Plan
Comptes Rendus

Internal Geophysics (Physics of Earth's Interior)
Resistivity saturation in liquid iron–light-element alloys at conditions of planetary cores from first principles computations
Comptes Rendus. Géoscience, Volume 351 (2019) no. 2-3, pp. 154-162.

Résumé

We present a comprehensive analysis of electrical resistivity for liquid Fe–Si, Fe–S, and Fe–O alloys from first principles computations, covering the pressure/temperature conditions and major light element candidates inside the cores of terrestrial planets. By fitting optical conductivity with the Drude formula, we explicitly calculate the effective electron mean free path, and show that it becomes comparable to the interatomic distance for high densities and Si/S concentrations (Ioffe–Regel criterion). In approaching the Ioffe–Regel criterion, the temperature coefficient of resistivity decreases with compression for all compositions, eventually vanishes (Fe–Si), or even changes sign (Fe–S). Differences in resistivity and the degree of saturation between the iron alloys studied are explained in terms of iron–light element coordination numbers and their density dependence. Due to competing temperature and pressure effects, resistivity profiles along proposed core adiabats exhibit a small negative pressure gradient.

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crte.2018.05.002
Mots clés : Electrical conductivity, Resistivity saturation, Mean free path, Planetary interiors, Earth's core, Coordination numbers
Fabian Wagle 1 ; Gerd Steinle-Neumann 1 ; Nico de Koker 1, 2

1 Bayerisches Geoinstitut, Universität Bayreuth, 95440 Bayreuth, Germany
2 School of Geosciences, University of the Witwatersrand, Private Bag 3, 2050 Wits, South Africa
@article{CRGEOS_2019__351_2-3_154_0,
     author = {Fabian Wagle and Gerd Steinle-Neumann and Nico de Koker},
     title = {Resistivity saturation in liquid iron{\textendash}light-element alloys at conditions of planetary cores from first principles computations},
     journal = {Comptes Rendus. G\'eoscience},
     pages = {154--162},
     publisher = {Elsevier},
     volume = {351},
     number = {2-3},
     year = {2019},
     doi = {10.1016/j.crte.2018.05.002},
     language = {en},
}
TY  - JOUR
AU  - Fabian Wagle
AU  - Gerd Steinle-Neumann
AU  - Nico de Koker
TI  - Resistivity saturation in liquid iron–light-element alloys at conditions of planetary cores from first principles computations
JO  - Comptes Rendus. Géoscience
PY  - 2019
SP  - 154
EP  - 162
VL  - 351
IS  - 2-3
PB  - Elsevier
DO  - 10.1016/j.crte.2018.05.002
LA  - en
ID  - CRGEOS_2019__351_2-3_154_0
ER  - 
%0 Journal Article
%A Fabian Wagle
%A Gerd Steinle-Neumann
%A Nico de Koker
%T Resistivity saturation in liquid iron–light-element alloys at conditions of planetary cores from first principles computations
%J Comptes Rendus. Géoscience
%D 2019
%P 154-162
%V 351
%N 2-3
%I Elsevier
%R 10.1016/j.crte.2018.05.002
%G en
%F CRGEOS_2019__351_2-3_154_0
Fabian Wagle; Gerd Steinle-Neumann; Nico de Koker. Resistivity saturation in liquid iron–light-element alloys at conditions of planetary cores from first principles computations. Comptes Rendus. Géoscience, Volume 351 (2019) no. 2-3, pp. 154-162. doi : 10.1016/j.crte.2018.05.002. https://comptes-rendus.academie-sciences.fr/geoscience/articles/10.1016/j.crte.2018.05.002/

Version originale du texte intégral

1 Introduction

In the generation of Earth's magnetic field by thermo-chemical convection in the outer core (OC) (Merrill et al., 1996), the electrical resistivity ρ of the convecting liquid determines the dissipation of the magnetic field. It constrains two critical aspects of the geodynamo: (i) the timescale of dipole decay (Wicht, 2002), and therefore the stability of the magnetic field in a specific polarity, and (ii) the fluid velocity required for generating a given magnetic field strength (Davies et al., 2015).

Resistivity and the energy budget of convection depend on chemical composition, for which the current understanding inside Earth's core is based on seismic observations in combination with cosmochemical constraints. Both the solid inner core (IC) and the liquid OC are less dense than pure iron or iron–nickel alloys, which requires the presence of one or more lighter elements (Birch, 1952). Furthermore, the density discontinuity at the inner core boundary (ICB) is too large to be caused by a liquid–solid phase transition alone, suggesting that at least one light element preferentially partitions into the OC (Masters and Gubbins, 2003). Light elements need to be sufficiently abundant in the solar system and readily alloy with iron under the conditions at which the Earth differentiated, makingsilicon, oxygen and sulfur primary candidates (Poirier, 1994). In addition to density, acoustic velocities need to match those of a seismic reference model, such as PREM (Dziewonski and Anderson, 1981), which cannot be achieved by a binary alloy of one of those elements with iron (Badro et al., 2015). As the Earth is depleted in sulfur compared to chondritic composition (Dreibus and Palme, 1996), silicon and oxygen are assumed to be the major light elements in the core (Tsuno et al., 2013) with estimated concentrations of 2–3.5% Si and 2.7–5% O (Badro et al., 2015). The presence of carbon and hydrogen is likely limited to below 0.2% due to their volatility (McDonough, 2003), and consequently these elements are not considered in the present study.

While the sulfur concentration in the Earth's core has been estimated to be below 2% (Dreibus and Palme, 1996), other terrestrial planets are expected to contain substantially more sulfur. Its high solubility in iron, its high abundance in iron meteorites, the absence of a giant impact that leads to a loss of volatiles from the Earth (Halliday, 2004), and more reducing conditions during planetary differentiation (Namur et al., 2016) support models for the cores of Mercury (Hauck et al., 2013; Rivoldini et al., 2009), Mars (Lodders and Fegley, 1997), and Ganymede (Rückriemen et al., 2015) which assume sulfur to be the sole light element.

