Next Article in Journal
Using Calcined Marls as Non-Common Supplementary Cementitious Materials—A Critical Review
Next Article in Special Issue
Compressibility and Phase Stability of Iron-Rich Ankerite
Previous Article in Journal
Study on Sintering Characteristics of Ultra-Poor Vanadium-Titanium Magnetite
Previous Article in Special Issue
Composition and Pressure Effects on Partitioning of Ferrous Iron in Iron-Rich Lower Mantle Heterogeneities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Compressibility of Novel Nickel Carbide at Pressures of Earth’s Outer Core

by
Timofey Fedotenko
1,*,
Saiana Khandarkhaeva
1,
Leonid Dubrovinsky
2,
Konstantin Glazyrin
3,
Pavel Sedmak
4 and
Natalia Dubrovinskaia
1,5
1
Material Physics and Technology at Extreme Conditions, Laboratory of Crystallography, University of Bayreuth, D-95440 Bayreuth, Germany
2
Bayerisches Geoinstitut, University of Bayreuth, D-95440 Bayreuth, Germany
3
Deutsches Elektronen Synchrotron, 22607 Hamburg, Germany
4
European Synchrotron Radiation Facility, F-38043 Grenoble, France
5
Department of Physics, Chemistry and Biology (IFM), Linköping University, SE-581 83 Linköping, Sweden
*
Author to whom correspondence should be addressed.
Minerals 2021, 11(5), 516; https://doi.org/10.3390/min11050516
Submission received: 16 April 2021 / Revised: 8 May 2021 / Accepted: 10 May 2021 / Published: 13 May 2021
(This article belongs to the Special Issue Minerals under Extreme Conditions)

Abstract

:
We report the high-pressure synthesis and the equation of state (EOS) of a novel nickel carbide (Ni3C). It was synthesized in a diamond anvil cell at 184(5) GPa through a direct reaction of a nickel powder with carbon from the diamond anvils upon heating at 3500 (200) K. Ni3C has the cementite-type structure (Pnma space group, a = 4.519(2) Å, b = 5.801(2) Å, c = 4.009(3) Å), which was solved and refined based on in-situ synchrotron single-crystal X-ray diffraction. The pressure-volume data of Ni3C was obtained on decompression at room temperature and fitted to the 3rd order Burch-Murnaghan equation of state with the following parameters: V0 = 147.7(8) Å3, K0 = 157(10) GPa, and K0′ = 7.8(6). Our results contribute to the understanding of the phase composition and properties of Earth’s outer core.

1. Introduction

Nickel is known as the second most abundant element in Earth’s core after iron [1,2]. Cosmochemical models and studies of meteorites suggest that Earth’s core apart from Fe contains also about 5 wt.% of Ni [3,4] and, in the inner core, up to 10 wt.% of light elements [5,6,7]. Which elements exactly and their amount is a subject of active discussions [3]. A large amount of carbon in iron meteorites [8], its high solubility in liquid Fe at high pressure [5,9], and high abundance in the solar system [5] suggest carbon to be one of the most important light elements in Earth’s core. Recent estimations of the inner core composition indicate up to 2.0 wt.% of carbon [3]. All these facts resulted in numerous high-pressure studies of the Fe-C system in recent decades. The intermediate Fe-C compounds Fe3C and Fe7C3 were suggested to be the most likely candidates to the carbon-bearing phases in Earth’s core, as they were found at relevant pressures and temperatures [2,5,10,11,12]. Although at room temperature Fe3C was shown to be stable up to 187 GPa, it decomposes into a mixture of solid Fe7C3 and hcp-Fe at above 145 GPa upon laser heating and transforms into Fe-C liquid and solid Fe7C3 at temperatures of above 3400 K [13]. Moreover, the high Poisson’s ratio of Fe7C3 at high pressures [2] indicates that the presence of carbon may significantly affect the elastic properties of iron. This corroborates the Preliminary Reference Earth Model (PREM) [14], which suggests the material of Earth’s inner core also has a high Poisson’s ratio.
Contrary to the binary iron-carbon system, the Fe–Ni–C, and Ni–C systems at high PT conditions are still poorly understood. Nickel can strongly modify the physical properties of pure Fe at elevated pressures and temperatures. Recent studies have shown that Ni alloying on Fe does not affect the melting temperature of Fe up to 100 GPa; however, it modifies its phase boundary by shifting the hcp/fcc/liquid triple point to the higher pressure-temperature region [6]. For example, for Fe-20 wt.% Ni alloy the triple point was found to be at 170(20) GPa and 4000(400) K [6] as compared to 100(10) GPa and 3500(200) K for pure Fe [15]. Pressure-induced Invar effect in Fe-Ni alloys was reported by Dubrovinsky et al. [16]. The thermal expansion of the alloys Fe0.55Ni0.45 and Fe0.20Ni0.80 was found to be extremely low in the temperature interval of 291 K to 500 K at pressures of 7.7 and 12.6 GPa, correspondingly [16]. It was also proven that alloys of Fe with Ni have significantly higher strength in comparison with pure Fe [17]. The mineral cohenite, (Fe, Ni)3C, which is isostructural to Fe3C, was found in iron meteorites [18] and predicted to be stable at high pressures [19]. However, a pure-Ni cementite-type phase (Ni3C) has never been reported before.
Here, we report the synthesis and EOS of a novel high-pressure phase of nickel carbide (Ni3C) in a laser-heated diamond anvil cell (LHDAC) at 184(5) GPa and 3500(200) K which was solved and refined using in-situ synchrotron single-crystal X-ray diffraction.

