Plan
Comptes Rendus

Internal Geophysics (Physics of Earth's interior)
Electrical resistivity of fcc phase iron hydrides at high pressures and temperatures
Comptes Rendus. Géoscience, Volume 351 (2019) no. 2-3, pp. 147-153.

Résumé

We measured the electrical resistivity of face-centered-cubic (fcc) structured iron hydrides at high pressures up to 65 GPa and high temperatures in a laser-heated diamond anvil cell. The results indicate that the resistivity of stoichiometric fcc FeHx (x ∼ 1.0) is smaller than that of fcc Fe at the same pressure and temperature conditions. The same behavior was also observed in fcc FeNiHx (x ∼ 1.0). On the other hand, hydrogen-poor fcc FeHx (x < 0.77) showed a resistivity comparable to that of the fcc phase of pure iron. Therefore, we conclude that the stoichiometric fcc Fe (–Ni) hydride is more conductive than Fe (–Ni) with the same crystal symmetry, and the impurity resistivity of hydrogen in Fe is vanishingly small. Even if hydrogen is the major light element in the Earth's core, it would have little influence on the electrical and thermal conductivity of Fe–Ni alloys, and hence the thermal evolution of the core.

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crte.2018.05.004
Mots clés : Iron hydride, Electrical resistivity, Earth's core, Core density deficit
Kenji Ohta 1 ; Sho Suehiro 1 ; Kei Hirose 2, 3 ; Yasuo Ohishi 4

1 Department of Earth and Planetary Sciences, Tokyo Institute of Technology, Meguro, Tokyo, Japan
2 Earth-Life Science Institute, Tokyo Institute of Technology, Meguro, Tokyo, Japan
3 Department of Earth and Planetary Science, The University of Tokyo, Bunkyo, Tokyo, Japan
4 Japan Synchrotron Radiation Research Institute, Sayo, Hyogo, Japan
@article{CRGEOS_2019__351_2-3_147_0,
     author = {Kenji Ohta and Sho Suehiro and Kei Hirose and Yasuo Ohishi},
     title = {Electrical resistivity of \protect\emph{fcc} phase iron hydrides at high pressures and temperatures},
     journal = {Comptes Rendus. G\'eoscience},
     pages = {147--153},
     publisher = {Elsevier},
     volume = {351},
     number = {2-3},
     year = {2019},
     doi = {10.1016/j.crte.2018.05.004},
     language = {en},
}
TY  - JOUR
AU  - Kenji Ohta
AU  - Sho Suehiro
AU  - Kei Hirose
AU  - Yasuo Ohishi
TI  - Electrical resistivity of fcc phase iron hydrides at high pressures and temperatures
JO  - Comptes Rendus. Géoscience
PY  - 2019
SP  - 147
EP  - 153
VL  - 351
IS  - 2-3
PB  - Elsevier
DO  - 10.1016/j.crte.2018.05.004
LA  - en
ID  - CRGEOS_2019__351_2-3_147_0
ER  - 
%0 Journal Article
%A Kenji Ohta
%A Sho Suehiro
%A Kei Hirose
%A Yasuo Ohishi
%T Electrical resistivity of fcc phase iron hydrides at high pressures and temperatures
%J Comptes Rendus. Géoscience
%D 2019
%P 147-153
%V 351
%N 2-3
%I Elsevier
%R 10.1016/j.crte.2018.05.004
%G en
%F CRGEOS_2019__351_2-3_147_0
Kenji Ohta; Sho Suehiro; Kei Hirose; Yasuo Ohishi. Electrical resistivity of fcc phase iron hydrides at high pressures and temperatures. Comptes Rendus. Géoscience, Volume 351 (2019) no. 2-3, pp. 147-153. doi : 10.1016/j.crte.2018.05.004. https://comptes-rendus.academie-sciences.fr/geoscience/articles/10.1016/j.crte.2018.05.004/

Version originale du texte intégral

1 Introduction

The Earth's core is not only composed of iron–nickel (Fe–Ni) alloy, but also needs other lighter impurities to explain the 8–10% density deficit with respect to pure Fe (e.g., Poirier, 1994). Hydrogen as a light element in the Earth's core has been a focus of interest for more than six decades (Badding et al., 1991; Birch, 1952; Fukai and Akimoto, 1983; Iizuka-Oku et al., 2017; Mao et al., 2004; Shibazaki et al., 2012; Stevenson, 1977). Therefore, the effects of hydrogen on crystal structures of Fe and the compressibility of FeHx under high pressure have been extensively studied. Stoichiometric FeHx (x ∼ 1.0) has been confirmed to show three types of high-pressure polymorph: body-centered cubic (bcc), double hexagonal close-packed (dhcp), and face-centered cubic (fcc) phases of FeHx (e.g., Sakamaki et al., 2009). FeH2, FeH3, and FeH5 were recently discovered at high pressures and high hydrogen concentration conditions (Pépin et al., 2014, 2017). Because of hydrogen's low atomic weight, about 1 wt.% (36 at.%) of hydrogen in Fe would be enough to account for the density deficit in the core (Poirier, 1994; Umemoto and Hirose, 2015). The required hydrogen content in the core becomes smaller in the case of Fe hydride with the other candidate light elements such as Si and S (Shibazaki et al., 2011; Tagawa et al., 2016; Terasaki et al., 2011).

