Next Article in Journal
Strength and Volume Change Characteristics of Clayey Soils: Performance Evaluation of Enzymes
Previous Article in Journal
Physicochemical Parameters and Geochemical Features of Ore-Forming Fluids for Orogenic Gold Deposits Throughout Geological Time
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interaction Between FeOOH and NaCl at Extreme Conditions: Synthesis of Novel Na2FeCl4OHx Compound

1
Bayerisches Geoinstitut, Bayreuth University, 95447 Bayreuth, Germany
2
Department of Earth and Planetary Materials Science, Graduate School of Science, Tohoku University, Sendai 980-8578, Japan
3
Deutsches Elektronen-Synchrotron, 22607 Hamburg, Germany
*
Author to whom correspondence should be addressed.
Minerals 2020, 10(1), 51; https://doi.org/10.3390/min10010051
Submission received: 25 November 2019 / Revised: 19 December 2019 / Accepted: 30 December 2019 / Published: 5 January 2020

Abstract

:
Iron(III) oxide-hydroxide, FeOOH, is abundant in the banded iron formations (BIFs). Recent studies indicate that BIFs may carry water down to the lower mantle with subducting slabs. The previous experiments investigating the properties of FeOOH at extreme pressures (P) and temperatures (T) were performed in diamond anvil cells (DACs), where it was compressed inside alkali metal halide pressure-transmitting media (2). Alkali metal halides such as NaCl or KCl are expected to be chemically inert; therefore, they are widely used in DAC experiments. Here, we report the chemical interaction between FeOOH and NaCl pressure medium at 107(2) GPa and 2400(200) K. By means of single-crystal X-ray diffraction (SC-XRD) analysis applied to a multigrain sample, we demonstrate the formation of a Na2FeCl4OHx phase and provide its structural solution and refinement. Our results demonstrate that at high P-T conditions, the alkali metal halides could interact with hydrous phases and thus cannot be used as a pressure transmitting and thermal insulating medium in DAC experiments dedicated to studies of hydroxyl or water-bearing materials at high P-T.

1. Introduction

Hydrous phases have been attracting great interest in the geological community due to their importance for the behavior and physical properties of the Earth’s mantle [1,2]. One of such compounds, α-FeOOH (stabilized as goethite at the ambient conditions, which is a common material in the banded iron formations, or BIFs), is expected to submerge into the lower mantle with subducting slabs, bringing water to the lowermost mantle and the core–mantle boundary (CMB), as it was suggested in some publications [3,4]. Recently, a number of studies were dedicated to studying the behavior of FeOOH at high P-T conditions, which show that FeOOH undergoes a series of phase transitions. A “pyrite-type” FeO2Hx phase is expected to be stable at the P-T conditions of lowermost part of a mantle and core–mantle boundary [5,6,7]. This phase is attracting the attention of the community and is being intensively investigated via density functional theory calculations and spectroscopic studies [5,6,7,8,9]. In some experiments performed in DACs, alkali halides (such as KCl, for example, ref. [9]) are used as a pressure-transmitting medium and thermal insulator for laser heating. It is assumed that alkali halides are fully chemically inert and cannot complicate data analysis. However, at very high pressures and temperatures, the chemistry of materials may change, and iron oxyhydroxide could potentially interact with the alkali halides. Information on the possible reactions between FeOOH and alkali metal halides at extreme conditions and the crystal structures of possible products is absent. Methodologically, it is of great importance to verify the reactivity between these species.