2. Materials and Methods

In our experiments, we used the BX90-type large X-ray aperture Diamond Anvil Cell (DAC) equipped with Boehler–Almax type diamonds with 80 µm culet diameter. To form the sample chamber, a rhenium gasket was preindented to ~20 µm thickness and a hole of 40 µm in diameter was drilled at the center of the indentation. A nickel powder was clamped between two thin layers of LiF inside the DAC’s sample chamber. LiF played a role of a pressure transmitting and thermal insulating medium in order to decrease temperature gradients in the sample during laser heating [20]. The pressure was determined using the equations of states (EOSes) of Ni [21] and monitored additionally using Raman signal from the diamond anvils [22].
The laser-heating (LH) of the samples was performed using in house laser heating setup [23]. The double-sided LH system is equipped with two YAG lasers (1064 nm central wavelength) and the IsoPlane SCT 320 spectrometer with a 1024 × 2560 PI-MAX 4 camera for the collection of thermal emission spectra from the heated spot. Temperatures were determined by fitting of thermal emission spectra of the sample to the grey body approximation of Planck’s radiation function in a given wavelength range (570–830 nm).
High-pressure single-crystal and powder X-ray diffraction (SCXRD) experiments were carried out at the extreme conditions beamline P02.2 (DESY, Hamburg, Germany) [24] and material science beamline ID11 (ESRF, Grenoble, France). The following beamline setups were used: At P02.2, λ = 0.29 Å, the beam size ~2 × 2 μm2, a Perkin Elmer XRD 1621 detector; at ID11, λ = 0.30996 Å, the beam size ~0.5 × 0.5 μm2, a Frelon4M detector. Single-crystal XRD data were collected during rotation of the DAC around the vertical ω-axis in a range ±35°. The diffraction images were collected with an exposure time of 5 s per frame with an angular step Δω = 0.5°.
To analyze the SCXRD data we used the CrysAlisPro software [25]. The analysis procedure includes the peak search, finding reflections belonging to a unique single-crystal domain, indexing, and data integration. The crystal structures were solved using ShelXT [26] structure solution program and refined with the JANA 2006 software [27].
Powder diffraction measurements were performed either without or upon continuous sample rotation about the ω axis of a diffractometer in the range of ±20°. The images were integrated into powder patterns with Dioptas software [28] and analyzed with Le Bail fitting technique using TOPAS 4.2. The parameters of the equation of state were obtained by fitting the pressure–volume data using EoSFit7-GUI software [29].