The electrical and thermal conductivity of the Earth's core is crucial for understanding the cooling history and thermal evolution of the Earth's interior (e.g., Labrosse, 2015). Recent studies reported that the conductivity of pure Fe is much higher than we have ever thought (de Koker et al., 2012; Gomi et al., 2013; Ohta et al., 2016; Pozzo et al., 2012), but contrasting results exist (Konôpková et al., 2016; Xu et al., 2018). Light elements and dilute Ni in the core act as additional electron scatterers in the Fe alloy, and thus reduce the conductivity of the host metal. Impurity effects of Si, O, S, C, and Ni on Fe conductivity have already been investigated at high pressures (de Koker et al., 2012; Gomi and Hirose, 2015; Gomi et al., 2013, 2016; Pozzo et al., 2012; Seagle et al., 2013; Secco, 2017; Suehiro et al., 2017; Wagle et al., 2018; Zhang et al., 2018). However, the effect of hydrogen solution on the conductivity of Fe and Fe–Ni alloy remains unknown.

In this study, we measured high-pressure (P) and temperature (T) electrical resistivity (the reciprocal of electrical conductivity) values of fcc FeHx up to 65 GPa and fcc FeNiHx at 47 GPa in a laser-heated diamond anvil cell (LHDAC). The hydrogen content (x) in our samples was estimated based on synchrotron X-ray diffraction (XRD) studies. From the obtained results, we constrain the transport properties in the Fe–H system at high PT conditions. Since the impurity resistivity of hydrogen in Fe is found to be very small, hydrogen would not play an important role in the conductivity of the Earth's core.

2 Experimental methods

Fe (–Ni) hydrides with fcc structure were prepared by laser heating of Fe or Fe–Ni alloy in the presence of an internal hydrogen source in a LHDAC with 300 μm culet anvils. The host metals were the pure Fe foil (99.99% purity) and the foil of Fe containing 10 wt.% (9.56 at.%) Ni, which are the same products as those used in the literature (Gomi and Hirose, 2015; Gomi et al., 2013; Ohta et al., 2016). Hydrogen was supplied into the sample chamber by a decomposition reaction of paraffin (CnH2n+2, n > 5). This method has been used for the successful synthesis of fcc FeHx in a LHDAC (Narygina et al., 2011; Thompson et al., 2018). The paraffin also served as a pressure transmitting medium and thermal insulator against diamond anvils.

All the experiments were performed at BL10XU, SPring-8. A monochromatic X-ray beam with an energy of 30 keV was directed at the sample. Angle-dispersive XRD spectra were acquired on a flat-panel detector (PerkinElmer). The sample-to-flat-panel distance, the wavelength of the incident X-ray, and the flat panel geometry were calibrated with a CeO2 standard. The high temperature was generated by using a couple of 100 W Yb-fiber lasers. We collected thermal radiation from the heated sample and analyzed it to convert the temperature (Ohishi et al., 2008). The pressures were determined at 300 K from a spectrum of ruby fluorescence (Dorogokupets and Oganov, 2007) for FeHx alloys and the diamond Raman edge method (Akahama and Kawamura, 2006) for FeNiHx alloy.

Methods for the high PT electrical resistivity measurements in a LHDAC are similar to our previous studies (Ohta et al., 2016; Suehiro et al., 2017). The Fe and Fe–10 wt.% Ni foils were shaped into the sample with four electrodes by using a focused-ion-beam (FIB) apparatus. The shaped metal foil sample and the paraffin were loaded into a sample chamber at the center of an insulated gasket consisting of Re and cBN. Four electrical leads made of Pt were connected to each Fe (or Fe–Ni) electrode outside the sample chamber. The electrical resistance of the sample was measured by the four-terminal method using a SourceMeter (Keithley 2450) under a constant direct current of 100 mA. Since temperature heterogeneity in a laser-heated area of the sample generates thermo-electric power, we measured the voltages of samples by passing direct current from both current directions and averaging these two voltage values to eliminate the effect of thermo-electric power.

We use an empirical method proposed by Fukai (1992) to estimate hydrogen content (x) in the prepared FeHx and FeNiHx:

x=VMHxVMΔVH,(1)
in which VMHx and VM are the lattice volume per number of metal atoms of the metal hydride and the host metal, respectively, and ΔVH is a predetermined volume expansion due to an interstitial hydrogen. In this study, Vfcc FeHx and VfccFeNiHx at each pressure were directly obtained by the XRD experiments. We used the PVT equation of state (EoS) of fcc Fe reported by Komabayashi (2014) for Vfcc Fe. The atomic volume of fcc Fe–10 wt.% Ni (Vfcc FeNi) was estimated using the thermoelastic parameters reported by Funamori et al. (1996) and Basinski et al. (1955). We assume that ΔVH is the same as ΔVD = 2.21 Å3, which was determined from the recent neutron diffraction experiment for fcc FeDx (Machida et al., 2014).