There have been considerable advances in constraining electronic transport properties in recent years, both experimentally and computationally. First principles calculations on liquid iron (de Koker et al., 2012; Pozzo et al., 2012) first predicted resistivity at conditions of the Earth's core to be substantially lower than previous estimates (Stacey and Anderson, 2001; Stacey and Loper, 2007). The results are consistent with dynamic compression experiments at high pressure (P) and temperature (T) (Keeler, 1971), while ρ is underestimated at ambient (van Zytveld, 1980) and low P conditions (Secco and Schlössin, 1989). Recently, similarly low resistivity values of hexagonal close packed iron at high P/high T have been reported by Ohta et al. (2016), using a laser-heated diamond anvil cell (DAC).

In addition to early static and dynamic compression experiments (Bridgman, 1957; Matassov, 1977), resistivity of Fe–Si alloys has been measured in large-volume press experiments at low P and high T (Kiarasi and Secco, 2015), and beyond 100 GPa in the DAC (Gomi et al., 2013; Seagle et al., 2013). High-temperature extrapolations based on DAC experiments at ambient T yield values consistent with computational results at conditions inside Earth's core.

Single-phase samples of iron alloys containing oxygen and sulfur are challenging to synthesize (Mori et al., 2017); therefore, resistivity data is scarce for compositions in these systems. A recent study circumvented this problem by calculating impurity resistivity of sulfur from measured values of Fe, Fe–Si (Gomi et al., 2013, 2016) and the ternary Fe–Si–S alloy using Matthiessen's rule (Suehiro et al., 2017).

An important assumption in fitting experimental data was that resistivity of iron and its alloys does not follow a linear trend with T, as expected from the Bloch–Grüneisen equation, but appears to saturate at high P/T conditions (Gomi et al., 2013, 2016; Ohta et al., 2016). Resistivity saturation occurs for highly resistive transition metal alloys (Gunnarsson et al., 2003), where the electron mean free path becomes comparable to the interatomic distance, which is known as the Ioffe–Regel criterion (Ioffe and Regel, 1960). First principles computations suggest a decrease in the slope of ρ(T) for hexagonal-close-packed iron at high compression (Pozzo and Alfè, 2016), approaching saturation.

With the incorporation of light elements, resistivity increases due to the additional impurity scattering mechanism. For crystalline alloys, the concentration (c) dependence of resistivity is generally described by Nordheim's rule as ρ ∝ c(1 − c) (Nordheim, 1928), but fundamental deviations from that relation have been observed in liquid and amorphous alloys (e.g., Enderby and Howe, 1968).

Faber–Ziman theory (Faber and Ziman, 1965; Ziman, 1961) relates the dependence of resistivity for liquid alloys on concentration to structural properties of the liquid, i.e. to the microscopic short-range order described by partial structure factors Sαβ(q). While qualitative features are described well by Faber–Ziman theory, there is substantial quantitative disagreement, especially for transition metal alloys. Iron and iron alloys have a complex electronic structure with partially filled d-bands at the Fermi level and are therefore not well described by Faber–Ziman theory, as far as absolute resistivity values are concerned.

We build our study on density functional theory-based molecular dynamics (DFT–MD) trajectories and resistivity results for liquid Fe, Fe–Si, and Fe–O alloys by de Koker et al. (2012), and Fe–S alloys by Wagle et al. (2018). In this paper, we consider in detail the relation between the liquid structure and computed values of ρ for binary iron alloys with the major light element candidates Si, O, and S. We do so by calculating the effective mean free path from the Drude response of optical conductivity, from which we identify different saturation mechanisms and show correlations with structural properties of the liquid. Finally, we discuss implications on conductivity profiles, and their influence on the dynamos of terrestrial planets.

2 Methods

DFT-MD simulations were performed on cells containing 128 atoms in the NVT ensemble, using the plane-wave code VASP (Kresse and Furthmüller, 1996a,b; Kresse and Hafner, 1993). Cubic cells in a volume range between 7.09 and 11.82 Å3/atom (V0)—covering the P-range of the Earth—and light element contents of 12.5 (Fe7X) and 25 at.% (Fe3X) were set up by randomly replacing iron with light element atoms (X) in molten configurations. Atomic coordinates were updated using a time step of 1 fs, and T was controlled by the Nosé thermostat (Nosé, 1984), with T ranging from 2000 K to 8000 K. At each time step, the electron density was computed using the projector-augmented-wave (PAW) method (Kresse and Joubert, 1999) with the PBE exchange-correlation functional (Perdew et al., 1996) and a plane wave cutoff energy of 400 eV. Electronic states were occupied according to Fermi–Dirac statistics at T of the thermostat. Brillouin zone sampling was restricted to the zone center. After equilibration of P, T, and the total energy (E) had been achieved (typically after a few hundred fs), the DFT–MD simulations were continued for at least 15 ps.

We observe non-vanishing slopes of the atomic mean square displacement at all simulation conditions, even for T below the liquidus at high density. This indicates that the liquid is metastable, i.e. cells are not frozen to a glass for which the mean square displacement would reach a constant value after some simulation time.

Partial radial distribution functions gαβ(r) were obtained by a time average over all equilibrated configurations from the DFT-MD trajectories. The unnormalized radial distribution function Gαβ(r) = 4πr2ϱgαβ(r), with number density ϱ, describes how the particle density of species β varies with distance to a reference particle of species α. We define the properties of the first coordination shell by Gαβ(r) up to the first minimum: the first coordination number is then represented by the integral and the interatomic distance by the median of the distribution.