3. Results and Discussion

Sample of Ni powder was pressurized in LiF pressure-transmitting medium up to 184(5) GPa and laser-heated up to 3500 (200) K by scanning of the Ni sample with a laser beam. A direct reaction between Ni and carbon from the diamond anvil resulted in the synthesis of a new compound indexed as orthorhombic (Figure 1).
In order to localize the point of interest, high-resolution two-dimensional X-ray diffraction mapping (40 × 40 steps of 1 µm each) through the whole sample was realized at the ID11 beamline at the ESRF (Figure 2).
The reaction products consist of numerous single-crystalline grains that were identified using synchrotron single-crystal XRD. For one of such grains (one crystallite domain), we were able to collect 182 independent reflections and reduce the data with Rint = 7.3% at 184(5) GPa. The structure solution and refinement (final R1 = 6.4%, see Table 1) revealed the cementite-type orthorhombic structure (space group Pnma, #62; a = 4.520(3) Å, b = 5.8014(17) Å, c = 4.009(4) Å at 184(5) GPa) and the Ni3C chemical composition (Table 1, Supplementary Material, Crystallographic Information File: Ni3C_184GPa.cif).
The structure can be described as built of distorted trigonal prisms formed by six nickel atoms coordinating a C atom (Figure 3). The Ni-C distances in the prism vary from 1.760(19) to 1.830(16) Å at 184(5) GPa. The trigonal prisms, interconnected through sharing edges and corners, form layers parallel to the ac plane stacking along the b direction. The previously observed trigonal Ni3C (R-3c space group), which is a product of the thermal decomposition of Ni succinate [30] is built of CNi6 octahedra with an average Ni-C distance of 1.86 Å. Thus, the average Ni-C distance depends on the coordination of C atoms. Our data suggest that at ambient pressure the average Ni-C distance in CNi6 trigonal prisms should be significantly lager compared to that inCNi6 octahedra.
The Ni3C sample was studied on a stepwise decompression. SCXRD data were collected at seven pressure points down to 84(2) GPa. Below 84(2) GPa no diffraction pattern from Ni3C was observed; however, the reason remained unclear. That means the question as to if the quality of the sample deteriorated or the phase decomposed or amorphized stays open. The pressure-volume data (Table 2) of Ni3C was fitted to the 3rd order Birch-Murnaghan (BM3) EOS and gave the following parameters: V0 = 147.7(8) Å3; K0 = 157(10) GPa, K′ = 7.8(6) (Figure 4).
Figure 5 demonstrates experimental data on Ni3C axial compression. The structure is most compressible along the b axis, the direction of stacking of the layers of interconnected CNi6 trigonal prisms. Compared to the predicted compressional behavior ofFe3C in the same pressure region [34], Ni3C is highly anisotropic.
Based on obtained data, we calculated the bulk sound velocity for Ni3C as a function of pressure at 293 K and compared it with those known for Fe, Ni, and possible carbon-bearing components of Earth’s core (Fe3C and Fe7C3). Figure 6 shows that within the errors Ni3C exhibits similar bulk velocities as Fe3C and Fe7C3 at pressures up to 400 GPa.
Thereby, the presence of Ni in the alloy likely should not affect the elastic properties of the Fe-Ni-C system at high pressure but potentially can change the carbon distribution. Due to the stability of Ni3C at conditions of Earth’s outer core, it may be considered as one of the likely candidates to carbon-bearing phases in the core along with Fe7C3.

4. Conclusions

In the presented work, we have synthesized a nickel carbide yet unknown at ambient conditions. It was shown that Ni reacts with carbon at high-pressure and high-temperature conditions that result in the formation of an orthorhombic Ni3C compound (space group Pnma, a = 4.520(3) Å, b = 5.8014(17) Å, c = 4.009(4) Å at 84(5) GPa) with the cementite-type structure revealed using synchrotron single-crystal X-ray diffraction. The Ni3C was studied on decompression down to 84(2) GPa. We have shown that in the pressure range 84(2)–185(5) GPa, Ni3C is less compressible than cementite (Fe3C); the calculated bulk sound velocities are similar to those known for Fe3C and Fe7C3 at pressures up to 400 GPa and 297 K. Ni3C remains stable at pressure-temperature conditions relevant to Earth’s core and thus can be considered as one of the likely candidates to carbon-bearing phases in the core along with Fe7C3.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/min11050516/s1, Crystallographic information file of Ni3C at 184(5) GPa.

Author Contributions

Conceptualization, T.F., L.D. and N.D.; methodology, T.F. and S.K.; validation, T.F. and S.K.; formal analysis, T.F.; investigation, T.F., S.K., K.G., P.S.; data curation, T.F.; writing—original draft preparation, T.F.; writing—review and editing, N.D. and L.D.; visualization, T.F.; supervision, N.D. and L.D.; project administration, T.F. and L.D.; funding acquisition, N.D. and L.D. All authors have read and agreed to the published version of the manuscript.