3 Results

We first synthesized the Fe (–Ni) hydrides at around 50–60 GPa and 1700 K for all runs that were monitored by XRD (Fig. 1). The obtained XRD patterns indicate that the hcp Fe sample and paraffin reacted and formed Fe carbides (Fe3C and Fe7C3) soon after the beginning of laser heating, and then the Fe carbides disappeared, and fcc FeHx and the diamond formed in one hour. Our experiments perfectly reproduced the previous study of fcc FeHx formation (Narygina et al., 2011). Incorporation of hydrogen expands the Fe lattice so that the lattice volume of fcc FeHx is much larger than that of fcc Fe at equivalent conditions.

Fig. 1

Representative XRD patterns obtained before and after laser heating for an Fe sample and paraffin pressure medium in the first run. (Upper pattern) The diffraction lines of hcp Fe obtained at 65 GPa before laser heating. (Lower) XRD peaks from fcc FeHx and diamond, which are reaction products of Fe and paraffin after laser heating. The calculated XRD peak positions of fcc Fe at 65 GPa and 300 K are shown by blue bars.

After the formation of Fe (–Ni) hydrides at around 50–60 GPa, we changed the pressures to perform electrical resistivity measurements on fcc FeHx at 34, 39, and 65 GPa and on fcc FeNiHx at 47 GPa, respectively. We did not conduct the resistivity measurements at higher pressures to avoid the occurrence of the Fe polyhydrides, although the stable P–T conditions of the Fe polyhydrides remain unknown (Pépin et al., 2014, 2017). The results of the resistivity measurements are shown in Fig. 2. For comparison, we also plot the resistivity data of Fe at 26 GPa up to 2610 K (Ohta et al., 2016) and the calculated resistivity values of Fe and Fe–10 wt.% Ni alloy based on the model proposed by Gomi and Hirose (2015). This model considers the resistivity saturation phenomenon and Matthiessen's rule for Ni impurity as follows. The resistivity saturation is expressed by the shunt resistor model (Wiesmann et al., 1977):

1ρtotalV,T=1ρidealV,T+1ρsatV,(2)
where ρsat(V) is the volume-dependent saturation resistivity (Gomi et al., 2013),
ρsatV=168VV013μΩ×cm(3)
and ρideal(V,T) is the ideal resistivity described by the Matthiessen rule and the Bloch–Grüneisen law:
ρidealV,T=ρFeV,T+ixiρiV(4)
where ρi(V) and χi are impurity resistivity and its concentration, and ρFe(V,T) is the F resistivity computed from the Bloch–Grüneisen law:
ρFeV,T=BVTΘDV50ΘDVTz5dzexpz11expz,(5)

Fig. 2

High-T electrical resistivity values of (a) fcc FeH0.77 at 34 GPa, (b) fcc FeH0.51 at 39 GPa, (c) fcc FeH0.93 at 65 GPa, and (d) fcc FeNiH0.94 at 47 GPa shown as open symbols in each panel. The crosses in Fig. 2a indicate the reported resistivity data of Fe at 26 GPa (Ohta et al., 2016). Black lines with gray band are the calculated resistivity values of hcp Fe and hcp Fe–10 wt.% Ni alloy using the model proposed in the literature (Eq. (2)–(5)) (Gomi et al., 2013; Gomi and Hirose, 2015). Black broken lines are for fcc Fe and fcc Fe–10 wt.% Ni alloy. The temperatures of hcp to fcc phase transition at each pressure are taken from Anzellini et al. (2013) for Fe and from Komabayashi et al. (2012) for the Fe–10 wt.% Ni alloy.

where ΘD(V) is the Debye temperature and B(V) is a material constant. The impurity resistivity of Ni has been determined using the following equation (Gomi and Hirose, 2015):

ρNiV=72503.51VV08.06µΩ×cm/at.%.(6)

The EoS of hcp Fe (Dewaele et al., 2006) is used to calculate the values of V and ΘD(V). We assume a twofold increase of the ρFe(V,T) due to the hcp–fcc transition as in 26 GPa data (Fig. 2a).

In the first run at 65 GPa, the resistance of the synthesized fcc FeHx sample increased by 2% compared to that of hcp Fe, which was measured before heating at 300 K. The hydrogen content of this sample was calculated as x = 0.93 using Eq. (1). Considering the increase in the resistance and the volume expansion due to the fcc FeH0.93 formation, we converted the high-P/room-T electrical resistivity of hcp Fe determined by Gomi et al. (2013) to high-P/room-T resistivity of the fcc FeH0.93 at 65 GPa. High-P/high-T resistivity of the fcc FeH0.93 was calculated from the ratio of resistance measured at high temperature up to 1770 K to that measured at room temperature, multiplied by the high-P/room-T resistivity of the fcc FeH0.93 (Fig. 2c). As shown in Fig. 2c, the resistivity of fcc FeH0.93 is 5–18% higher than that of hcp Fe, but they are comparable within the experimental uncertainty.