For calculating electronic transport properties, at least six uncorrelated snapshots were extracted from the MD simulations (i.e. separated by time periods greater than that required for the velocity autocorrelation function to decay to zero) and Kohn–Sham wavefunctions ψk, their energy eigenvalues ϵk, and the gradient of the Hamiltonian with respect to reciprocal wave vector H/k were computed using the Abinit software package (Gonze, 1997; Gonze et al., 2009; Torrent et al., 2008). From those, optical conductivity was calculated within the Kubo–Greenwood formalism as

σ(ω)=e2Vcellk,k[f(ϵk)f(ϵk)]δ(ϵkϵkω)ψk|vˆ|ψkψk|vˆ|ψk,(1)
implemented in the conducti-module of Abinit (Recoules and Crocombette, 2005). In Eq. (1), ℏ denotes the reduced Planck constant, e the elementary charge, Vcell the cell volume, ω the frequency of the external field and vˆ=1/·H/k the velocity operator.

By fitting the Drude formula

R[σ(ω)]=σ01+(ωτ)2(2)
to the Kubo–Greenwood results for each snapshot, we extracted the DC limit of electrical resistivity ρ=σ01 and the effective relaxation time τ. We averaged σ and τ over the snapshots and took one standard deviation as uncertainty. Calculations with denser grids of 2 × 2 × 2 and 3 × 3 × 3 k-points showed that σ(ω) is sufficiently converged (to within 3%) using a single k-point (Wagle et al., 2018).

The resulting ρ(V, T) values were fit with a parallel resistor model to interpolate between results and extrapolate to conditions not investigated:

1ρ(V,T)=1ρBG(V,T)+1ρsat(V)+1ρel(T),(3)
where
ρBG=ρ0VV0a+ρ1VV0bTT0(4)
is the empirical expression used by de Koker et al. (2012) based on the Bloch–Grüneisen formula.
ρsat=cVV013(5)
is a term accounting for resistivity saturation, similar to expressions used by Gomi et al. (2013) and Suehiro et al. (2017), and
ρel=dT0T(6)
describes the effect of thermal broadening of the density of states, which, for the Fe–S system, is analyzed in more detail by Wagle et al. (2018). The fitting parameters are listed in Table 1.

Table 1

Fit parameters of the models for ρ(V, T) (Eq. (3)) for the liquid iron–light-element alloys considered. As the individual terms (Eqs. (4)–(6)) for describing resistivity results are difficult to separate, the fit is not well constrained and uncertainties in the fitting parameters exceed their values in most cases. Parameters for Fe and the Fe–S alloys have been reported in a previous study (Wagle et al., 2018).

Fe Fe7O Fe3O Fe7Si Fe3Si Fe7S Fe3S
ρ 0R [μΩ cm] 75.10 87.80 100.3 97.09 193.4 89.03 105.2
ρ 1R [μΩ cm] 21.48 11.88 7.979 4.377 76.74 12.73 12.06
a 0.792 0.645 0.717 0.607 0.465 0.389 0.124
b 1.479 1.695 1.482 1.617 2.495 1.804 2.686
c [μΩ cm] 747.2 1823 7.771 × 105 6030 254.0 2077 6609
d [μΩ cm] 1405 4202 2.356 × 106 24810 2006 2829 2910

In order to describe electrical resistivity as a function of P, suitable for comparison to experiments and for applications in planetary models, we fit a thermodynamic model to the DFT-MD results that is based on a separation of the Helmholtz energy into ideal gas, electronic, and excess terms (de Koker and Stixrude, 2009; Vlček et al., 2012). The volume dependence of the excess term is represented by Eulerian finite strain (f) with exponent n = 2 and a similarly reduced T-term (θ) with exponent m = 0.79 and expansion orders Of=3 and Oθ=2, parameters that best describe the results for liquid iron (de Koker et al., 2012). The thermodynamic properties at reference conditions are summarized in Table 2.

Table 2

Parameters of the modified thermodynamic model by de Koker and Stixrude (2009) for V0 = 11.82 Å3/atom and T0 = 2000 K (reference conditions). Values for extensive variables are per mol of formula units. Parameters for Fe and the Fe–S alloys have been reported in a previous study (Wagle et al., 2018).

Fe Fe7O Fe3O Fe7Si Fe3Si Fe7S Fe3S
P xs0 [GPa] −2.335 −1.186 5.534 0.846 5.534 0.846 5.534
K T,xs0 [GPa] 131.4 101.0 42.56 137.3 133.1 137.8 140.0
K T , xs 0 5.161 5.706 9.223 5.031 5.122 4.694 4.736
αK T,xs0 [GPa/K] 8.822 × 10−3 5.939 × 10−3 1.252 × 10−3 8.363 × 10−3 6.138 × 10−3 8.620 × 10−3 7.194 × 10−3
V 0 α K T V T , xs 0 [GPa/K] −1.563 × 10−2 −1.518 × 10−2 −2.131 × 10−2 −1.229 × 10−2 −9.111 × 10−3 −1.660 × 10−2 −1.327 × 10−2
T 0 α K T T T , xs 0 [GPa/K] −3.348 × 10−3 −1.352 × 10−3 4.956 × 10−4 −3.288 × 10−3 −1.739 × 10−3 −2.376 × 10−3 −1.808 × 10−3
V 0 2 2 α K T V 2 T , xs 0 [GPa/K] 2.840 × 10−2 3.401 × 10−2 5.741 × 10−2 2.115 × 10−2 9.670 × 10−3 5.115 × 10−2 3.534 × 10−2
C V,xs0 [J/(mol K)] 18.50 183.1 100.5 189.5 102.3 185.1 92.90
V 0 C V V T , xs 0 [kJ/(mol K)] 15.84 313.6 86.07 324.7 87.66 317.1 79.57
V 0 2 2 C V V 2 T , xs 0 [kJ/(mol K)] 2.113 × 10−2 1.626 × 10−1 1.193 × 10−1 6.850 × 10−2 −3.449 × 10−2 3.094 × 10−1 1.033 × 10−1
ζ 0 [J/(mol K2)] 3.486 32.67 14.51 27.18 12.35 30.04 12.36
ξ 0.843 0.911 0.671 0.730 0.925 1.096 1.006

3 Results and discussion

3.1 Liquid structure

The results of our structural analysis in terms of gαβ(r), interatomic distances and coordination numbers (Figs. 1 and 2) for the binary alloys in the Fe–S, Fe–O, and Fe–Si systems reveal the following key observations.

