appl. Cryst. 32, 864–870 (1999).
39
T. Kurashina, K. Hirose, S. Ono, N. Sata, Y. Ohishi, Phase transition in Al-bearing CaSiO3 perovskite: Implications for seismic discontinuities in the lower mantle. Phys. Earth Planet. Inter. 145, 67–74 (2004).
40
R. J. Angel, M. L. Mazzucchelli, M. Alvaro, P. Nimis, F. Nestola, Geobarometry from host-inclusion systems: The role of elastic relaxation. Am. Mineral. 99, 2146–2149 (2014).

(0)eLetters

eLetters is a forum for ongoing peer review. eLetters are not edited, proofread, or indexed, but they are screened. eLetters should provide substantive and scholarly commentary on the article. Embedded figures cannot be submitted, and we discourage the use of figures within eLetters in general. If a figure is essential, please include a link to the figure within the text of the eLetter. Please read our Terms of Service before submitting an eLetter.

Log In to Submit a Response

No eLetters have been published for this article yet.

Information & Authors

Information

Published In

Science
Volume 374 | Issue 6569
12 November 2021

Submission history

Received: 9 August 2021
Accepted: 21 September 2021
Published in print: 12 November 2021

Permissions

Request permissions for this article.

Acknowledgments

We thank N. Tomioka and an anonymous reviewer for their helpful comments.
Funding: This work was supported by awards NSF-EAR-1838330, -EAR-1942042, and -EAR-1322082; the NSF Cooperative Agreement No. DMR-1644779; and the State of Florida. Use of the Advanced Photon Source and the Advanced Light Source were supported by the US Department of Energy, Basic Energy Sciences, contracts DE-AC02-06CH11357 and DE-AC02-05CH11231, respectively.
Author contributions: O.T., S.H., S.Y., and M.H. participated in design, interpretation, data collection, and analysis of the reported results and in drafting and revising the manuscript. W.L., S.N.G.C., H.A.B., J.T., and G.R.R. participated in data collection and revising the manuscript.
Competing interests: The authors have no competing interests.
Data and materials availability: Additional chemical and crystallographic information about davemaoite is provided in the supplementary materials. Raw data are deposited at Dryad (23). Crystallographic and chemical information on type davemaoite is deposited in the Inorganic Crystal Structure Database. The type material is deposited with the NHMLA under accession number 74541.

Authors

Affiliations

Department of Geoscience, University of Nevada, Las Vegas, NV 89154, USA.
Roles: Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Resources, Supervision, Validation, Visualization, and Writing - original draft.
Department of Geoscience, University of Nevada, Las Vegas, NV 89154, USA.
Roles: Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing - original draft, and Writing - review & editing.
National High Magnetic Field Laboratory and Department of Earth, Ocean, and Atmospheric Science, Florida State University, Tallahassee, FL 32310, USA.
Role: Investigation.
National High Magnetic Field Laboratory and Department of Earth, Ocean, and Atmospheric Science, Florida State University, Tallahassee, FL 32310, USA.
Roles: Investigation, Methodology, Project administration, Supervision, Writing - original draft, and Writing - review & editing.
Wenjun Liu
Advanced Photon Source, Argonne National Laboratory, Lemont, IL 60439, USA.
Roles: Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Resources, Software, and Validation.
Stephanie N. Gilbert Corder https://orcid.org/0000-0002-5041-8670
Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA.
Roles: Data curation, Investigation, Project administration, Resources, Supervision, and Writing - review & editing.
Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA.
Roles: Investigation, Methodology, Resources, Validation, and Writing - review & editing.
Advanced Photon Source, Argonne National Laboratory, Lemont, IL 60439, USA.
Roles: Formal analysis, Methodology, Software, and Writing - review & editing.
Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91105, USA.
Roles: Data curation, Resources, and Visualization.

Funding Information

National Science Foundation: NSF-EAR-1838330
U.S. Department of Energy: DE-AC02-06CH11357
U.S. Department of Energy: DE-AC02-05CH11231
U.S. National Science Foundation: DMR-1644779
This: NSF-EAR-1838330
This: EAR-1942042

Notes

*
Corresponding author. Email: [email protected]

Metrics & Citations

Metrics

Article Usage

Altmetrics

Citations

Cite as

Export citation

Select the format you want to export the citation of this publication.

Cited by

  1. Davemaoite as the mantle mineral with the highest melting temperature, Science Advances, 9, 49, (2023)./doi/10.1126/sciadv.adj2660
    Abstract
  2. Imaging of the electronic bonding of diamond at pressures up to 2 million atmospheres, Science Advances, 9, 20, (2023)./doi/10.1126/sciadv.adg4159
    Abstract
  3. Equation of State and Spin Crossover of (Al, Fe)‐Phase H, Journal of Geophysical Research: Solid Earth, 128, 4, (2023).https://doi.org/10.1029/2022JB026291
    Crossref
  4. Na-bearing bridgmanite: Synthesis, phase relations and application to the origin of alkaline melts in the uppermost lower mantle, Lithos, 444-445, (107101), (2023).https://doi.org/10.1016/j.lithos.2023.107101
    Crossref
  5. Change in carbonate budget and composition during subduction below metal saturation boundary, Geoscience Frontiers, 14, 1, (101463), (2023).https://doi.org/10.1016/j.gsf.2022.101463
    Crossref
  6. Transformation microstructures in pyrolite under stress: Implications for anisotropy in subducting slabs below the 660 km discontinuity, Earth and Planetary Science Letters, 604, (118015), (2023).https://doi.org/10.1016/j.epsl.2023.118015
    Crossref
  7. Equation-of-state of magmatic liquids, Mass Transport in Magmatic Systems, (755-820), (2023).https://doi.org/10.1016/B978-0-12-821201-1.00011-0
    Crossref
  8. Wie innerer Druck die Erde formt, Physik in unserer Zeit, 54, 1, (37-43), (2023).https://doi.org/10.1002/piuz.202201655
    Crossref
  9. Mimicking Photosynthesis: A Natural Z‐Scheme Photocatalyst Constructed from Red Mud Bauxite Waste for Overall Water Splitting, Angewandte Chemie International Edition, (2023).https://doi.org/10.1002/anie.202302050
    Crossref
  10. Mimicking Photosynthesis: A Natural Z‐Scheme Photocatalyst Constructed from Red Mud Bauxite Waste for Overall Water Splitting, Angewandte Chemie, (2023).https://doi.org/10.1002/ange.202302050
    Crossref
  11. See more
Loading...

View Options

View options

PDF format

Download this article as a PDF file

Download PDF

Check Access

Log in to view the full text

AAAS ID LOGIN

AAAS login provides access to Science for AAAS Members, and access to other journals in the Science family to users who have purchased individual subscriptions.

Log in via OpenAthens.
Log in via Shibboleth.

More options

Register for free to read this article

As a service to the community, this article is available for free. Login or register for free to read this article.

Purchase this issue in print

Buy a single issue of Science for just $15 USD.

Media

Figures

Multimedia

Tables

Share

Share

Share article link

Share on social media