According to the previous studies, the hydrogen solubility in fcc Fe and the fcc Fe–Ni alloy decreases with increasing the temperature in a certain pressure range (Hiroi et al., 2005; Shibazaki et al., 2014). Therefore, we expected to be able to vary the hydrogen content in the pre-synthesized FeH0.93 to make the pressure and temperature conditions deviate from those in the pre-synthesis situation. Thus, we decompressed the same sample to 34 GPa, and annealed it at about 1700 K. After annealing, the hydrogen content decreased to x = 0.77 as we had expected, implying that the process of hydrogen dissolution into fcc Fe is exothermic (x decreases with increasing T) at 34 GPa. We determined the resistivity values of fcc FeH0.77 at 34 GPa up to 1690 K (Fig. 2a). These values are intermediate between those of hcp and fcc Fe in the same PT conditions.

The resistivity of the fcc FeHx was also examined at 39 GPa up to 2100 K in a different run (Fig. 2b). We synthesized the fcc FeHx at 48 GPa and 1750 K in 1 hour, and then decreased the pressure to 39 GPa and annealed it. We found that this sample's x value was 0.51 after annealing. The obtained resistivity of fcc FeH0.51 is very similar to that of fcc Fe, conversely to hcp Fe, which is different from the case for fcc FeHx, with a higher hydrogen concentration, as shown in the first run (Fig. 2a, c).

A similar experiment was also carried out on an Fe–Ni-H system (Fig. 2d). fcc Fe–Ni hydride was formed at 47 GPa and around 1700 K in 1.5 h. Its hydrogen content was determined as FeNiH0.94. The reported experimental result (Gomi and Hirose, 2015) was taken as the reference value of the resistivity of Fe–10 wt.% Ni alloy at 47 GPa and 300 K. This nearly stoichiometric fcc Fe–Ni hydride showed a resistivity value similar to that of hcp Fe with 10 wt.% Ni alloy, which was calculated using the model proposed by Gomi and Hirose (2015) (Table 1).

Table 1

Experimental conditions, observed volumes, hydrogen contents, and electrical resistivities.

Table 1
Run P (GPa)a T (K) Phase Observed lattice V3) x Resistivity (μΩ·cm)
1 65 300 fcc FeHx 44.51 (6) 0.93 8.0 (+1.0/−1.6)
1690 (40) fcc FeHx 41.8 (+5.4/−8.4)
1700 (30) fcc FeHx 41.6 (+5.4/−8.3)
1720 (30) fcc FeHx 42.2 (+5.5/−8.4)
1720 (40) fcc FeHx 44.4 (+5.8/−8.9)
1730 (40) fcc FeHx 43.7 (+5.7/−8.7)
1770 (30) fcc FeHx 46.5 (+6.0/−9.3)
After decompression and annealing
34 300 fcc FeHx 46.13 (4) 0.77 12.6 (+1.3/−1.8)
1440 (30) fcc FeHx 53.3 (+5.3/−7.5)
1520 (30) fcc FeHx 60.1 (+6.0/−8.4)
1690 (30) fcc FeHx 72.0 (+7.2/−10.1)
2 48 300 fcc FeHx 44.95 (14) 0.81 9.5 (+1.1/−1.4)
After decompression and annealing
39 300 fcc FeHx 43.27 (8) 0.51 n.d.
1780 (50) fcc FeHx 76.9 (+8.5/−13.1)
1810 (50) fcc FeHx 77.9 (+8.6/−13.3)
1880 (50) fcc FeHx 80.5 (+8.9/−13.7)
2010 (160) fcc FeHx 86.2 (+9.5/−14.7)
3 47 300 fcc FeNiHx 46.21 (6) 0.94 28.1 (+1.6/−2.6)
1490 (20) fcc FeNiHx 58.3 (+3.3/−5.4)
1500 (20) fcc FeNiHx 59.0 (+3.3/−5.4)
1530 (20) fcc FeNiHx 59.9 (+3.4/−5.5)
1540 (20) fcc FeNiHx 60.3 (+3.4/−5.5)
1570 (20) fcc FeNiHx 60.6 (+3.4/−5.6)
1590 (20) fcc FeNiHx 61.5 (+3.5/−5.6)
1620 (20) fcc FeNiHx 62.4 (+3.5/−5.7)
1660 (30) fcc FeNiHx 62.3 (+3.5/−5.7)
1680 (30) fcc FeNiHx 62.8 (+3.5/−5.8)
1690 (30) fcc FeNiHx 62.0 (+3.5/−5.7)
1700 (30) fcc FeNiHx 62.3 (+3.5/−5.8)
1710 (20) fcc FeNiHx 63.7 (+3.6/−5.9)

a Pressures were determined from the ruby method in Runs 1 and 2 and from the diamond Raman method in Run 3.