  • (i) Silicon incorporates substitutionally into liquid iron, with average nearest-neighbor distances very similar for Fe–Fe, Fe–Si, and Si–Si (Fig. 1). Fe–Si alloys are in the vicinity of the close-packing limit of a twelve-fold coordination, which does not change with P (Fig. 2), in agreement with previous results (Posner et al., 2017b). Structural data from in situ X-ray diffraction show the same trend in the low-P range (Morard et al., 2008; Sanloup et al., 2002).
  • (ii) The average Fe–O distance is substantially shorter than the Fe–Fe distance, and two oxygen atoms are typically separated by one iron atom. This difference in size has been attributed to account for the observation that oxygen preferentially partitions into the liquid over the solid (Alfè et al., 2002; Morard et al., 2014). With increasing P, the six-fold coordination of iron around oxygen at ambient conditions increases to eight-fold at core conditions. The transition of the compression mechanism from changing coordination to shortening the interatomic distance has been shown to affect diffusive transport for the Fe–O system (Posner et al., 2017a,c).
  • (iii) Fe–S alloys at high P exhibit a net S–S repulsion, as shown by the virtual absence of a typical nearest-neighbor peak for gSS(r) between 2.0 and 2.5 Å and the presence of a next-nearest neighbor maximum at ∼3.5 Å (Fig. 1a). This has been linked to a strengthening of covalent bonds between Fe and S with increasing P, resulting in high Fe–S coordination (Alfè and Gillan, 1998). Fig. 2 shows increasing Fe–S coordination with compression, suggesting sulfur to shift from interstitial-like to substitutional incorporation in the liquid. At the highest compressions, Fe–S approaches a close-packed short-range structure.

Fig. 1

Partial radial distribution functions for (a) Fe3S, (b) Fe3Si, and (c) Fe3O at 8000 K and two unit cell volumes (V0 and V/V0 = 0.6).

Fig. 2

First iron coordination numbers around an atom of species X. The vertical scatter at each volume arises from including results for different light element concentrations at different temperatures. Linear regressions with propagated uncertainty bands are shown in matching colors to guide the eye.

3.2 Resistivity saturation and electron mean free path

As a structural dependence of resistivity is expected from transport theory of liquid metals (Faber and Ziman, 1965; Ziman, 1961), it is our goal to describe the correlations between the observations made in Section 3.1 and resistivity calculated from first principles (Fig. 3).

Fig. 3

Electrical resistivity of liquid iron and various binary iron alloys as a function of temperature along isochores. The resistivity model fit for ρ(V, T) (Eq. (3)) is shown by dashed lines.

Generally, ρ increases with T and c. Larger vibrational amplitudes and shorter distances between impurity atoms lead to a decrease in the effective electron mean free path (xeff), approaching the Ioffe–Regel saturation limit, where xeff becomes comparable to the interatomic distance. However, the addition of the same mole fraction of different light elements leads to resistivity increases by different amounts, depending on volume and temperature.

For the low impurity concentrations considered (Fe7X), resistivity can be described reasonably well by a linear T-dependence above the Debye temperature (ΘD) (∼1000 K at low compression), consistent with Bloch–Grüneisen theory. With compression, ΘD increases based on the thermodynamic parameters from the DFT–MD simulation, and the values of ρ decrease. This behavior is well captured by the resistivity model (Eq. (3)).

For higher impurity concentrations (Fe3X), we find that ρ is higher (Fig. 3) and that the Bloch–Grüneisen behavior breaks down for Fe3Si and Fe3S. The temperature coefficient of resistivity (TCR) decreases with compression for all alloys, vanishes within uncertainty for the three smallest volumes of Fe3Si and changes sign for Fe3S for the smallest two volumes we consider. We have previously explained the negative TCR for Fe3S by changes in the electronic structure that are only noticeable once the liquid has reached the Ioffe–Regel condition (Wagle et al., 2018).

In order to illustrate resistivity saturation mechanisms from a semi-classical picture of electron transport, we calculate the effective electron mean free path as xeff=vFτ, where vF=(/m)·(3π2neff)1/3 is the Fermi velocity, neff = (0)/(e2τ) the effective number density of conduction electrons, and m the electron mass. Fig. 4 reveals three distinctive features.

  • (i) For ambient P volumes (V0 = 11.82 Å3/atom), xeff approaches the mean interatomic distance asymptotically with increasing T, consistent with the thermally driven saturation process found in previous studies (Mooij, 1973; Pozzo and Alfè, 2016).
  • (ii) At the lowest cell V considered (V = 7.09 Å3/atom), xeff becomes shorter than at lower compression, due to the increased density of scattering centers. At first glance, this observation appears to be inconsistent with the fact that ρ decreases with compression (Fig. 3), but can be understood in terms of electronic structure (Wagle et al., 2018).
  • (iii) With increasing light element concentration, xeff decreases significantly. This reflects the expected behavior of an increased probability of impurity-caused scattering. For the highest compression, the Ioffe–Regel condition is satisfied for Fe3Si and Fe3S as xeff becomes equal to the mean interatomic distance within uncertainty.