Funding

N.D. and L.D. thank the Federal Ministry of Education and Research, Germany (BMBF, grant No. 05K19WC1) and the Deutsche Forschungsgemeinschaft (DFG projects DU 954-11/1, DU 393-9/2, and DU 393-13/1) for financial support. N.D. thanks the Swedish Government Strategic Research Area in Materials Science on Functional Materials at Linköping University (Faculty Grant SFO-Mat-LiU No. 2009 00971).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

We acknowledge DESY (Hamburg, Germany), a member of the Helmholtz Association HGF, for the provision of experimental facilities, photon beamline P02.2 (Petra III).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Birch, F. Elasticity and constitution of the Earth’s interior. J. Geophys. Res. 1952, 57, 227–286. [Google Scholar] [CrossRef]
  2. Prescher, C.; Dubrovinsky, L.; Bykova, E.; Kupenko, I.; Glazyrin, K.; Kantor, A.; McCammon, C.; Mookherjee, M.; Nakajima, Y.; Miyajima, N.; et al. High Poisson’s ratio of Earth’s inner core explained by carbon alloying. Nat. Geosci. 2015, 8, 220–223. [Google Scholar] [CrossRef]
  3. Litasov, K.D.; Shatskiy, A.F. Composition of the Earth’s core: A review. Russ. Geol. Geophys. 2016, 57, 22–46. [Google Scholar] [CrossRef]
  4. McDonough, W.F. Compositional Model for the Earth’s Core. In Treatise on Geochemistry; Elsevier: Amsterdam, The Netherlands, 2003; pp. 547–568. [Google Scholar]
  5. Wood, B.J. Carbon in the core. Earth Planet. Sci. Lett. 1993, 117, 593–607. [Google Scholar] [CrossRef]
  6. Torchio, R.; Boccato, S.; Miozzi, F.; Rosa, A.D.; Ishimatsu, N.; Kantor, I.; Sévelin-Radiguet, N.; Briggs, R.; Meneghini, C.; Irifune, T.; et al. Melting Curve and Phase Relations of Fe-Ni Alloys: Implications for the Earth’s Core Composition. Geophys. Res. Lett. 2020, 47, e2020GL088169. [Google Scholar] [CrossRef]
  7. Poirier, J.-P. Light elements in the Earth’s outer core: A critical review. Phys. Earth Planet. Inter. 1994, 85, 319–337. [Google Scholar] [CrossRef]
  8. Bashir, N.; Beckett, J.R.; Hutcheon, I.D.; Stolper, E.M. Carbon in the Metal of Iron Meteorites. In Proceedings of the Lunar and Planetary Science Conference, Houston, TX, USA, 18–22 March 1996; Volume 27, p. 63. [Google Scholar]
  9. Hirayama, Y.; Fujii, T.; Kurita, K. The melting relation of the system, iron and carbon at high pressure and its bearing on the early stage of the Earth. Geophys. Res. Lett. 1993, 20, 2095–2098. [Google Scholar] [CrossRef]
  10. Lord, O.T.; Walter, M.J.; Dasgupta, R.; Walker, D.; Clark, S.M. Melting in the Fe–C system to 70 GPa. Earth Planet. Sci. Lett. 2009, 284, 157–167. [Google Scholar] [CrossRef] [Green Version]
  11. Nakajima, Y.; Takahashi, E.; Suzuki, T.; Funakoshi, K. “Carbon in the core” revisited. Phys. Earth Planet. Inter. 2009, 174, 202–211. [Google Scholar] [CrossRef]
  12. Chen, B.; Li, Z.; Zhang, D.; Liu, J.; Hu, M.Y.; Zhao, J.; Bi, W.; Alp, E.E.; Xiao, Y.; Chow, P.; et al. Hidden carbon in Earth’s inner core revealed by shear softening in dense Fe7C3. Proc. Natl. Acad. Sci. USA 2014, 111, 17755–17758. [Google Scholar] [CrossRef] [Green Version]
  13. Liu, J.; Lin, J.-F.; Prakapenka, V.B.; Prescher, C.; Yoshino, T. Phase relations of Fe3C and Fe7C3 up to 185 GPa and 5200 K: Implication for the stability of iron carbide in the Earth’s core. Geophys. Res. Lett. 2016, 43, 12415–12422. [Google Scholar] [CrossRef]
  14. Dziewonski, A.M.; Anderson, D.L. Preliminary reference Earth model. Phys. Earth Planet. Inter. 1981, 25, 297–356. [Google Scholar] [CrossRef]