2. Materials and Methods

A FeOOH sample (Alfa Aesar, CAS: 20344-49-4) with a lateral diameter of ~10 μm and thickness of ~5 μm was loaded inside a sample chamber of a BX90-type diamond anvil cell (DAC) [10] equipped with two beveled Boehler-Almax type diamonds (culet diameter 80 μm). The sample chamber was made by pre-indenting Re oil-free gasket (initial thickness 200 µm, thickness after indentation ~15 µm) and drilling a hole of ~35 μm diameter in it. A NaCl powder was used as both a pressure-transmitting medium and a reactant to check its possible interaction with FeOOH. The DAC was compressed to the desired pressure (the pressure was determined using the equation of state (EoS) of B2-structured NaCl [11]) and laser-heated at BL-10XU at SPring-8 (Sayo, Hyogo, Japan) [12]. The analysis of the reaction products was performed on the T-quenched sample at P2.02 DESY beamline [13] employing the single-crystal X-ray diffraction technique for laser-heated DACs as developed at Bayreuth University [14]. In order to locate the grain of a reaction product and collect the single-crystal X-ray diffraction (SC-XRD) dataset from it, the XRD mapping was performed by scanning the sample in vertical and horizontal directions with steps of three micrometers. After localizing grain(s) of interest, the step scans were recorded on a Perkin Elmer detector (XRD1621, Waltham, MA, USA) (sets of individual photographs (with equal exposure time, from 1 to 10 s) taken each 0.5° of DAC rotation in a range of ± 36° around the ω axis of the goniometer).
Analysis of collected diffraction images was performed using CrysAlis PRO software (Agilent Technologies, Version 1.171.37.35) [15]. The structures were solved using the dual-space method and refined in the isotropic approximation by full-matrix least-squares refinements using SHELX software package (SHELX-2019, Universität Göttingen, Göttingen, Germany) software [16,17]. Crystal structure visualization made with the VESTA (Version 3.4.4) software [18].

3. Results

The compression of FeOOH in NaCl at ambient temperature to 107(2) GPa did not result in any chemical reaction. At this pressure, after double-side laser heating of the sample at ~2400(200) K, new sets of reflections on the XRD patterns of the T-quenched sample were observed. Consideration of powder and single-crystal diffraction data collected from the laser-heated sample suggests that several different phases are formed in the DAC (Figure S1). From the powder-diffraction data (Figure 1, Figure S1), we identified a cubic FeO2Hx phase with a unit cell parameter 4.403(3) Å, which is in agreement with a recent literature reports [3,4], as well as a signal from a cubic B2-type NaCl (a = 2.911(2) Å) and hcp-structured Re (a = 2.6697(8) Å, c = 3.923(2) Å) from the gasket. Remaining peaks on the powder XRD pattern belong to an unknown phase. Through the further analysis of the single-crystal data of decent quality obtained for several grains, we were able to index reflections belonging to two different co-existing orthorhombic phases. For the first phase, we found unit cell parameters a = 2.5467(5) Å, b = 9.640(2) Å, and c = 11.580(2) Å, and a suggested space group Imm2 (#44) (further denoted as “oI-phase”) as determined by analysis of the 298 unique reflections and systematic absences. Unfortunately, the intensity of the Bragg reflections of this phase is too low, and the structure solution is not possible.
For the second new phase, analysis of single-crystal data allows determining the orthorhombic unit cell with parameters a = 8.725(2) Å, b = 6.180(3) Å, c = 3.0679(12) Å. Further integration of a dataset allowed extracting the intensities of 315 Bragg reflections, which belong to the single grain, and a space group Pbam (#55) was determined (Figure S2). The quality of the data (Table 1) allowed performing structure solution and refinement. We found that this phase with a chemical formula Na2FeCl4OHx is the product of the chemical reaction between FeOOH and NaCl. Its crystal structure is shown in Figure 2 (crystal data, information on the data collection and structure refinement details summarized in Table 1 and Table 2). We refined atomic displacement parameters in isotropic approximation due to the limited quality of the data collected in the DAC at pressure over megabar. Due to the same reason, positions of hydrogen atoms cannot be freely refined, and we fixed H-atoms at Wyckoff 8i position with a 1 Å distance to the oxygen atoms and site occupancy fixed at 0.25.
In the orthorhombic Na2FeCl4OHx compound, the iron atoms coordinated with two crystallographically distinct chlorine (Cl1 and Cl2) atoms, forming just a slightly distorted octahedra (distortion index = 0.00125) with an average bond length of 2.044 Å. These FeCl6 octahedra are stacked into the columns along the c direction, sharing the edges (Figure 2). The Na atoms are coordinated with seven chlorine atoms and two oxygen atoms, forming a monocapped antiprismatic polyhedron (Figure 2). Two oxygen and two chlorine atoms form a parallelogram base of antiprism (with an O1–Cl2–O1 angle of 74.42(8)°). These polyhedra are connected to each other through a common triangular and parallelogram faces and share the common edges with FeCl6 octahedra (Figure 2). The coordination of Na in Na2FeCl4OHx is unusual for the Na and resembles the one described for La in LaOCl oxyhalides [19].