4 Discussion

dhcp FeHx (x ∼ 1.0) has been confirmed to have its metallic property denoted by the resistivity increase with increasing temperature (Matsuoka et al., 2011), but it is unknown whether other phases of the Fe hydride are metallic or insulator. The present study revealed the electrical transport properties of nearly stoichiometric fcc FeHx and FeNiHx (x ∼ 1.0), and nonstoichiometric fcc FeHx (x = 0.77 and 0.51) at high pressures and high temperatures, and all the examined fcc Fe (–Ni) hydrides showed a metallic nature (Fig. 2). In addition, our results indicate that hydrogen-rich fcc FeHx (x > 0.5) becomes more conductive with increasing the hydrogen content.

The core density deficit requires about 1 wt.% (36 at.%) of hydrogen as the sole light element in the core (Poirier, 1994; Umemoto and Hirose, 2015). Therefore, in order to discuss the effect of hydrogen on the Earth's core conductivity, the impurity resistivity of the minor amount of hydrogen in Fe is important to constrain. Note, however, that the Matthiessen's rule is only valid at a low impurity concentration up to about 20 at.% in general (Gomi et al., 2016). In some metallic binary continuous solid solutions, the dependence of the resistivity on the composition is characterized by a convex parabolic curve. We predicted the resistivity in an Fe–FeH system from the present data and the form of Nordheim's rule:

ρFeHx=1xρFe+xρFeH1.0+kx1x,(7)

where ρFe and ρFeH1.0 are the electrical resistivities of Fe and FeH1.0, respectively, and k is a positive constant that depends on the alloy system. Here we assume the resistivity of fcc FeH1.0 is the same as that of hcp Fe based on our results (Fig. 2). Fig. 3 shows the predicted relation of the resistivity in a fcc Fe–fcc FeH1.0 system at 34 GPa and 1690 K with that at 39 GPa and 2100 K. The k values in Eq. (7), i.e. 61 at 34 GPa and 55 at 39 GPa, fit our experimental data. The volume-dependent impurity resistivity values of Si, Ni, S, and C (ρSi, ρNi, ρS, and ρC) have been experimentally determined (Gomi and Hirose, 2015; Gomi et al., 2016; Suehiro et al., 2017; Zhang et al., 2018). The obtained ρSi, ρNi, ρS, and ρC are 7.2, 2.9, 4.6, and 12.1 μΩ·cm/at% at 34 GPa, and are 6.8, 2.8, 4.4, and 11.4 μΩ·cm/at% at 39 GPa, respectively. We calculated the resistivity values of fcc Fe with Si, Ni, S, and C using their impurity resistivity values and Eqs. (2)–(5) (Fig. 3). The effect of a small amount of hydrogen on Fe resistivity is likely to be weaker than that of these four elements. Gomi et al. (2013) also predicted the impurity resistivity of O in Fe (ρO) based on the Norbury–Linde rule. According to this rule, the ρO is one-fourth of the ρSi, although the validity of the Norbury–Linde rule for iron alloys at high pressures is unclear. To the best of our knowledge, the strongest impurity scatterer in iron among the core light element candidates would be carbon, while the weakest one would be hydrogen in the present experimental pressures range.

Fig. 3

The electrical resistivity in an FeHx system (a) at 34 GPa and 1690 K, and (b) at 39 GPa and 2100 K. The open symbols are the present data (Fig. 2a, b). The black lines with gray band indicate the resistivity of FeHx as a function of the hydrogen content x estimated from the form of Nordheim's rule. The effects of C, Si, S, and Ni on the Fe resistivity are calculated from Matthiessen's rule and the saturation model (Eq. (2)–(5)). Orange, Fe–C alloy from Zhang et al. (2018); blue, Fe–Si alloy from Gomi et al. (2016); red, Fe–S alloy from Suehiro et al. (2017); green, Fe–Ni alloy from Gomi and Hirose (2015).

Incorporation of hydrogen significantly decreases the density and the melting temperature of Fe, even when in low quantities. Such characteristics of hydrogen appear to elucidate the density deficit of the core (e.g., Badding et al., 1991) and the low solidus temperature of the pyrolytic lowermost mantle (Nomura et al., 2014). However, the present study indicates the minor contribution of hydrogen incorporation in the core to electrical conductivity, and hence the electronic thermal conductivity. The 1 wt.% (36 at.%) hydrogen incorporation enhances the electrical resistivity of Fe a few % at 39 GPa and 2100 K (Fig. 3b). Hydrogen impurity resistivity would decrease with increasing pressure, as it does for other light elements (Gomi et al., 2016; Suehiro et al., 2017; Zhang et al., 2018), so that the influence of hydrogen in the Earth's core resistivity should be smaller. The hydrogen content in the uppermost core could increase as a result of the interaction between the core and hydrous minerals such as the δ-AlOOH, phase H of MgSiO4H2 and pyrite-type FeOOH in the mantle and subducting slab (Nishi et al., 2014, 2017; Terasaki et al., 2012), which would induce the compositional layering in the outer core (Helffrich, 2014). However, the increase in hydrogen solubility at the top of the core has a very limited influence on the core conductivity and thus the thermal evolution of the core.

