Next Article in Journal
Anomalous Behaviors of Spin Waves Studied by Inelastic Light Scattering
Next Article in Special Issue
From Molecules to Carbon Materials—High Pressure Induced Polymerization and Bonding Mechanisms of Unsaturated Compounds
Previous Article in Journal
Mg-Fe Layered Double Hydroxides Enhance Surfactin Production in Bacterial Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Evidence for Partially Dehydrogenated ε-FeOOH

1
Center for High Pressure Science and Technology Advanced Research, Beijing 100094, China
2
Hawai’i Institute of Geophysics and Planetology, School of Ocean and Earth Science and Technology, University of Hawai’i at Manoa, Honolulu, HI 96822, USA
3
Geophysical Laboratory, Carnegie Institution of Washington, Washington, DC 20015, USA
*
Author to whom correspondence should be addressed.
Crystals 2019, 9(7), 356; https://doi.org/10.3390/cryst9070356
Submission received: 18 June 2019 / Revised: 5 July 2019 / Accepted: 12 July 2019 / Published: 13 July 2019
(This article belongs to the Special Issue High Pressure Synthesis in Crystalline Materials)

Abstract

:
Hydrogen in hydrous minerals becomes highly mobile as it approaches the geotherm of the lower mantle. Its diffusion and transportation behaviors under high pressure are important in order to understand the crystallographic properties of hydrous minerals. However, they are difficult to characterize due to the limit of weak X-ray signals from hydrogen. In this study, we measured the volume changes of hydrous ε-FeOOH under quasi-hydrostatic and non-hydrostatic conditions. Its equation of states was set as the cap line to compare with ε-FeOOH reheated and decompression from the higher pressure pyrite-FeO2Hx phase with 0 < x < 1. We found the volumes of those re-crystallized ε-FeOOH were generally 2.2% to 2.7% lower than fully hydrogenated ε-FeOOH. Our observations indicated that ε-FeOOH transformed from pyrite-FeO2Hx may inherit the hydrogen loss that occurred at the pyrite-phase. Hydrous minerals with partial dehydrogenation like ε-FeOOHx may bring it to a shallower depth (e.g., < 1700 km) of the lower mantle.

1. Introduction

The major components of the lower mantle are made of nominally anhydrous minerals, which contain no more than 1 weight percent of water [1,2]. However, recent discoveries of deep hydrous phase (DHP) including δ-AlOOH [3,4,5], phase H [6], HH-phase [7], and pyrite-type FeO2Hx [8,9], which were synthesized at conditions of cold mantle geotherm from natural minerals like diaspore (α-AlOOH) and goethite (α-FeOOH), provide possible mechanisms to transport a significant amount of water to the bottom of the mantle. The potential presence of hydrogen is likely to contribute a variety of seismological features observed at Earth’s lower mantle. For example, dehydration melting at the top of the lower mantle could dramatically decrease the seismic velocities below a depth of 660 kilometers [10]. Accumulation of iron-enriched hydrous pyrite-type phase would reduce the speed of seismic waves at the core-mantle boundary, which may be detected at large low shear velocity provinces and ultra-low velocity zones [11,12,13]. Reservoirs of H also produce hydrides that would possibly infiltrate to the outer core [9,14]. Although H is an influential volatile component, the budget of H in the lower mantle is still under debate [15]. Large uncertainties in the abundance are probably due to the scarcity to find natural samples derived from the deep mantle [16]. However, the number of DHPs revealed by laboratory experiments continues to grow with the development of high-pressure synchrotron-based experiments [17]. The extraordinary thermal stabilities of DHPs suggest that the lower mantle can hold more water than previously expected.
Discoveries of novel DHPs have attracted an appreciable amount of research efforts [5,18,19,20]. What is equally important is the diffusion and transportation behaviors of hydrogen in those DHPs [21]. Since hydrogen is the lightest element and is highly mobile, an outstanding question is how we can quantify the hydrogen content in DHP under high pressure. It becomes even more challenging when DHP is partially dehydrogenated [22,23]. In this case, we measured the equation of state of ε-FeOOH, a typical DHP in the Fe-O-H ternary system, and studied its hydrogen content based on its volume variation. This method has been testified to determine the hydrogen content in hydride [24] and pyrite-type FeO2Hx [22]. We provide evidence that the ε-FeOOH phase could undergo partial dehydrogenation and provide evidence that a new form of ε-FeOOHx (0 < x < 1) can be stabilized at lower mantle conditions. Similar to pyrite-type FeO2Hx (0 < x < 1), ε-FeOOHx may exist as a solid solution of ε-FeOOH and FeO2 [8].