4. Discussion

While the stoichiometry of heavier atoms in the Na2FeCl4OHx compound is uniquely defined by structural refinement, the amount of hydrogen is not known. One could assign common (usual) oxidation states for Na, Cl, and O atoms (+1, –1, and –2, respectively), but for iron as a transition metal, several oxidation states are known to range from 0 to +4. The trivial charge balance considerations could not be easily implemented: the refinement of hydrogen atoms’ positions and occupancies directly from the SC-XRD data is a complex task even at ambient conditions; for the experiments performed in the DACs, it is practically impossible in the most cases. Mössbauer or X-ray absorption near edge structure (XANES) spectroscopies could help to define the oxidation state of iron and therefore establish the amount of hydrogen in phase; however, in case of multiphase samples, the interpretation of spectroscopy data would be most probably ambiguous.
Another way would be to perform the crystallochemical analysis and compare the bond distances in Na2FeCl4OHx with Fe–Cl distances in other known compounds. Unfortunately, such a type of analysis is greatly complicated by very limited data on the high-P behavior of Fe,Cl-bearing compounds. By fitting the pressure–volume literature data on Fe2+Cl2 and Fe3+ClO with the Birch–Murnaghan equation of state (EoS), we described the Fe–Cl bond evolution in FeCl6 octahedra upon compression. Assuming that the structures of these compounds remain the same at high pressures, we compare extrapolated Fe–Cl bond lengths for Fe2+Cl2 and Fe3+OCl and our experimental values at 107 GPa (Figure 3). The observed average Fe–Cl bond length in FeCl6 octahedra in the Na2FeCl4OHx is 2.044 Å (see above). Extrapolated values for the Fe–Cl contact in FeOCl and FeCl2 are ~2.6 Å and ~2.0 Å at 107 GPa.
While the coordination of iron in FeOCl is not the same as in Na2FeCl4OHx and the extrapolated Fe–Cl bond length is much higher, the iron in FeCl2 is coordinated similarly, and the estimated Fe–Cl bonds of ~2 Å are very close to the observed Fe–Cl contact in our experiment. Therefore, one could suggest that iron in Na2FeCl4OHx is in the 2+ oxidation state, and therefore x = 2. Still, the iron oxidation state 3+ cannot be excluded: the reported values of Fe–Cl bond lengths for ferric iron in Fe3+Cl6 octahedra of FeCl3 and K2FeCl5(H2O) compounds are obviously much lower than those of ferrous iron chlorides at ambient conditions. Unfortunately, the lack of information on the compressibility of Fe3+ atoms octahedrally coordinated by Cl does not allow us to estimate the Fe–Cl distances at 107(2) GPa. Therefore, our crystallochemical considerations cannot confidently assign an oxidation state to the Fe atoms.

5. Conclusions

Our results demonstrate that at high P-T conditions, sodium halide reacts with FeOOH, forming a novel Na2FeCl4OHx compound. Thus, the alkali metal halides could interact with hydrous phases and thus cannot be used as a pressure-transmitting and thermal-insulating medium in DAC experiments dedicated to studies of hydroxyl or water-bearing materials at high pressure–high temperature consitions.

Supplementary Materials

The following are available online at https://www.mdpi.com/2075-163X/10/1/51/s1, Figure S1: Representative X-ray diffraction pattern collected after laser-heating of FeOOH in NaCl at 2000(200) K and 107(2) GPa (λ = 0.2892 Å, P02.2 beamline at PETRAIII). (a) As collected 2D XRD image acquired with Perkin-Elmer detector during rotation of the DAC from −20° to +20° around the ω axis (“wide scan”). (b) Rolled up (‘cake’) representation of a selected 2D wide scan XRD pattern. Figure S2: Examples of reconstructed reciprocal planes for one of the domains of Na2FeCl4OHx phase at 107(2) GPa. (a), (b), and (c) panels show (hk-1), (h1l), and (2kl) planes correspondingly.