Acknowledgments

We thank Dr. Takahiro Matsuoka for discussion. In situ high PT synchrotron XRD measurements were performed at BL10XU, SPring-8 (proposals No. 2014B0087, 2015A0087, and 2017B0072). This work was supported by JSPS KAKENHI, Grant Numbers 15H05827 and 17H04861.


Bibliographie

[Akahama and Kawamura, 2006] Y. Akahama; H. Kawamura Pressure calibration of diamond anvil Raman gauge to 310 GPa, J. Appl. Phys., Volume 100 (2006), p. 043516

[Anzellini et al., 2013] S. Anzellini; A. Dewaele; M. Mezouar; P. Loubeyre; G. Morard Melting of iron at Earth's inner core boundary based on fast X-ray diffraction, Science, Volume 340 (2013), pp. 464-466

[Badding et al., 1991] J.V. Badding; R.J. Hemley; H.-K. Mao High-pressure chemistry of hydrogen in metals: in situ study of iron hydride, Science, Volume 253 (1991), pp. 421-424

[Basinski et al., 1955] Z.S. Basinski; W. Hume-Rothery; A.L. Sutton The lattice expansion of iron, Proc. R. Soc. A, Volume 229 (1955), pp. 459-467

[Birch, 1952] F. Birch Elasticity and constitution of the Earth's interior, J. Geophys. Res., Volume 52 (1952), pp. 227-286

[de Koker et al., 2012] N. de Koker; G. Steinle-Neumann; V. Vlcek Electrical resistivity and thermal conductivity of liquid Fe alloys at high P and T, and heat flux in Earth's core, Proc. Natl. Acad. Sci. USA, Volume 109 (2012), pp. 4070-4073

[Dewaele et al., 2006] A. Dewaele; P. Loubeyre; F. Occelli; M. Mezouar; P. Dorogokupets; M. Torrent Quasihydrostatic equation of state of iron above 2 Mbar, Phys. Rev. Lett., Volume 97 (2006), p. 215504

[Dorogokupets and Oganov, 2007] P.I. Dorogokupets; A.R. Oganov Ruby, metals, and MgO as alternative pressure scales: a semiempirical description of shock-wave, ultrasonic, X-ray, and thermochemical data at high temperatures and pressures, Phys. Rev. B, Volume 75 (2007), p. 024115

[Fukai, 1992] Y. Fukai Some properties of the Fe–H system at high pressures and temperatures, and their implications for the Earth's core (Y. Syono; M.H. Manghnani, eds.), High-Pressure Research: Application to Earth and Planetary Sciences, TERRAPUB, Washington, DC, 1992, pp. 373-385

[Fukai and Akimoto, 1983] Y. Fukai; S. Akimoto Hydrogen in the earth's core. Experimental approach, Proc. Japan Acad. Ser. B, Volume 59 (1983), pp. 158-162

[Funamori et al., 1996] N. Funamori; T. Yagi; T. Uchida High-pressure and high-temperature in situ X-ray diffraction study of iron to above 30 GPa using MA8-type apparatus, Geophys. Res. Lett., Volume 23 (1996), pp. 953-956

[Gomi et al., 2013] H. Gomi; K. Ohta; K. Hirose; S. Labrosse; R. Caracas; M.J. Verstraete; J.W. Hernlund The high conductivity of iron and thermal evolution of the Earth's core, Phys. Earth Planet. Inter., Volume 224 (2013), pp. 88-103

[Gomi and Hirose, 2015] H. Gomi; K. Hirose Electrical resistivity and thermal conductivity of hcp Fe–Ni alloys under high pressure: Implications for thermal convection in the Earth's core, Phys. Earth Planet. Inter., Volume 247 (2015), pp. 2-10

[Gomi et al., 2016] H. Gomi; K. Hirose; H. Akai; Y. Fei Electrical resistivity of substitutionally disordered hcp Fe–Si and Fe–Ni alloys: chemically-induced resistivity saturation in the Earth's core, Earth Planet. Sci. Lett., Volume 451 (2016), pp. 51-61

[Helffrich, 2014] G. Helffrich Outer core compositional layering and constraints on core liquid transport properties, Earth Planet. Sci. Lett., Volume 391 (2014), pp. 256-262

[Hiroi et al., 2005] T. Hiroi; Y. Fukai; K. Mori The phase diagram and superabundant vacancy formation in Fe–H alloys revisited, J. Alloy Compd., Volume 404 (2005), pp. 252-255

[Iizuka-Oku et al., 2017] R. Iizuka-Oku; T. Yagi; H. Gotou; T. Okuchi; T. Hattori; A. Sano-Furukawa Hydrogenation of iron in the early stage of Earth's evolution, Nat. Commun., Volume 8 (2017), p. 14096

[Komabayashi, 2014] T. Komabayashi Thermodynamics of melting relations in the system Fe–FeO at high pressure: implications for oxygen in the Earth's core, J. Geophys. Res., Volume 119 (2014), pp. 4164-4177