2. Materials and Methods

We started our experiment by synthesizing ε-FeOOH from goethite (α-FeOOH, CAS: 20344-49-4). Goethite comprises double chains of edge-shared octahedra that form 2 × 1 channels hosting hydrogen bonds at an ambient condition. Above 5 GPa, it transforms to a high-pressure phase of ε-FeOOH, which consists of corner-shared single bands of octahedral. ε-FeOOH is thermodynamically stable in cold subducted slabs [25]. It is quenchable to ambient conditions once the crystal is synthesized in a multi-anvil press [26,27]. We followed the recipe by Suzuki [26,27] by compacting goethite powder in a gold capsule, rolling in a rhenium heater, and placing it in a Kawai-type multi-anvil press at the Geophysical Laboratory, Carnegie Institution of Washington. The use of the gold capsule was to seal water from out-of-capsule diffusion. The ε-FeOOH sample was synthesized at 10 GPa and 800 °C for 4 hours. After quenching to ambient conditions, the recovered samples were examined and confirmed on a diffractometer (Figure 1). The lattice parameters and volume at ambient conditions of the synthesized ε-FeOOH phase (orthorhombic, P21nm) were listed in Table 1 when compared with ε-FeOOH synthesized in other laboratories [26]. We found the ambient lattice parameters of our sample were consistent with other studies, which confirmed the composition of the synthesized ε-FeOOH was fully hydrogenated.
We then loaded powder ε-FeOOH in a diamond anvil cell (DAC) with Ne as the pressure medium. A second piece of ε-FeOOH from the same source was loaded without any pressure medium to check the effect of anisotropic compression [28]. The sample chamber was built by drilling a 100 to 120 μm hole in a tungsten gasket, which was squeezed between two diamond anvils with a 260 μm culet. Angular dispersive X-ray diffraction (XRD) patterns were obtained at the 13BM-C station of the GeoSoilEnviroCARS at the Advanced Phonon Source, Argonne National Laboratory (Argonne, IL, USA). The wavelength of the incident X-ray was 0.434 Å and the initial data reduction was performed by the Dioptas program [29]. We chose both ruby and gold for pressure calibration [30,31].