Author Contributions

Conceptualization, E.K. and L.D.; methodology, E.K., E.B., K.G, L.D., E.O. and L.Y.; investigation, E.K., L.Y., E.B., K.G, E.O., and L.D.; resources, E.O. and L.D.; data curation, E.K., L.Y., E.B., E.O., and L.D.; writing—original draft preparation, E.K. and L.D.; writing—review and editing, E.K., L.Y., E.B., K.G., E.O. and L.D.; funding acquisition, E.K., E.O., L.Y., and L.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Deutsche Forschungsgemeinschaft (DFG), grant number GRK 2156/1.

Acknowledgments

Authors thank the Deutsches Elektronen-Synchrotron for the provision of synchrotron beamtimes (P02.2). Authors are grateful for the SPring-8 for provision of synchrotron beamtime at BL10XU.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Peslier, A.H.; Schönbächler, M.; Busemann, H.; Karato, S.I. Water in the Earth’s interior: Distribution and origin. Space Sci. Rev. 2017, 212, 1–68. [Google Scholar]
  2. Hirschmann, M.M. Water, melting, and the deep Earth H2O cycle. Annu. Rev. Earth Planet. Sci. 2006, 34, 629–653. [Google Scholar] [CrossRef] [Green Version]
  3. Hu, Q.; Kim, D.Y.; Yang, W.; Yang, L.; Meng, Y.; Zhang, L.; Mao, H.K. FeO2 and FeOOH under deep lower-mantle conditions and Earth’s oxygen-hydrogen cycles. Nature 2016, 534, 241–244. [Google Scholar] [CrossRef] [PubMed]
  4. Hu, Q.; Kim, D.Y.; Liu, J.; Meng, Y.; Yang, L.; Zhang, D.; Mao, W.L.; Mao, H. Dehydrogenation of goethite in Earth’s deep lower mantle. Proc. Natl. Acad. Sci. USA 2017, 114, 1498–1501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Mao, H.K.; Hu, Q.; Yang, L.; Liu, J.; Kim, D.Y.; Meng, Y.; Zhang, L.; Prakapenka, V.B.; Yang, W.; Mao, W.L. When water meets iron at Earth’s core-mantle boundary. Natl. Sci. Rev. 2017, 4, 870–878. [Google Scholar] [CrossRef]
  6. Liu, J.; Hu, Q.; Bi, W.; Yang, L.; Xiao, Y.; Chow, P.; Meng, Y.; Prakapenka, V.B.; Mao, H.K.; Mao, W.L. Altered chemistry of oxygen and iron under deep Earth conditions. Nat. Commun. 2019, 10, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Nishi, M.; Kuwayama, Y.; Tsuchiya, J.; Tsuchiya, T. The pyrite-Type high-pressure form of FeOOH. Nature 2017, 547, 205–208. [Google Scholar] [CrossRef] [PubMed]
  8. Streltsov, S.S.; Shorikov, A.O.; Skornyakov, S.L.; Poteryaev, A.I.; Khomskii, D.I. Unexpected 3+ valence of iron in FeO2, a geologically important material lying “in between” oxides and peroxides. Sci. Rep. 2017, 7, 13005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Boulard, E.; Harmand, M.; Guyot, F.; Lelong, G.; Morard, G.; Cabaret, D.; Boccato, S.; Rosa, A.D.; Briggs, R.; Pascarelli, S.; et al. Ferrous iron under oxygen-rich conditions in the deep mantle. Geophys. Res. Lett. 2019, 46, 1348–1356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Kantor, I.; Prakapenka, V.; Kantor, A.; Dera, P.; Kurnosov, A.; Sinogeikin, S.; Dubrovinskaia, N.; Dubrovinsky, L. BX90: A new diamond anvil cell design for X-ray diffraction and optical measurements. Rev. Sci. Instrum. 2012, 83, 125102. [Google Scholar] [CrossRef] [PubMed]
  11. Fei, Y.; Ricolleau, A.; Frank, M.; Mibe, K.; Shen, G.; Prakapenka, V. High-pressure geoscience special feature: Toward an internally consistent pressure scale. Proc. Natl. Acad. Sci. USA 2007, 104, 9182–9186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Ohishi, Y.; Hirao, N.; Sata, N.; Hirose, K.; Takata, M. Highly intense monochromatic X-ray diffraction facility for high-pressure research at SPring-8. High Press. Res. 2008, 28, 163–173. [Google Scholar] [CrossRef]