  15. Morard, G.; Boccato, S.; Rosa, A.D.; Anzellini, S.; Miozzi, F.; Henry, L.; Garbarino, G.; Mezouar, M.; Harmand, M.; Guyot, F.; et al. Solving Controversies on the Iron Phase Diagram Under High Pressure. Geophys. Res. Lett. 2018, 45, 11-074. [Google Scholar] [CrossRef] [Green Version]
  16. Dubrovinsky, L.; Dubrovinskaia, N.; Abrikosov, I.A.; Vennström, M.; Westman, F.; Carlson, S.; van Schilfgaarde, M.; Johansson, B. Pressure-Induced Invar Effect in Fe-Ni Alloys. Phys. Rev. Lett. 2001, 86, 4851–4854. [Google Scholar] [CrossRef] [PubMed]
  17. Reagan, M.M.; Gleason, A.E.; Liu, J.; Krawczynski, M.J.; Van Orman, J.A.; Mao, W.L. The effect of nickel on the strength of iron nickel alloys: Implications for the Earth’s inner core. Phys. Earth Planet. Inter. 2018, 283, 43–47. [Google Scholar] [CrossRef]
  18. Brett, R. Cohenite in Meteorites: A Proposed Origin. Science 1966, 153, 60–62. [Google Scholar] [CrossRef]
  19. Ringwood, A.E. Cohenite as a pressure indicator in iron meteorites. Geochim. Cosmochim. Acta 1960, 20, 155–158. [Google Scholar] [CrossRef]
  20. Dong, H.; Dorfman, S.M.; Holl, C.M.; Meng, Y.; Prakapenka, V.B.; He, D.; Duffy, T.S. Compression of lithium fluoride to 92 GPa. High Press. Res. 2014, 34, 39–48. [Google Scholar] [CrossRef]
  21. Dewaele, A.; Torrent, M.; Loubeyre, P.; Mezouar, M. Compression curves of transition metals in the Mbar range: Experiments and projector augmented-wave calculations. Phys. Rev. B 2008, 78, 104102. [Google Scholar] [CrossRef]
  22. Akahama, Y.; Kawamura, H. Pressure calibration of diamond anvil Raman gauge to 310GPa. J. Appl. Phys. 2006, 100, 043516. [Google Scholar] [CrossRef]
  23. Fedotenko, T.; Dubrovinsky, L.; Aprilis, G.; Koemets, E.; Snigirev, A.; Snigireva, I.; Barannikov, A.; Ershov, P.; Cova, F.; Hanfland, M.; et al. Laser heating setup for diamond anvil cells for in situ synchrotron and in house high and ultra-high pressure studies. Rev. Sci. Instrum. 2019, 90, 104501. [Google Scholar] [CrossRef]
  24. Liermann, H.-P.; Konôpková, Z.; Morgenroth, W.; Glazyrin, K.; Bednarčik, J.; McBride, E.E.; Petitgirard, S.; Delitz, J.T.; Wendt, M.; Bican, Y.; et al. The Extreme Conditions Beamline P02.2 and the Extreme Conditions Science Infrastructure at PETRA III. J. Synchrotron Radiat. 2015, 22, 908–924. [Google Scholar] [CrossRef] [Green Version]
  25. CrysAlis Pro (v. 171.39.46). Available online: https://www.rigaku.com/products/smc/crysalis (accessed on 3 January 2020).
  26. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Petříček, V.; Dušek, M.; Plášil, J. Crystallographic computing system Jana2006: Solution and refinement of twinned structures. Zeitschrift für Krist. Cryst. Mater. 2016, 231, 583–599. [Google Scholar] [CrossRef]
  28. Prescher, C.; Prakapenka, V.B. DIOPTAS: A program for reduction of two-dimensional X-ray diffraction data and data exploration. High Press. Res. 2015, 35, 223–230. [Google Scholar] [CrossRef]
  29. Gonzalez-Platas, J.; Alvaro, M.; Nestola, F.; Angel, R. EosFit7-GUI: A new graphical user interface for equation of state calculations, analyses and teaching. J. Appl. Crystallogr. 2016, 49, 1377–1382. [Google Scholar] [CrossRef]
  30. Huba, Z.J.; Carpenter, E.E. Monitoring the formation of carbide crystal phases during the thermal decomposition of 3d transition metal dicarboxylate complexes. Dalt. Trans. 2014, 43, 12236–12242. [Google Scholar] [CrossRef]