3. Results

In Figure 2, the high-pressure XRD patterns can be indexed to ε-FeOOH, gasket, and ruby. We collected XRD patterns up to 56.3 GPa and ε-FeOOH is still stable. Figure S1 (in Supplementary Materials) shows the selected XRD pattern of compared ε-FeOOH without a pressure medium. The molar volume of ε-FeOOH in Ne pressure medium is 24.7(3) Å3 at 56.3 GPa, which is consistent with literature [32]. The volumes as a function of pressure were plotted in Figure 3a. Under a hydrostatic condition, the change of volume became much more incompressible at around 45 GPa, which may attribute to the spin transition of iron [32,33]. Likewise, spin transition occurred about 4–5 GPa in advance under non-hydrostatic conditions, which is possibly affected by the deviatoric stress [32]. In Figure 3b, we showed the pressure dependence of the relative lattice constants of ε-FeOOH. By comparing their evolution of the edge length along the lattice axis, we found the a and b axes were more affected by hydrostaticity and, thus, became more compressible under non-hydrostatic conditions. The elastic anisotropy may be the result of pressure induced hydrogen bond symmetrization in ε-FeOOH [32]. The centering of H atoms in two O atoms induced the shortening of O-O distance along the b axis. An interesting coincidence was the same H-bond behavior in the iso-structured δ-AlOOH, whose thermal stability is significantly improved due to the H-bond symmetry [34]. The same symmetrical hydrogen bonds in ε-FeOOH may also expand its thermal stability field to higher temperatures.
The elastic parameters were derived by the fitting of PV data to the third order Birch-Murnaghan equation of states (EOS).
P = 3 2 K T 0 [ ( V 0 V ) 7 3 ( V 0 V ) 5 3 ] × { 1 + 3 4 ( K 0 4 ) [ ( V 0 V ) 2 3 1 ] }
where P is the pressure, KT0 is the isothermal bulk modulus, V0 and V are the volumes at high pressure and ambient condition, respectively, and K0 is the pressure derivative of KT0 at 1 bar. The data were listed in Table 2 [35]. It should be noted that the EOS of ε-FeOOH was separated to two regions due to the high-spin to low-spin transition. For the high-spin phase, the bulk modulus K0 equals to 125.2(4) GPa with V0 per formula unit (f.u.) of 32.4(4) Å at hydrostatic condition, which was consistent with previous studies [26]. The low-spin state has a much higher bulk modulus K0 = 248.5(2) GPa. Consequently, the sample became more stiffened and incompressible in the low-spin state. The parallel non-hydrostatic experiments generally reproduced the results from hydrostatic conditions, with slightly lower bulk modulus of K0 = 114.3(8) GPa in a high spin (Table 2). While the calculated V0 is almost the same, a smaller K0 from non-hydrostatic compression means that anisotropic compression may facilitate the formation of symmetric H bond in compressed ε-FeOOH.
The equation of the state is the key to study the diffusion and transportation behaviors of H in ε-FeOOH. Our previous simulation results suggested that ε-FeOOH with 0–75% of H defects is still energetically stable and the H loss may promote its phase transition to the pyrite-type FeOOH by lowering the transition barrier [36]. Removing all the hydrogen in ε-FeOOH leads to a FeO2 stoichiometry, which is a pyrite-type phase synthesized above 74 GPa [8,37]. However, FeO2 is also reported to have a few low-pressure polymorphs. For example, below ~50 GPa, the pyrite-FeO2 may transform to an orthorhombic FeO2 with Pnnm [38] or Pbcn [39] phase, which are both very similar to the crystal structure of ε-FeOOH. Would it be possible to form a complete solid-solution of ε-FeOOH and the corresponding orthorhombic FeO2 [8]? While ε-FeOOH synthesized from α-FeOOH and multi-anvil press is guaranteed to be fully hydrogenated, ε-FeOOH transformed from the pyrite-FeO2Hx might inherit the hydrogen loss from its high-pressure polymorph. We took data from our previous experiments [22] and compared it with the compression curve of ε-FeOOH. In Figure 4, we used the EOS of ε-FeOOH as the cap line of the fully hydrogenated phase. The volumes by subtracting 0.5 mole of hydrogen (Phase I [40]) from ε-FeOOH served as the baseline for hydrogen depleted ε-FeOOHx (x = 0). By comparing their molar volumes, we can calculate the H content x for possible H depletion in ε-FeOOHx through the following equation.
x = [VxVe0]/VH
where Vx is the observed volume of ε-FeOOHx from the experiment, Ve0 is the volume of hydrogen depleted ε-FeOOHx (the baseline with x = 0), and VH is the volume per formula unit of phase-I solid H2, which is stable up to the pressures of Earth’s core-mantle boundary [41]. The unit cell volumes of phase-I H2 was obtained from Reference [40]. The same scheme was widely used in estimating hydrogen numbers in iron hydrides [24]. In Table 3, we compiled the volumes of FeOOHx and evaluated x from our previous experiments and the literature [9,22,36]. The values of x are within the range of 0.47–0.75, which are in the same range of our previous estimated x in pyrite-type FeO2Hx [22]. The coincidence suggests the loss of H in the recrystallized ε-FeOOHx samples may come from its high-pressure polymorph. We also noticed that the volume of ε-FeOOH reported by Nishi et al. [9] was 2.2% to 2.7% below the capline of ε-FeOOH, which is equivalent to 0.69 < x < 0.74. It might be due to a dehydrogenation process by reheating a pyrite-type FeOOH sample at a relatively low pressure. We summarized the transportation of H in the polymorphic transition between ε-FeOOHx and pyrite-type FeO2Hx in Figure 4, where the O-H-O framework is re-established in ε-FeOOHx with the same amount of H defects. As a result, ε-FeOOHx can be regarded as a solid solution of ε-FeOOH and orthorhombic FeO2. Even if there is no evidence that temperature-induced dehydrogenation will occur in ε-FeOOH, the H loss may be exhibited as an intrinsic property of pyrite-type FeO2Hx.