Fig. 4

Electron mean free path for liquid iron with 12.5 at.% (top) and 25 at.% (bottom) light element content for two cell volumes, near ambient P (left) and largest compression (right), obtained by a Drude fit to optical conductivity (Eq. (2)). The mean free path approaches the interatomic Fe–Fe distance in the alloys (median, thick gray lines) with increasing temperature, compression, and impurity concentration. For Fe3Si and Fe3S at the smallest cell volume, the Ioffe–Regel condition is satisfied.

These observations match the resistivity behavior shown in Fig. 3 and explain the temperature effect on resistivity: while xeff in pure iron and Fe–O alloys is still somewhat larger than the average spacing of scattering centers, resistivity is not yet saturated, and its T-dependence is retained. High concentrations of Si and S in the liquid iron alloy, on the other hand, drive the system into the Ioffe–Regel limit, resulting in a vanishing TCR.

A possible reason for the short mean free path of Fe–S alloys can be found in the S–S avoidance (Fig. 1) and high Fe–S coordination (Fig. 2), leading to a quasi-ordered short-range distribution of impurity scatterers in the liquid. Sulfur is the only alloying element that leads to a negative TCR. As saturation with compression is reached, secondary electronic effects at the Fermi surface are noticeable (Wagle et al., 2018). By contrast, silicon substitutes for iron in a random fashion to a good approximation, as shown by the similarity of pair distribution functions and coordination numbers compared to pure iron (Figs. 1 and 2). Impurity resistivity of oxygen is the lowest among the light elements investigated. It occupies interstitial positions between iron atoms and has the smallest scattering cross section, with little influence on total resistivity.

4 Resistivity profiles of the cores of Mars and the Earth

The significance of saturation mechanisms on resistivity profiles of planetary cores is presented in Fig. 5, where we show ρ(P) along proposed core isentropes of Mars (Fei and Bertka, 2005) and the Earth (Kamada et al., 2012). The simultaneous increase of T and P along an isentrope leads to competing effects on resistivity (Section 3.2): ρ increases with T and decreases with P, with a net effect of slowly decreasing resistivity as a function of pressure. Fe–S alloys exhibit a significantly shallower slope than pure Fe, Fe–O and Fe–Si alloys at both low and high P conditions. This results in a resistivity crossover of the corresponding silicon and sulfur concentrations at conditions of the Earth's OC, an effect that is related to the ongoing coordination change in the highly saturated Fe–S alloy. Considering that xeff is similarly shortened in both Fe–S and Fe–Si alloys (Fig. 4), the increase in Fe–S coordination with P (Fig. 2) and the effective S–S repulsion (Fig. 1) leads to more efficient impurity scattering than the randomly distributed Si atoms.

Fig. 5

Electrical resistivity profiles for various iron–light-element alloys along the proposed core isentropes of Mars (Fei and Bertka, 2005) and of the Earth (Kamada et al., 2012). The error bars shown have been calculated from the largest relative uncertainty of all resistivity datapoints of a given composition (Fig. 3). Results for Fe77.7S22.3 (Mars) and Fe80.8S19.2 (Earth) by Suehiro et al. (2017), computational results for the ternary Fe–Si–O system by Pozzo et al. (2012, 2013), as well as results from studies of various Fe–Si alloys made by Matassov (1977) (dynamic compression) and Gomi et al. (2013) (extrapolation of ambient temperature experiments) are included for comparison.

Fig. 5 contains datapoints for Fe–Si alloys, which have been directly measured in dynamic compression experiments or based on low-T static experiments. Shock data for comparable Fe–Si alloys by Matassov (1977) show somewhat larger values than our computed resistivity profiles near the core–mantle boundary (CMB) pressure. Although the temperature corresponding to these data differs from the proposed isentrope by up to 1000 K, T is expected to have a small effect on resistivity in the highly resistive systems considered. We find our results to be in excellent agreement with the extrapolation of Gomi et al. (2013) at CMB pressure, and slightly higher at ICB pressure (∼10%).

Previous computational results for ternary liquid Fe–Si–O alloys along a different core adiabat (Pozzo et al., 2013) plot between our Fe3Si and Fe3O profiles, suggesting that the presence of O in an Fe–Si alloy lowers resistivity significantly. As a consequence of the Ioffe–Regel limit, the difference in T between the different adiabats used here and that in Pozzo et al. (2013) is not expected to have a large effect on resistivity. For pure iron, without the effect of saturation, their numbers along a hotter core adiabat (TICB ≈ 6350 K) are somewhat higher than ours.

Based on an analysis of experimental data for the Fe–Si–S system, Suehiro et al. (2017) calculated substantially lower and steeper resistivity profiles along the same isentropes, and, in contrast to our results, find sulfur to cause a smaller impurity resistivity than silicon. While the underlying extrapolated resistivity data for pure Fe and Fe–Si alloys (Gomi et al., 2013, 2016) are consistent with the DFT–MD-based results by de Koker et al. (2012) at conditions of the Earth's core, it is surprising that the (inferred) conductivity profile of Fe80.8S19.2 is very close to the computational results for pure iron. However, Matthiessen's rule—which Suehiro et al. (2017) use for calculating the impurity resistivity of sulfur from pure Fe, Fe–Si, and Fe–Si–S resistivity—has been shown to break down in the saturation limit (Gomi et al., 2016), which might in part account for the discrepancy.

As resistivity of a liquid iron alloy with a large amount of light elements is saturated, and, as a consequence, the resistivity profile is virtually independent of T, it is not expected to change significantly over geological time as the planet cools. This likely remains valid across the melting point upon the formation of an inner core, for two reasons. (i) Si and S, the elements that are mostly responsible for resistivity saturation, partition almost equally between the liquid and the solid (Alfè et al., 2002). Oxygen, which preferentially remains in the liquid portion of the core, contributes only little to total impurity resistivity. (ii) The resistivity discontinuity across the melting curve is expected to become negligible at high P (Wagle and Steinle-Neumann, 2018).