[Komabayashi et al., 2012] T. Komabayashi; K. Hirose; Y. Ohishi In situ X-ray diffraction measurements of the fcc–hcp phase transition boundary of an Fe–Ni alloy in an internally heated diamond anvil cell, Phys. Chem. Miner., Volume 39 (2012), pp. 329-338

[Konôpková et al., 2016] Z. Konôpková; R. McWilliams; N. Gómez-Pérez; A. Goncharov Direct measurement of thermal conductivity in solid iron at planetary core conditions, Nature, Volume 534 (2016), pp. 99-101

[Labrosse, 2015] S. Labrosse Thermal evolution of the core with a high thermal conductivity, Phys. Earth Planet. Inter., Volume 247 (2015), pp. 36-55

[Machida et al., 2014] A. Machida; H. Saitoh; H. Sugimoto; T. Hattori; A. Sano-Furukawa; N. Endo; Y. Katayama; R. Iizuka; T. Sato; M. Matsuo; S. Orimo; K. Aoki Site occupancy of interstitial deuterium atoms in face-centred cubic iron, Nat. Commun., Volume 5 (2014), p. 5063

[Mao et al., 2004] W. Mao; W. Sturhahn; D. Heinz; H. Mao; J. Shu; R. Hemley Nuclear resonant X-ray scattering of iron hydride at high pressure, Geophys. Res. Lett., Volume 31 (2004), p. L15618

[Matsuoka et al., 2011] T. Matsuoka; N. Hirao; Y. Ohishi; K. Shimizu; A. Machida; K. Aoki Structural and electrical transport properties of FeHx under high pressures and low temperatures, High Pressure Res., Volume 31 (2011), pp. 64-67

[Narygina et al., 2011] O. Narygina; L. Dubrovinsky; C. McCammon; A. Kurnosov; I. Kantor; V. Prakapenka; N. Dubrovinskaia X-ray diffraction and Mössbauer spectroscopy study of fcc iron hydride FeH at high pressures and implications for the composition of the Earth's core, Earth Planet. Sci. Lett., Volume 307 (2011), pp. 409-414

[Nishi et al., 2014] M. Nishi; T. Irifune; J. Tsuchiya; Y. Tange; Y. Nishihara; K. Fujino; Y. Higo Stability of hydrous silicate at high pressures and water transport to the deep lower mantle, Nature Geosci., Volume 7 (2014), pp. 224-227

[Nishi et al., 2017] M. Nishi; Y. Kuwayama; J. Tsuchiya; T. Tsuchiya The pyrite-type high-pressure form of FeOOH, Nature, Volume 547 (2017), pp. 205-208

[Nomura et al., 2014] R. Nomura; K. Hirose; K. Uesugi; Y. Ohishi; A. Tsuchiyama; A. Miyake; Y. Ueno Low core-mantle boundary temperature inferred from the solidus of pyrolite, Science, Volume 343 (2014), pp. 522-525

[Ohishi et al., 2008] Y. Ohishi; N. Hirao; N. Sata; K. Hirose; M. Takata Highly intense monochromatic X-ray diffraction facility for high-pressure research at SPring-8, High Pressure Res., Volume 28 (2008), pp. 163-173

[Ohta et al., 2016] K. Ohta; Y. Kuwayama; K. Hirose; K. Shimizu; Y. Ohishi Experimental determination of the electrical resistivity of iron at Earth's core conditions, Nature, Volume 534 (2016), pp. 95-98

[Pépin et al., 2014] C. Pépin; A. Dewaele; G. Geneste; P. Loubeyre; M. Mezouar New iron hydrides under high pressure, Phys. Rev. Lett., Volume 113 (2014), p. 265504

[Pépin et al., 2017] C.M. Pépin; G. Geneste; A. Dewaele; M. Mezouar Synthesis of FeH5: a layered structure with atomic hydrogen slabs, Science, Volume 357 (2017), pp. 382-385

[Poirier, 1994] J.-P. Poirier Light elements in the Earth's outer core: a critical review, Phys. Earth Planet. Inter., Volume 85 (1994), pp. 319-337

[Pozzo et al., 2012] M. Pozzo; C. Davies; D. Gubbins; D. Alfè Thermal and electrical conductivity of iron at Earth's core conditions, Nature, Volume 485 (2012), pp. 355-358

[Sakamaki et al., 2009] K. Sakamaki; E. Takahashi; Y. Nakajima; Y. Nishihara; K. Funakoshi; T. Suzuki; Y. Fukai Melting phase relation of FeHx up to 20 GPa: Implication for the temperature of the Earth's core, Phys. Earth Planet. Inter., Volume 174 (2009), pp. 192-201

[Seagle et al., 2013] C. Seagle; E. Cottrell; Y. Fei; D. Hummer; V. Prakapenka Electrical and thermal transport properties of iron and iron-silicon alloy at high pressure, Geophys. Res. Lett., Volume 40 (2013), pp. 5377-5381

[Secco, 2017] R. Secco Thermal conductivity and Seebeck coefficient of Fe and Fe–Si alloys: implications for variable Lorenz number, Phys. Earth Planet. Inter., Volume 265 (2017), pp. 23-34