4. Discussion

During the transformation from ε-FeOOH to the pyrite structured FeO2H, the emission of H is still a controversy in experiments [9,22]. The current work used fully hydrogenated ε-FeOOH as the cap line that was derived from the starting composition and compared with the volumes of ε-FeOOHx. In comparison, our previous work on pyrite-type FeO2Hx used the baseline of FeO2 [22]. The volume added to the baseline might be slightly underestimated due to the volume collapse amid the phase transition. Therefore, our current estimation on ε-FeOOHx was more reliable and sensitive to H loss. The results confirmed both ε-FeOOHx and pyrite-type FeOOHx are partially dehydrogenated.
Our EOS of ε-FeOOH under different hydrostatic environments extended the studies of ε-FeOOH [26,27,32] to a wider pressure range. The XRD experiments of ε-FeOOH and ε-FeOOHx have many implications for hydrogen diffusion and transportation in the deep Earth. H loss in transformed ε-FeOOH indicated H transportation is not affected by the phase transition. Therefore, recycled hydrogenated phases near the core-mantle boundary will retain the loss of hydrogen when they were subducted to a shallower part of the lower mantle even when they transformed to other phases like ε-FeOOH. It indicates ε-FeOOH in the lower mantle may have a significant amount of H defects. The total H content in the lower mantle may not be estimated by the H amount in full hydrogenated DHP. Instead, the case of dehydrogenation needs to be considered. Our experimental results support the H loss phenomenon in the high-pressure phases of FeOOH. The released hydrogen may escape or dissolve in the surrounding Earth materials as a part of the hydrogen cycle in Earth’s deep interiors.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4352/9/7/356/s1, Figure S1: Selected x-ray diffraction patterns of compressed ε-FeOOH under non-hydrostatic condition.

Author Contributions

Y.Z., Z.C., and D.Z. conducted the experiment. R.T. synthesized the sample. Y.Z., J.L., and Q.H. wrote the paper. Q.H. conceived the work. All authors discussed the results.

Funding

GeoSoilEnviroCARS (GSECARS) is supported by National Science Foundation - Earth Sciences (EAR-1128799) and Department of Energy - Geosciences (DE-FG02-94ER14466). YZ is supported by China Postdoctoral Science Foundation with grant 18NZ021-0213-216308. Operations of Center for High Pressure Science and Technology Advanced Research (HPSTAR) is partially supported by NSAF (Grant No: U1530402).

Acknowledgments

The authors acknowledge the use of synchrotron X-ray diffraction at the 13BM-C of GSECARS, Advanced Photon Source, Argonne National Laboratory.

Conflicts of Interest

The authors declare no conflict of interest.

Angular Dispersive X-ray Diffraction Experiments

Angular dispersive x-ray diffraction experiments were performed at 13BM-C station of the GSECARS at the APS, ANL. Samples of ε-FeOOH powders were grinded and pre-compressed to ~15 (T) × 60 (W) × 60 (L) μm3 before placing on the culet of DAC. High pressure was achieved by using diamond anvils with 260 μm culet diameter. The sample chamber was a drilled hole with 100–120 μm diameter in a tungsten gasket. For quasi-hydrostatic condition, neon gas was pumped into the sample chamber using a gas-loading system at HPSTAR. For non-hydrostatic condition, no pressure medium was used. One or two pieces of ruby were placed around the sample to calibrate pressure. The ruby pressure scale was compared with the equation of state of gold [30] and Ne [42]. Pressure uncertainty is up to ± 1 GPa throughout the experiment.