  31. Li, J.; Mao, H.K.; Fei, Y.; Gregoryanz, E.; Eremets, M.; Zha, C.S. Compression of Fe 3 C to 30 GPa at room temperature. Phys. Chem. Miner. 2002, 29, 166–169. [Google Scholar] [CrossRef]
  32. Prescher, C.; Dubrovinsky, L.; McCammon, C.; Glazyrin, K.; Nakajima, Y.; Kantor, A.; Merlini, M.; Hanfland, M. Structurally hidden magnetic transitions in Fe3C at high pressures. Phys. Rev. B 2012, 85, 140402. [Google Scholar] [CrossRef] [Green Version]
  33. Scott, H.P.; Williams, Q.; Knittle, E. Stability and equation of state of Fe 3 C to 73 GPa: Implications for carbon in the Earth’s core. Geophys. Res. Lett. 2001, 28, 1875–1878. [Google Scholar] [CrossRef]
  34. Mookherjee, M. Elasticity and anisotropy of Fe3C at high pressures. Am. Miner. 2011, 96, 1530–1536. [Google Scholar] [CrossRef]
  35. Fei, Y.; Murphy, C.; Shibazaki, Y.; Shahar, A.; Huang, H. Thermal equation of state of hcp-iron: Constraint on the density deficit of Earth’s solid inner core. Geophys. Res. Lett. 2016, 43, 6837–6843. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Reconstructed reciprocal lattice planes of orthorhombic Ni3C compound with the cementite-type structure. Reflections highlighted by red circles correspond to diamond, diffraction rings correspond to Re, Ni, and LiF.
Figure 1. Reconstructed reciprocal lattice planes of orthorhombic Ni3C compound with the cementite-type structure. Reflections highlighted by red circles correspond to diamond, diffraction rings correspond to Re, Ni, and LiF.
Minerals 11 00516 g001
Figure 2. (a) Two-dimensional X-Ray diffraction mapping of the sample chamber. The color map allowing to distinguish between the present phases. The intensity of the colors is proportional to the intensity of particular reflections: the dark purple region beyond the pressure chamber corresponds to the (100) and (101) reflections of the Re gasket; the blue region—the (200) reflection of the Ni; The orange region—the (111) reflection of LiF; (020) and (301) reflection of Ni3C for the red region. (b) A comparison view of the sample chamber under an optical microscope. (c) Powder diffraction pattern the temperature quenched sample at 184(5) GPa at the position highlighted by a black dotted square on the XRD color map.
Figure 2. (a) Two-dimensional X-Ray diffraction mapping of the sample chamber. The color map allowing to distinguish between the present phases. The intensity of the colors is proportional to the intensity of particular reflections: the dark purple region beyond the pressure chamber corresponds to the (100) and (101) reflections of the Re gasket; the blue region—the (200) reflection of the Ni; The orange region—the (111) reflection of LiF; (020) and (301) reflection of Ni3C for the red region. (b) A comparison view of the sample chamber under an optical microscope. (c) Powder diffraction pattern the temperature quenched sample at 184(5) GPa at the position highlighted by a black dotted square on the XRD color map.
Minerals 11 00516 g002
Figure 3. Crystal structure of the cementite type Ni3C at 184(5) GPa and room temperature. Purple and black spheres designate nickel and carbon atoms, correspondingly.
Figure 3. Crystal structure of the cementite type Ni3C at 184(5) GPa and room temperature. Purple and black spheres designate nickel and carbon atoms, correspondingly.
Minerals 11 00516 g003