  13. 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]
  14. Bykova, E. Single-crystal X-ray Diffraction at Extreme Conditions in Mineral Physics and Material Sciences. Ph.D. Thesis, Universität Bayreuth, Bayreuth, Germany, 2015. [Google Scholar]
  15. CrysalisPRO (Rigaku Products Website). Available online: https://www.rigaku.com/products/smc/crysalis (accessed on 3 January 2020).
  16. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found Crystallogr. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  17. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  18. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  19. Hölsä, J.; Säilynoja, E.; Koski, K.; Rahiala, H.; Valkonen, J. X-ray powder diffraction study of the stability of solid solutions in (La1−xGdx)OCl. Powder Diffr. 1996, 11, 129–133. [Google Scholar] [CrossRef]
  20. Rozenberg, G.K.; Pasternak, M.P.; Gorodetsky, P.; Xu, W.M.; Dubrovinsky, L.S.; Le Bihan, T.; Taylor, R.D. Pressure-induced structural, electronic, and magnetic phase transitions in FeCl2 studied by X-ray diffraction and resistivity measurements. Phys. Rev. B 2009, 79, 1–7. [Google Scholar] [CrossRef]
  21. Bykov, M. Structural Aspects of Pressure- and Temperature-Induced Phase Transitions in Low-Dimensional Systems. Ph.D. Thesis, Bayreuth University, Bayreuth, Germany, June 2015. [Google Scholar]
  22. Troyanov, S.I. Crystalline structure of FeCl3 polytype modifications. Zhurnal Neorganicheskoi Khimii 1993, 38, 1946–1949. [Google Scholar]
  23. Schultz, A.J.; Carlin, R.L. Single-crystal pulsed neutron diffraction structure of the antiferromagnet K2[FeCl5(H2O)] with and without applied pressure. Acta Crystallogr. Sect. B 1995, 51, 43–47. [Google Scholar] [CrossRef]
Figure 1. Le-Bail fitting of an XRD pattern of a mixture of NaCl and FeOOH at 107(2) GPa after laser-heating to 2400(200) K. Black line—convoluted experimental data from a Perkin-Elmer detector (diamond reflections masked), red line—calculated XRD signal from several phases. Color ticks below the diffraction curve represent the reflections of the phases described in a legend. Rp = 12.49%.
Figure 1. Le-Bail fitting of an XRD pattern of a mixture of NaCl and FeOOH at 107(2) GPa after laser-heating to 2400(200) K. Black line—convoluted experimental data from a Perkin-Elmer detector (diamond reflections masked), red line—calculated XRD signal from several phases. Color ticks below the diffraction curve represent the reflections of the phases described in a legend. Rp = 12.49%.
Minerals 10 00051 g001
Figure 2. The crystal structure of Na2FeCl4OHx at 107(2) GPa. (a) Polyhedral structural model. Yellow spheres—Na atoms, brown spheres—Fe atoms, Cl and O atoms shown as green and red spheres, correspondingly. Yellow and brown polyhedra represent the coordination of Na and Fe irons. White spheres show the possible (partially occupied) positions of H atoms. Dashed black lines—unit cell boundaries. (b) Fe coordination polyhedra with bond lengths. (c) Na coordination polyhedra with bond lengths.