The evolution of a magnetic field generated by a convection-driven dynamo is therefore—to the first order—determined by the vigor of thermo-chemical convection, changing spatial constraints and magnetic boundary conditions as the inner core grows, not by changes in resistivity.

5 Conclusions

By using first principles simulations, we compute the structural properties and the electrical resistivity of liquid iron alloys up to pressure and temperature of the Earth's inner core, with different concentrations of silicon, oxygen, and sulfur. We find correlations between short-range order and the effective electron mean free path.

At high pressure, silicon and sulfur substitute for iron in the liquid structure, and are found to substantially shorten the mean free path, which is reflected by high coordination numbers, near the close packing limit. As the mean free path becomes comparable to the mean interatomic distance, resistivity saturates. Once the system is close to saturation, resistivity becomes independent of temperature. Confirming results of previous studies, sulfur atoms prefer to be isolated from one another (Alfè and Gillan, 1998; Morard et al., 2014) and therefore distribute more evenly in liquid iron than silicon. Compared to a randomly substituted configuration, electron scattering is then more efficient, as scattering cross sections of impurity atoms are less likely to overlap. In the saturation limit at high densities and sulfur concentrations, this leads to a negative temperature coefficient of resistivity due to secondary electronic effects at the Fermi level, which cannot be observed in the Fe–Si system.

Oxygen, which is incorporated into the liquid on interstitial positions and has a considerably smaller scattering cross section, does not affect the mean free path compared to pure iron in a significant way. Correspondingly, the Fe–O system is not close to the Ioffe–Regel limit and retains a positive temperature coefficient of resistivity throughout the pressure and temperature range investigated.

Computed resistivity profiles along core adiabats for Mars and the Earth are considerably shallower than estimates based on experimental data using Matthiessen's rule. As liquid iron alloys with high concentration of sulfur or silicon are in the Ioffe–Regel limit at high densities, resistivity will be independent of temperature and is therefore unlikely to change substantially with time during core evolution.

Acknowledgments

This work was supported by Deutsche Forschungsgemeinschaft (German Science Foundation, DFG) in the Focus Program “Planetary Magnetism” (SPP 1488) with Grant STE1105/10-1 and Research Unit “Matter under Planetary Interior Conditions” (FOR 2440) with Grant STE1105/13-1. Computing and data resources for the current project were provided by the Leibniz Supercomputing Centre of the Bavarian Academy of Sciences and Humanities (www.lrz.de). Comments by two anonymous reviewers helped to significantly improve the manuscript.


Bibliographie

[Alfè and Gillan, 1998] D. Alfè; M.J. Gillan First-principles simulations of liquid Fe-S under Earth's core conditions, Phys. Rev. B, Volume 58 (1998), pp. 8248-8256 | DOI

[Alfè et al., 2002] D. Alfè; M.J. Gillan; G.D. Price Composition and temperature of the Earth's core constrained by combining ab initio calculations and seismic data, Earth Planet. Sci. Lett., Volume 195 (2002), pp. 91-98 | DOI

[Badro et al., 2015] J. Badro; J.P. Brodholt; H. Piet; J. Siebert; F.J. Ryerson Core formation and core composition from coupled geochemical and geophysical constraints, Proc. Natl. Acad. Sci. USA, Volume 112 (2015), pp. 12310-12314 | DOI

[Birch, 1952] F. Birch Elasticity and constitution of the Earth's interior, J. Geophys. Res., Volume 57 (1952), pp. 227-286 | DOI

[Bridgman, 1957] P.W. Bridgman Effects of pressure on binary alloys V fifteen alloys of metals of moderately high melting point, Proc. Am. Acad. Arts Sci., Volume 84 (1957), pp. 131-177

[Davies et al., 2015] C. Davies; M. Pozzo; D. Gubbins; D. Alfè Constraints from material properties on the dynamics and evolution of Earth's core, Nat. Geosci., Volume 8 (2015), pp. 678-685 | DOI

[Dreibus and Palme, 1996] G. Dreibus; H. Palme Cosmochemical constraints on the sulfur content in the Earth's core, Geochim. Cosmochim. Acta, Volume 60 (1996), pp. 1125-1130 | DOI

[Dziewonski and Anderson, 1981] A.M. Dziewonski; D.L. Anderson Preliminary reference Earth model, Phys. Earth Planet. Inter., Volume 25 (1981), pp. 297-356 | DOI

[Enderby and Howe, 1968] J.E. Enderby; R.A. Howe Electron transport in liquid Cu-Sn, Philos. Mag., Volume 18 (1968), pp. 923-927 | DOI

[Faber and Ziman, 1965] T.E. Faber; J.M. Ziman A theory of the electrical properties of liquid metals. III. The resistivity of binary alloys, Philos. Mag., Volume 11 (1965), pp. 153-173 | DOI

[Fei and Bertka, 2005] Y. Fei; C. Bertka The interior of Mars, Science, Volume 308 (2005), pp. 1120-1121 | DOI

[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 | DOI

[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 | DOI

[Gonze, 1997] X. Gonze Dynamical matrices. Born effective charges, dielectric permittivity tensors, and interatomic force constants from density-functional perturbation theory, Phys. Rev. B, Volume 55 (1997), pp. 10355-10368 | DOI

[Gonze et al., 2009] X. Gonze; B. Amadon; P.M. Anglade; J.M. Beuken; F. Bottin; P. Boulanger; F. Bruneval; D. Caliste; R. Caracas; M. Côté; T. Deutsch; L. Genovese; P. Ghosez; M. Giantomassi; S. Goedecker; D.R. Hamann; P. Hermet; F. Jollet; G. Jomard; S. Leroux; M. Mancini; S. Mazevet; M.J.T. Oliveira; G. Onida; Y. Pouillon; T. Rangel; G.M. Rignanese; D. Sangalli; R. Shaltaf; M. Torrent; M.J. Verstraete; G. Zerah; J.W. Zwanziger Abinit: First-principles approach to material and nanosystem properties, Comput. Phys. Commun., Volume 180 (2009), pp. 2582-2615 | DOI