References

  1. Bell, D.R.; Rossman, G.R. Water in Earth’s Mantle: The Role of Nominally Anhydrous Minerals. Science 1992, 255, 1391–1397. [Google Scholar] [CrossRef]
  2. Kaminsky, F.V. Water in the Earth’s Lower Mantle. Geochem. Int. 2018, 56, 1117–1134. [Google Scholar] [CrossRef]
  3. Tsuchiya, J.; Tsuchiya, T.; Tsuneyuki, S.; Yamanaka, T. First Principles Calculation of a High-Pressure Hydrous Phase, δ-AlOOH. Geophys. Res. Lett. 2002, 29, 15:1–15:4. [Google Scholar] [CrossRef]
  4. Duan, Y.; Sun, N.; Wang, S.; Li, X.; Guo, X.; Ni, H.; Prakapenka, V.B.; Mao, Z. Phase Stability and Thermal Equation of State of δ-AlOOH: Implication for Water Transportation to the Deep Lower Mantle. Earth Planet. Sci. Lett. 2018, 494, 92–98. [Google Scholar] [CrossRef]
  5. Sano, A.; Ohtani, E.; Kondo, T.; Hirao, N.; Sakai, T.; Sata, N.; Ohishi, Y.; Kikegawa, T. Aluminous Hydrous Mineral δ-AlOOH as a Carrier of Hydrogen into the Core-Mantle Boundary. Geophys. Res. Lett. 2008, 35, L03303:1–L03303:5. [Google Scholar] [CrossRef]
  6. Nishi, M.; Irifune, T.; Tsuchiya, J.; Tange, Y.; Nishihara, Y.; Fujino, K.; Higo, Y. Stability of Hydrous Silicate at High Pressures and Water Transport to the Deep Lower Mantle. Nat. Geosci. 2014, 7, 224–227. [Google Scholar] [CrossRef]
  7. Zhang, L.; Yuan, H.; Meng, Y.; Mao, H.K. Discovery of a Hexagonal Ultradense Hydrous Phase in (Fe,Al)OOH. Proc. Natl. Acad. Sci. 2018, 115, 2908–2911. [Google Scholar] [CrossRef] [PubMed]
  8. 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]
  9. Nishi, M.; Kuwayama, Y.; Tsuchiya, J.; Tsuchiya, T. the Pyrite-Type High-Pressure Form of FeOOH. Nature 2017, 547, 205–208. [Google Scholar] [CrossRef]
  10. Schmandt, B.; Jacobsen, S.D.; Becker, T.W.; Liu, Z.; Dueker, K.G. Dehydration Melting at the Top of the Lower Mantle. Science 2014, 344, 1265–1268. [Google Scholar] [CrossRef]
  11. 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]
  12. Liu, J.; Hu, Q.; Kim, D.Y.; Wu, Z.; Wang, W.; Xiao, Y.; Chow, P.; Meng, Y.; Prakapenka, V.B.; Mao, H.K.; et al. Hydrogen-Bearing Iron Peroxide and the Origin of Ultralow-Velocity Zones. Nature 2017, 551, 494–497. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, J.; Lv, J.; Li, H.; Feng, X.; Lu, C.; Redfern, S.A.T.; Liu, H.; Chen, C.; Ma, Y. Rare Helium-Bearing Compound FeO2He Stabilized at Deep-Earth Conditions. Phys. Rev. Lett. 2018, 121, 255703. [Google Scholar] [CrossRef] [PubMed]
  14. Thompson, E.C.; Davis, A.; Bi, W.; Zhao, J.; Alp, E.; Zhang, D.; Greenberg, E.; Prakapenka, V.B.; Campbell, A.J. High-Pressure Geophysical Properties of Fcc Phase FeHX. Geochem. Geophys. Geosyst. 2018, 19, 305–314. [Google Scholar] [CrossRef]
  15. Ni, H.; Zheng, Y.; Mao, Z.; Wang, Q.; Chen, R.; Zhang, L. Distribution, Cycling and Impact of Water in the Earth’s Interior. Natl. Sci. Rev. 2017, 4, 879–891. [Google Scholar] [CrossRef]
  16. Tschauner, O.; Huang, S.; Greenberg, E.; Prakapenka, V.B.; Ma, C.; Rossman, G.R.; Shen, A.H.; Zhang, D.; Newville, M.; Lanzirotti, A.; et al. Ice-VII Inclusions in Diamonds: Evidence for Aqueous Fluid in Earth’s Deep Mantle. Science 2018, 359, 1136–1139. [Google Scholar] [CrossRef]