Figure 2. The crystal structure of Na2FeCl4OHx at 107(2) GPa. (a) Polyhedral structural model. Yellow spheres—Na atoms, brown spheres—Fe atoms, Cl and O atoms shown as green and red spheres, correspondingly. Yellow and brown polyhedra represent the coordination of Na and Fe irons. White spheres show the possible (partially occupied) positions of H atoms. Dashed black lines—unit cell boundaries. (b) Fe coordination polyhedra with bond lengths. (c) Na coordination polyhedra with bond lengths.
Minerals 10 00051 g002
Figure 3. The estimation of a Fe–Cl bond length in different compounds at high pressures. FeCl2 data are from [20], FeOCl data are from [21]. Experimental values of Fe–Cl bond length for FeCl3 and K2FeCl5(H2O) are from [22] and [23], correspondingly. Red star—experimental point for Na2FeCl4OHx.
Figure 3. The estimation of a Fe–Cl bond length in different compounds at high pressures. FeCl2 data are from [20], FeOCl data are from [21]. Experimental values of Fe–Cl bond length for FeCl3 and K2FeCl5(H2O) are from [22] and [23], correspondingly. Red star—experimental point for Na2FeCl4OHx.
Minerals 10 00051 g003
Table 1. Crystallographic information on the Na2FeCl4OHx.
Table 1. Crystallographic information on the Na2FeCl4OHx.
Crystal Data
Chemical formulaNa2FeCl4OHx
Mr (g/mol)260.6
Crystal system, space groupOrthorhombic, Pbam (#55)
Temperature (K), Pressure (GPa)293, 107(2)
a,b,c (Å)8.725 (2), 6.180 (3), 3.0679 (12)
V (Å3)165.41 (11)
Z2
Radiation typeSynchrotron, λ = 0.2892 Å
μ (mm−1)0.61
Crystal shapeCube
Crystal size (mm)0.01 × 0.01 × 0.01
Data collection
DiffractometerP02.2, DESY
Absorption correctionMulti-scan (ABSPACK, [15])
Tmin, Tmax0.538, 1
No. of measured, independent and observed [I > 3σ(I)] reflections894, 432, 315
Rint0.038
θ values (°)θ = 3.2–18.4
(sin θ/λ)max (Å-1)1.089
Range of h, k, lh = −16→13, k = −8→10, l = −5→4
Refinement
R[F2 > 2σ(F2)], wR(F2), S0.070, 0.071, 2.93
No. of reflections315
No. of parameters27
Δρmax, Δρmin (eÅ−3)2.07, –1.63
Software: Crysalis PRO [15], SHELXT [16], SHELXL2014 [17].
Table 2. Atomic coordinates in the Na2FeCl4OHx structure.
Table 2. Atomic coordinates in the Na2FeCl4OHx structure.
Atom NumberAtom NameAtomic Coordinates (x, y, z)OccupancyIsotropic Atomic Displacement Parameter (U, Å2)Wyckoff
1Fe10.5, 0.0, 0.010.012 c
2Cl10.2076, 0.3455, 0.010.0124 g
3Cl20.4251, 0.1907, 0.510.0114 h
4Na10.1645, 0.0701, 0.510.0124 h
5O10.0, 0.0, 0.010.0022 a
6H10.00380, 0.11683, 0.17495 0.250.08 i

Share and Cite

MDPI and ACS Style

Koemets, E.; Yuan, L.; Bykova, E.; Glazyrin, K.; Ohtani, E.; Dubrovinsky, L. Interaction Between FeOOH and NaCl at Extreme Conditions: Synthesis of Novel Na2FeCl4OHx Compound. Minerals 2020, 10, 51. https://doi.org/10.3390/min10010051

AMA Style

Koemets E, Yuan L, Bykova E, Glazyrin K, Ohtani E, Dubrovinsky L. Interaction Between FeOOH and NaCl at Extreme Conditions: Synthesis of Novel Na2FeCl4OHx Compound. Minerals. 2020; 10(1):51. https://doi.org/10.3390/min10010051

Chicago/Turabian Style

Koemets, Egor, Liang Yuan, Elena Bykova, Konstantin Glazyrin, Eiji Ohtani, and Leonid Dubrovinsky. 2020. "Interaction Between FeOOH and NaCl at Extreme Conditions: Synthesis of Novel Na2FeCl4OHx Compound" Minerals 10, no. 1: 51. https://doi.org/10.3390/min10010051

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