[Gunnarsson et al., 2003] O. Gunnarsson; M. Calandra; J.E. Han Saturation of electrical resistivity, Rev. Mod. Phys., Volume 75 (2003), pp. 1085-1099 | DOI

[Halliday, 2004] A.N. Halliday Mixing, volatile loss and compositional change during impact-driven accretion of the Earth, Nature, Volume 427 (2004), pp. 505-509 | DOI

[Hauck et al., 2013] S.A. Hauck; J.L. Margot; S.C. Solomon; R.J. Phillips; C.L. Johnson; F.G. Lemoine; E. Mazarico; T.J. McCoy; S. Padovan; S.J. Peale; M.E. Perry; D.E. Smith; M.T. Zuber The curious case of Mercury's internal structure, J. Geophys. Res., Volume 118 (2013), pp. 1204-1220 | DOI

[Ioffe and Regel, 1960] A.F. Ioffe; A.R. Regel Non-crystalline, amorphous and liquid electronic semiconductors, Prog. Semicond., Volume 4 (1960), pp. 237-291

[Kamada et al., 2012] S. Kamada; E. Ohtani; H. Terasaki; T. Sakai; M. Miyahara; Y. Ohishi; N. Hirao Melting relationships in the Fe-Fe3S system up to the outer core conditions, Earth Planet. Sci. Lett., Volume 359 (2012), pp. 26-33 | DOI

[Keeler, 1971] R.N. Keeler Electrical conductivity of condensed media at high pressures (P. Caldirola; H. Knoepfel, eds.), Physics of High Energy Density, Academic Press, New York, 1971, pp. 106-125

[Kiarasi and Secco, 2015] S. Kiarasi; R.A. Secco Pressure-induced electrical resistivity saturation of Fe17Si, Phys. Status Solidi B, Volume 252 (2015), pp. 2034-2042 | DOI

[de Koker et al., 2012] N. de Koker; G. Steinle-Neumann; V. Vlček 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 | DOI

[de Koker and Stixrude, 2009] N. de Koker; L. Stixrude Self-consistent thermodynamic description of silicate liquids, with application to shock melting of MgO periclase and MgSiO3 perovskite, Geophys. J. Int., Volume 178 (2009), pp. 162-179 | DOI

[Kresse and Furthmüller, 1996a] G. Kresse; J. Furthmüller Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set, Comput. Mat. Sci., Volume 6 (1996), pp. 15-50 | DOI

[Kresse and Furthmüller, 1996b] G. Kresse; J. Furthmüller Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B, Volume 54 (1996), pp. 11169-11186 | DOI

[Kresse and Hafner, 1993] G. Kresse; J. Hafner Ab initio molecular dynamics for liquid metals, Phys. Rev. B, Volume 47 (1993), pp. 558-561 | DOI

[Kresse and Joubert, 1999] G. Kresse; D. Joubert From ultrasoft pseudopotentials to the projector augmented-wave method, Phys. Rev., Volume 59 (1999), pp. 1758-1775 | DOI

[Lodders and Fegley, 1997] K. Lodders; B. Fegley An oxygen isotope model for the composition of Mars, Icarus, Volume 126 (1997), pp. 373-394 | DOI

[Masters and Gubbins, 2003] G. Masters; D. Gubbins On the resolution of density within the Earth, Phys. Earth Planet. Inter., Volume 140 (2003), pp. 159-167 | DOI

[Matassov, 1977] G. Matassov The Electrical Conductivity of Iron-Silicon Alloys at High Pressures and the Earth's Core (Ph.D. thesis), University of California, CA, USA, 1977

[McDonough, 2003] W.F. McDonough, Volume vol. 2, Elsevier, Amsterdam (2003), pp. 547-568 | DOI

[Merrill et al., 1996] R.T. Merrill; M.W. McElhinny; P.L. McFadden The Magnetic Field of the Earth: Paleo-magnetism, the Core and the Deep Mantle, Academic Press, San Diego, CA, USA, 1996

[Mooij, 1973] J.H. Mooij Electrical conduction in concentrated disordered transition metal alloys, Phys. Status Solidi A, Volume 17 (1973), pp. 521-530 | DOI

[Morard et al., 2014] G. Morard; D. Andrault; D. Antonangeli; J. Bouchet Properties of iron alloys under the Earth's core conditions, C. R. Geoscience, Volume 346 (2014), pp. 130-139 | DOI

[Morard et al., 2008] G. Morard; C. Sanloup; B. Guillot; G. Fiquet; M. Mezouar; J.P. Perrillat; G. Garbarino; K. Mibe; T. Komabayashi; K. Funakoshi In situ structural investigation of Fe-S-Si immiscible liquid system and evolution of Fe-S bond properties with pressure, J. Geophys. Res. Solid Earth, Volume 113 (2008), p. B10205 | DOI

[Mori et al., 2017] Y. Mori; H. Ozawa; K. Hirose; R. Sinmyo; S. Tateno; G. Morard; Y. Ohishi Melting experiments on Fe-Fe3S system to 254 GPa, Earth Planet. Sci. Lett., Volume 464 (2017), pp. 135-141 | DOI

[Namur et al., 2016] O. Namur; B. Charlier; F. Holtz; C. Cartier; C. McCammon Sulfur solubility in reduced mafic silicate melts: implications for the speciation and distribution of sulfur on Mercury, Earth Planet. Sci. Lett., Volume 448 (2016), pp. 102-114 | DOI