  17. Mao, H.K.; Chen, B.; Chen, J.; Li, K.; Lin, J.F.; Yang, W.; Zheng, H. Recent Advances in High-Pressure Science and Technology. Matter Radiat. Extremes 2016, 1, 59–75. [Google Scholar] [CrossRef]
  18. Suzuki, A.; Ohtani, E.; Kamada, T. A New Hydrous Phase δ-AlOOH Synthesized at 21 GPa and 1000 °C. Phys. Chem. Miner. 2000, 27, 689–693. [Google Scholar] [CrossRef]
  19. Walter, M.J.; Thomson, A.R.; Wang, W.; Lord, O.T.; Ross, J.; McMahon, S.C.; Baron, M.A.; Melekhova, E.; Kleppe, A.K.; Kohn, S.C. The Stability of Hydrous Silicates in Earth’s Lower Mantle: Experimental Constraints from the Systems MgO–SiO2–H2O and MgO–Al2O3–SiO2–H2O. Chem. Geol. 2015, 418, 16–29. [Google Scholar] [CrossRef]
  20. 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, 153. [Google Scholar] [CrossRef]
  21. 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]
  22. Hu, Q.; Kim, D.Y.; Liu, J.; Meng, Y.; Yang, L.; Zhang, D.; Mao, W.L.; Mao, H.K. Dehydrogenation of Goethite in Earth’s Deep Lower Mantle. Proc. Natl. Acad. Sci. USA 2017, 114, 1498–1501. [Google Scholar] [CrossRef] [PubMed]
  23. Boulard, E.; Guyot, F.; Menguy, N.; Corgne, A.; Auzende, A.; Perrillat, J.; Fiquet, G. CO2-Induced Destabilization of Pyrite-Structured FeO2Hx in the Lower Mantle. Natl. Sci. Rev. 2018, 5, 870–877. [Google Scholar] [CrossRef]
  24. Pépin, C.M.; Geneste, G.; Dewaele, A.; Mezouar, M.; Loubeyre, P. Synthesis of FeH5: A Layered Structure with Atomic Hydrogen Slabs. Science 2017, 357, 382–385. [Google Scholar] [CrossRef] [PubMed]
  25. Gleason, A.E.; Jeanloz, R.; Kunz, M. Pressure-Temperature Stability Studies of FeOOH Using X-Ray Diffraction. Am. Miner. 2008, 93, 1882–1885. [Google Scholar] [CrossRef]
  26. Suzuki, A. Pressure–Volume–Temperature Equation of State of ε-FeOOH to 11 GPa and 700 K. J. Miner. Petrol. Sci. 2016, 111, 420–424. [Google Scholar] [CrossRef]
  27. Suzuki, A. High-Pressure X-Ray Diffraction Study of ε-FeOOH. Phys. Chem. Miner. 2009, 37, 153–157. [Google Scholar] [CrossRef]
  28. Zhuang, Y.; Dai, L.; Wu, L.; Li, H.; Hu, H.; Liu, K.; Yang, L.; Pu, C. Pressure-Induced Permanent Metallization with Reversible Structural Transition in Molybdenum Disulfide. Appl. Phys. Lett. 2017, 110, 122103. [Google Scholar] [CrossRef]
  29. 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]
  30. Mao, H.K.; Xu, J.; Bell, P.M. Calibration of the Ruby Pressure Gauge to 800 Kbar under Quasi-Hydrostatic Conditions. J. Geophys. Res. 1986, 91, 4673–4676. [Google Scholar] [CrossRef]
  31. Ye, Y.; Shim, S.H.; Prakapenka, V.; Meng, Y. Equation of State of Solid Ne Inter-Calibrated with the MgO, Au, Pt, NaCl-B2, and Ruby Pressure Scales up to 130 GPa. High Press. Res. 2018, 38, 377–395. [Google Scholar] [CrossRef]
  32. Gleason, A.E.; Quiroga, C.E.; Suzuki, A.; Pentcheva, R.; Mao, W.L. Symmetrization Driven Spin Transition in ε-FeOOH at High Pressure. Earth Planet. Sc. Lett. 2013, 379, 49–55. [Google Scholar] [CrossRef]
  33. Liu, J.; Dorfman, S.M.; Zhu, F.; Li, J.; Wang, Y.; Zhang, D.; Xiao, Y.; Bi, W.; Alp, E.E. Valence and Spin States of Iron Are Invisible in Earth’s Lower Mantle. Nat. Commun. 2018, 9, 1284. [Google Scholar] [CrossRef] [PubMed]