[Shibazaki et al., 2011] Y. Shibazaki; E. Ohtani; H. Terasaki; R. Tateyama; T. Sakamaki; T. Tsuchiya; K. Funakoshi Effect of hydrogen on the melting temperature of FeS at high pressure: implications for the core of Ganymede, Earth Planet. Sci. Lett., Volume 301 (2011), pp. 153-158

[Shibazaki et al., 2012] Y. Shibazaki; E. Ohtani; H. Fukui; T. Sakai; S. Kamada; D. Ishikawa; S. Tsutsui; A. Baron; N. Nishitani; N. Hirao; K. Takemura Sound velocity measurements in dhcp-FeH up to 70 GPa with inelastic X-ray scattering: Implications for the composition of the Earth's core, Earth Planet. Sci. Lett., Volume 313–314 (2012), pp. 79-85

[Shibazaki et al., 2014] Y. Shibazaki; H. Terasaki; E. Ohtani; R. Tateyama; K. Nishida; K. Funakoshi; Y. Higo High-pressure and high-temperature phase diagram for Fe0.9Ni0.1–H alloy, Phys. Earth Planet. Inter., Volume 228 (2014), pp. 192-201

[Stevenson, 1977] D.J. Stevenson Hydrogen in the Earth's core, Nature, Volume 268 (1977), pp. 130-131

[Suehiro et al., 2017] S. Suehiro; K. Ohta; K. Hirose; G. Morard; Y. Ohishi The influence of sulfur on the electrical resistivity of hcp iron: Implications for the core conductivity of Mars and Earth, Geophys. Res. Lett., Volume 44 (2017), pp. 8254-8259

[Tagawa et al., 2016] S. Tagawa; K. Ohta; K. Hirose; C. Kato; Y. Ohishi Compression of Fe–Si–H alloys to core pressures, Geophys. Res. Lett., Volume 43 (2016), pp. 3686-3692

[Terasaki et al., 2011] H. Terasaki; Y. Shibazaki; T. Sakamaki; R. Tateyama; E. Ohtani; K. Funakoshi; Y. Higo Hydrogenation of FeSi under high pressure, Am. Mineral., Volume 96 (2011), pp. 93-99

[Terasaki et al., 2012] H. Terasaki; E. Ohtani; T. Sakai; S. Kamada; H. Asanuma; Y. Shibazaki; N. Hirao; N. Sata; Y. Ohishi; T. Sakamaki; A. Suzuki; K. Funakoshi Stability of Fe–Ni hydride after the reaction between Fe–Ni alloy and hydrous phase (δ-AlOOH) up to 1.2 Mbar: possibility of H contribution to the core density deficit, Phys. Earth Planet. Inter., Volume 194–195 (2012), pp. 18-24

[Thompson et al., 2018] E. Thompson; A. Davis; W. Bi; J. Zhao; E. Alp; D. Zhang; E. Greenberg; V. Prakapenka; A. Campbell High-pressure geophysical properties of fcc phase FeHx, Geochem. Geophys. Geosys., Volume 19 (2018), pp. 305-314

[Umemoto and Hirose, 2015] K. Umemoto; K. Hirose Liquid iron-hydrogen alloys at outer core conditions by first-principles calculations, Geophys. Res. Lett., Volume 42 (2015), pp. 7513-7520

[Wagle et al., 2018] F. Wagle; G. Steinle-Neumann; N. Koker Saturation and negative temperature coefficient of electrical resistivity in liquid iron-sulfur alloys at high densities from first-principles calculations, Phys. Rev. B, Volume 97 (2018), p. 094307

[Wiesmann et al., 1977] H. Wiesmann; M. Gurvitch; H. Lutz; A. Ghosh; B. Schwarz; M. Strongin; P. Allen; J. Halley Simple model for characterizing the electrical resistivity in A-15 superconductors, Phys. Rev. Lett., Volume 38 (1977), pp. 782-785

[Xu et al., 2018] J. Xu; P. Zhang; K. Haule; J. Minar; S. Wimmer; H. Ebert; R.E. Cohen Thermal conductivity and electrical resistivity of solid iron at Earth's core conditions from first-principles, 2018 (ar Xiv:1710.03564)

[Zhang et al., 2018] C. Zhang; J.-F. Lin; Y. Liu; S. Feng; C. Jin; M. Hou; T. Yoshino Electrical resistivity of Fe–C alloy at high pressure: effects of carbon as a light element on the thermal conductivity of the Earth's core, J. Geophys. Res., Volume 123 (2018), pp. 3564-3577 | DOI


Commentaires - Politique


Ces articles pourraient vous intéresser

Resistivity saturation in liquid iron–light-element alloys at conditions of planetary cores from first principles computations

Fabian Wagle; Gerd Steinle-Neumann; Nico de Koker

C. R. Géos (2019)


Charge transport in carbon nanotubes based materials: a Kubo–Greenwood computational approach

Hiroyuki Ishii; François Triozon; Nobuhiko Kobayashi; ...

C. R. Phys (2009)