[Nordheim, 1928] L. Nordheim Über den Widerstand von Legierungen, Naturwiss, Volume 16 (1928), pp. 1042-1043 | DOI

[Nosé, 1984] S. Nosé A unified formulation of the constant temperature molecular dynamics methods, J. Chem. Phys., Volume 81 (1984), pp. 511-519 | DOI

[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 | DOI

[Perdew et al., 1996] J.P. Perdew; K. Burke; M. Ernzerhof Generalized gradient approximation made simple, Phys. Rev. Lett., Volume 77 (1996), pp. 3865-3868 | DOI

[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 | DOI

[Posner et al., 2017a] E.S. Posner; D.C. Rubie; D.J. Frost; G. Steinle-Neumann Experimental determination of oxygen diffusion in liquid iron at high pressure, Earth Planet. Sci. Lett., Volume 464 (2017), pp. 116-123 | DOI

[Posner et al., 2017b] E.S. Posner; D.C. Rubie; D.J. Frost; V. Vlček; G. Steinle-Neumann High PT experiments and first principles calculations of the diffusion of Si and Cr in liquid iron, Geochim. Cosmochim. Acta, Volume 203 (2017), pp. 323-342 | DOI

[Posner et al., 2017c] E.S. Posner; G. Steinle-Neumann; V. Vlček; D.C. Rubie Structural changes and anomalous self-diffusion of oxygen in liquid iron at high pressure, Geophys. Res. Lett., Volume 44 (2017), pp. 3526-3534 | DOI

[Pozzo and Alfè, 2016] M. Pozzo; D. Alfè Saturation of electrical resistivity of solid iron at Earth's core conditions, SpringerPlus, Volume 5 (2016), p. 256 | DOI

[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 | DOI

[Pozzo et al., 2013] M. Pozzo; C. Davies; D. Gubbins; D. Alfè Transport properties for liquid silicon-oxygen-iron mixtures at Earth's core conditions, Phys. Rev. B, Volume 87 (2013), p. 014110 | DOI

[Recoules and Crocombette, 2005] V. Recoules; J.-P. Crocombette Ab initio determination of electrical and thermal conductivity of liquid aluminum, Phys. Rev. B, Volume 72 (2005), p. 104202 | DOI

[Rivoldini et al., 2009] A. Rivoldini; T. Van Holst; O. Verhoeven The interior structure of Mercury and its core sulfur content, Icarus, Volume 201 (2009), pp. 12-30 | DOI

[Rückriemen et al., 2015] T. Rückriemen; D. Breuer; T. Spohn The Fe snow regime in Ganymede's core: a deep-seated dynamo below a stable snow zone, J. Geophys. Res., Volume 120 (2015), pp. 1095-1118 | DOI

[Sanloup et al., 2002] C. Sanloup; F. Guyot; P. Gillet; Y. Fei Physical properties of liquid Fe alloys at high pressure and their bearings on the nature of metallic planetary cores, J. Geophys. Res. Solid Earth, Volume 107 (2002), p. 2272 | DOI

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

[Secco and Schlössin, 1989] R.A. Secco; H.H. Schlössin The electrical resistivity of solid and liquid Fe at pressures up to 7 GPa, J. Geophys. Res., Volume 94 (1989), pp. 5887-5894 | DOI

[Stacey and Anderson, 2001] F.D. Stacey; O.L. Anderson Electrical and thermal conductivities of Fe-Ni-Si alloy under core conditions, Phys. Earth Planet. Inter., Volume 124 (2001), pp. 153-162 | DOI

[Stacey and Loper, 2007] F.D. Stacey; D.E. Loper A revised estimate of the conductivity of iron alloy at high pressure and implications for the core energy balance, Phys. Earth Planet. Inter., Volume 161 (2007), pp. 13-18 | DOI

[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 | DOI

[Torrent et al., 2008] M. Torrent; F. Jollet; F. Bottin; G. Zérah; X. Gonze Implementation of the projector augmented-wave method in the abinit code: application to the study of iron under pressure, Comput. Mater. Sci., Volume 42 (2008), pp. 337-351 | DOI

[Tsuno et al., 2013] K. Tsuno; D.J. Frost; D.C. Rubie Simultaneous partitioning of silicon and oxygen into the Earth's core during early Earth differentiation, Geophys. Res. Lett., Volume 40 (2013), pp. 66-71 | DOI

[Vlček et al., 2012] V. Vlček; N. de Koker; G. Steinle-Neumann Electrical and thermal conductivity of al liquid at high pressures and temperatures from ab initio computations, Phys. Rev. B, Volume 85 (2012), p. 184201 | DOI

[Wagle and Steinle-Neumann, 2018] F. Wagle; G. Steinle-Neumann Electrical resistivity discontinuity of iron along the melting curve, Geophys. J. Int., Volume 213 (2018), pp. 237-243 | DOI

[Wagle et al., 2018] F. Wagle; G. Steinle-Neumann; N. de 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 | DOI

[Wicht, 2002] J. Wicht Inner-core conductivity in numerical dynamo simulations, Phys. Earth Planet. Inter., Volume 132 (2002), pp. 281-302 | DOI

[Ziman, 1961] J.M. Ziman A theory of the electrical properties of liquid metals. I: The monovalent metals, Philos. Mag., Volume 6 (1961), pp. 1013-1034 | DOI

[van Zytveld, 1980] J.B. van Zytveld Electrical resistivities of liquid transition metals, J. Phys. (Paris), Volume 41 (1980), p. C8 | DOI


Commentaires - Politique


Ces articles pourraient vous intéresser

Electrical resistivity of fcc phase iron hydrides at high pressures and temperatures

Kenji Ohta; Sho Suehiro; Kei Hirose; ...

C. R. Géos (2019)


Electronic scattering effects in europium-based iron pnictides

Sina Zapf; David Neubauer; Kirk W. Post; ...

C. R. Phys (2016)