  34. Sano-Furukawa, A.; Komatsu, K.; Vanpeteghem, C.B.; Ohtani, E. Neutron Diffraction Study of δ-AlOOD at High Pressure and Its Implication for Symmetrization of the Hydrogen Bond. Am. Mineral. 2008, 93, 1558–1567. [Google Scholar] [CrossRef]
  35. Thompson, E.C.; Campbell, A.J.; Tsuchiya, J. Elasticity of ε-FeOOH: Seismic implications for Earth’s lower mantle. J. Geophys. Res: Sol. Ea. 2017, 122, 5038–5047. [Google Scholar] [CrossRef]
  36. Zhu, S.C.; Hu, Q.; Mao, W.L.; Mao, H.K.; Sheng, H. Hydrogen-Bond Symmetrization Breakdown and Dehydrogenation Mechanism of FeO2H at High Pressure. J. Am. Chem. Soc. 2017, 139, 12129–12132. [Google Scholar] [CrossRef] [PubMed]
  37. Zhu, S.; Liu, J.; Hu, Q.; Mao, W.L.; Meng, Y.; Zhang, D.; Mao, H.-K.; Zhu, Q. Structure-Controlled Oxygen Concentration in Fe2O3 and FeO2. Inorg. Chem. 2019, 58, 5476–5482. [Google Scholar] [CrossRef]
  38. Tang, M.; Niu, Z.W.; Zhang, X.L.; Cai, L.C. Structural Stability of FeO2 in the Pressure Range of Lower Mantle. J. Alloys Compd. 2018, 765, 271–277. [Google Scholar] [CrossRef]
  39. Lu, C.; Amsler, M.; Chen, C. Unraveling the Structure and Bonding Evolution of the Newly Discovered Iron Oxide FeO2. Phys. Rev. B 2018, 98, 054102. [Google Scholar] [CrossRef]
  40. Loubeyre, P.; LeToullec, R.; Hausermann, D.; Hanfland, M.; Hemley, R.J.; Mao, H.K.; Finger, L.W. X-Ray Diffraction and Equation of State of Hydrogen at Megabar Pressures. Nature 1996, 383, 702–704. [Google Scholar] [CrossRef]
  41. Howie, R.T.; Dalladay-Simpson, P.; Gregoryanz, E. Raman spectroscopy of hot hydrogen above 200 GPa. Nat. Mater. 2015, 14, 495–499. [Google Scholar] [CrossRef] [PubMed]
  42. Fei, Y.; Ricolleau, A.; Frank, M.; Mibe, K.; Shen, G.; Prakapenka, V. Toward an Internally Consistent Pressure Scale. Proc. Nat. Acad. Sci. 2007, 104, 9182–9186. [Google Scholar] [CrossRef] [PubMed]
Figure 1. X-ray diffraction pattern of synthetic ε-FeOOH at ambient conditions. The sample was checked under a Brucker D8 diffractometer with a wavelength of 1.5406 Å. The diffraction pattern shows a pure phase of ε-FeOOH sample along with residual Au from the capsule.
Figure 1. X-ray diffraction pattern of synthetic ε-FeOOH at ambient conditions. The sample was checked under a Brucker D8 diffractometer with a wavelength of 1.5406 Å. The diffraction pattern shows a pure phase of ε-FeOOH sample along with residual Au from the capsule.
Crystals 09 00356 g001
Figure 2. Selected x-ray diffraction patterns of compressed ε-FeOOH in Ne.
Figure 2. Selected x-ray diffraction patterns of compressed ε-FeOOH in Ne.
Crystals 09 00356 g002
Figure 3. Equation of state for ε-FeOOH. (a) Volume versus pressure for ε-FeOOH in this study and literatures. Pressure-volume data for their high-spin or low-spin states were fitted to the third order Birch-Murnaghan equation of states. “X” and reversed triangle symbol were ε-FeOOH obtained by reheating Py-FeO2Hx out of its stability range. Open spheres were fully hydrogenated ε-FeOOH. Pressure was calibrated by ruby and Au (gold) with up to ± 1 GPa uncertainty. (b) The lattice parameters of ε-FeOOH as a function of pressure under different pressure environments. a1, b1, and c1 for hydrostatic conditions and a2, b2, and c2 for non-hydrostatic conditions. PM, pressure medium.