Figure 4. The pressure-volume dependence of Ni3C. Red dots represent experimental data, the dashed red curve is the BM3 EOS fit (V0 = 147.7(8) Å3; K0 = 157(10) GPa, K′ = 7.8(6)). Solid purple, blue and green lines correspond to the EOSes of Fe3C from studies of Li et al. (K0 = 174(6) GPa, K′ = 4.8(8)) [31], Prescher et al. (K0 = 161(2) GPa, K′ = 5.9(2)) [32] and Scott et al. (K0 = 165(4) GPa, K′ = 5.99(9)) [33].
Figure 4. The pressure-volume dependence of Ni3C. Red dots represent experimental data, the dashed red curve is the BM3 EOS fit (V0 = 147.7(8) Å3; K0 = 157(10) GPa, K′ = 7.8(6)). Solid purple, blue and green lines correspond to the EOSes of Fe3C from studies of Li et al. (K0 = 174(6) GPa, K′ = 4.8(8)) [31], Prescher et al. (K0 = 161(2) GPa, K′ = 5.9(2)) [32] and Scott et al. (K0 = 165(4) GPa, K′ = 5.99(9)) [33].
Minerals 11 00516 g004
Figure 5. The pressure dependence of the normalized unit cell parameters of Ni3C at 300 K.
Figure 5. The pressure dependence of the normalized unit cell parameters of Ni3C at 300 K.
Minerals 11 00516 g005
Figure 6. Calculated bulk sound velocity as a function of pressure for Ni3C (this study, black solid line with circles); Fe3C (green line with diamonds [33]) and Fe7C3 (blue line with squares [2]); Ni (red line with triangles [21]); Fe (purple line with pentagons [35]) at 293 K.
Figure 6. Calculated bulk sound velocity as a function of pressure for Ni3C (this study, black solid line with circles); Fe3C (green line with diamonds [33]) and Fe7C3 (blue line with squares [2]); Ni (red line with triangles [21]); Fe (purple line with pentagons [35]) at 293 K.
Minerals 11 00516 g006
Table 1. Crystallographic data for the Ni3C at 184(5) GPa and 293 K.
Table 1. Crystallographic data for the Ni3C at 184(5) GPa and 293 K.
Chemical FormulaNi3C
Crystal systemOrthorhombic
Space groupPnma
Pressure (GPa)184(5)
Temperature (K)293
a (Å)4.520(3)
b (Å)5.8014(17)
c (Å)4.009(4)
V3)105.12(13)
Z4
Density(g·cm−3)11.884
Radiation typesynchrotron, λ = 0.2895 Å
DiffractometerP02.2 @DESY
No. of measured, Independent and observed [I > 3σ(I)] reflections366, 182, 83
Rint7.3%
Refinement methodFull matrix least-squares on F
R [F > 3σ(F)], wR(F), S6.43, 8.42, 1.43
No. of parameters19
Δρmax, Δρmin(e·Å−3) 3.09, −3.51
Table 2. The pressure dependence of the unit cell parameter of Ni3C. Values in parentheses correspond to experimental uncertainties.
Table 2. The pressure dependence of the unit cell parameter of Ni3C. Values in parentheses correspond to experimental uncertainties.
Pressure, GPaVolume, Å3
84 (2)117.1 (6)
101 (2)114.7 (3)
123 (3)111.7 (4)
142 (3)108.9 (4)
160 (4)107.4 (4)
170 (4)106.3 (3)
184 (5)105.1 (2)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fedotenko, T.; Khandarkhaeva, S.; Dubrovinsky, L.; Glazyrin, K.; Sedmak, P.; Dubrovinskaia, N. Synthesis and Compressibility of Novel Nickel Carbide at Pressures of Earth’s Outer Core. Minerals 2021, 11, 516. https://doi.org/10.3390/min11050516

AMA Style

Fedotenko T, Khandarkhaeva S, Dubrovinsky L, Glazyrin K, Sedmak P, Dubrovinskaia N. Synthesis and Compressibility of Novel Nickel Carbide at Pressures of Earth’s Outer Core. Minerals. 2021; 11(5):516. https://doi.org/10.3390/min11050516

Chicago/Turabian Style

Fedotenko, Timofey, Saiana Khandarkhaeva, Leonid Dubrovinsky, Konstantin Glazyrin, Pavel Sedmak, and Natalia Dubrovinskaia. 2021. "Synthesis and Compressibility of Novel Nickel Carbide at Pressures of Earth’s Outer Core" Minerals 11, no. 5: 516. https://doi.org/10.3390/min11050516

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