Figure 3. Equation of state for ε-FeOOH. (a) Volume versus pressure for ε-FeOOH in this study and literatures. Pressure-volume data for their high-spin or low-spin states were fitted to the third order Birch-Murnaghan equation of states. “X” and reversed triangle symbol were ε-FeOOH obtained by reheating Py-FeO2Hx out of its stability range. Open spheres were fully hydrogenated ε-FeOOH. Pressure was calibrated by ruby and Au (gold) with up to ± 1 GPa uncertainty. (b) The lattice parameters of ε-FeOOH as a function of pressure under different pressure environments. a1, b1, and c1 for hydrostatic conditions and a2, b2, and c2 for non-hydrostatic conditions. PM, pressure medium.
Crystals 09 00356 g003
Figure 4. A schematic figure showing partially dehydrogenated ε-FeOOHx. The ε-FeOOHx phase inherits the hydrogen loss from pyrite-type FeO2Hx.
Figure 4. A schematic figure showing partially dehydrogenated ε-FeOOHx. The ε-FeOOHx phase inherits the hydrogen loss from pyrite-type FeO2Hx.
Crystals 09 00356 g004
Table 1. Structural data for the synthesized ε-FeOOH under an ambient environment.
Table 1. Structural data for the synthesized ε-FeOOH under an ambient environment.
PhaseV0/Z (A3)Lattice Parameters (Å, Degree)Reference
abc
ε-FeOOH33.139(3)4.9544(2)4.4594(3)2.9999(1)Suzuki et al. 2016 [26]
ε-FeOOH33.10(3)4.954(1)4.4540(9)3.0001(8)Suzuki et al. 2009 [27]
ε-FeOOH32.636(3)5.273(4)4.423(3)2.798(2)This study
Table 2. Compressibility parameters of ε-FeOOH. Abbreviations: PM, Pressure medium. AH, asymmetric O-H bonding. SH, symmetric O-H-O bonding. Calc., first-principles calculation.
Table 2. Compressibility parameters of ε-FeOOH. Abbreviations: PM, Pressure medium. AH, asymmetric O-H bonding. SH, symmetric O-H-O bonding. Calc., first-principles calculation.
Method/PMNoteK0 (GPa)K0V0/Z (A3)Reference
Calc.AH/high spin188(4)5.19(12)28.7(2)Thompson et al. 2017 [35]
Calc.SH/low spin223(2)4.07(3)29.3(1)Thompson et al. 2017 [35]
No PMhigh spin124(4)433Gleason et al. 2013 [32]
No PMlow spin1624-Gleason et al. 2013 [32]
NaClhigh spin126(3)10(1)33.1(3)Suzuki et al. 2009 [27]
NaClhigh spin135(3)6.1(9)33.1(3)Suzuki et al. 2016 [26]
Nehigh spin125.2(4)3.9(2)32.4(4)This study
Nelow spin248.5(2)3.9(8)29.1(2)This study
no PMhigh spin114.3(8)3.7(8)32.6(7)This study
Table 3. Hydrogen content x in ε-FeOOHx.
Table 3. Hydrogen content x in ε-FeOOHx.
Pressure (GPa)V0/f.u. (A3)xReference
7223.040.47Hu et al. 2017 [22]
7822.760.46Hu et al. 2017 [22]
87.523.150.75Zhu et al. 2017 [36]
52~24.20.69Nishi et al. 2017 [9]

Share and Cite

MDPI and ACS Style

Zhuang, Y.; Cui, Z.; Zhang, D.; Liu, J.; Tao, R.; Hu, Q. Experimental Evidence for Partially Dehydrogenated ε-FeOOH. Crystals 2019, 9, 356. https://doi.org/10.3390/cryst9070356

AMA Style

Zhuang Y, Cui Z, Zhang D, Liu J, Tao R, Hu Q. Experimental Evidence for Partially Dehydrogenated ε-FeOOH. Crystals. 2019; 9(7):356. https://doi.org/10.3390/cryst9070356

Chicago/Turabian Style

Zhuang, Yukai, Zhongxun Cui, Dongzhou Zhang, Jin Liu, Renbiao Tao, and Qingyang Hu. 2019. "Experimental Evidence for Partially Dehydrogenated ε-FeOOH" Crystals 9, no. 7: 356. https://doi.org/10.3390/cryst9070356

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