Next Article in Journal
Single-Objective Surrogate Models for Continuous Metaheuristics: An Overview
Previous Article in Journal
Algorithms and Methods for Designing and Scheduling Smart Manufacturing Systems Volume II
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermodynamic Properties of Liquid Fe-Mg Alloys Under Outer-Core Conditions Using First-Principles Molecular Dynamics

Institute of High-Pressure Physics, School of Physical Science and Technology, Ningbo University, Ningbo 315211, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2025, 15(16), 9065; https://doi.org/10.3390/app15169065
Submission received: 9 July 2025 / Revised: 3 August 2025 / Accepted: 10 August 2025 / Published: 18 August 2025
(This article belongs to the Section Earth Sciences)

Abstract

Magnesium (Mg) partitioning behavior between solid and liquid iron (Fe) shows that Mg slightly favors liquid Fe under conditions at the Earth’s core. This means that the Earth’s outer core may contain more Mg than previously thought. However, geophysical properties such as density (ρ) and sound velocity (VP) of liquid Fe-Mg alloys under outer-core conditions have yet to be explored. Considering that the liquid outer core includes approximately 10 wt.% light elements, here we established an equation of state (EoS) of liquid Fe-Mg alloys with Mg less than 10 wt.% to study the thermodynamic properties under outer-core conditions using first-principles molecular dynamics. The results show that adding Mg to liquid Fe will clearly reduce the ρ, isothermal bulk modulus (KT), and adiabatic bulk modulus (KS). Meanwhile, it will increase the VP. In order to access the Mg content in the outer core, the ρ and VP of liquid Fe-Mg alloys along the geotherm are compared with the preliminary reference Earth model. Assuming Mg is the only light element, the maximum content of Mg required is approximately 2.9–6.1 wt.% due to the temperature uncertainty at the inner core boundary (ICB). Further, considering the geochemical constraints (the partition coefficient between liquid and solid Fe, molten Fe-alloy, and silicate melts), the content of Mg is further constrained to below 0.5 wt.%.

1. Introduction

The early Earth was a magma ocean. With the Earth cooling, siderophile and lithophile elements began to separate, and Earth’s solidification led to the formation of the core and mantle until metal–silicate equilibrium was achieved [1]. The liquid outer core, which is the intermediate layer between the Earth’s inner core and the mantle, is primarily composed of liquid iron (Fe) [2,3]. However, seismic observations indicate that the ρ and VP of pure Fe are approximately 10% higher and 5% lower than those of the outer core, respectively [4]. Therefore, the liquid outer core may contain some impurities—light elements, such as sulfur (S), silicon (Si), oxygen (O), hydrogen (H), carbon (C), and nitrogen (N)—which reduce ρ and increase VP [5,6,7,8,9]. However, as one of the abundant elements in Earth, the role of Mg in the composition and dynamics of the outer core remains understudied.
Emerging evidence indicates that early precipitation of Mg-bearing minerals from the core could have supplied the energy required to sustain the geodynamo [10,11]. As the Earth crystallized and solidified, the majority of Mg concentrated in the lower mantle [12]. Theoretical investigations of Mg partitioning between solid and liquid Fe nevertheless reveal a modest preference for liquid Fe, yet a non-negligible fraction of Mg still enters the solid Fe state. This implies that the outer core could host approximately ~6 wt.% in earlier times and ~1 wt.% by now, with a higher Mg inventory than previously thought [13]. But the experimental studies about the partitioning of Mg between molten Fe-alloys and silicate melts at high pressure and temperature indicate that Mg incorporates less than 0.5 wt.% in the core formation process [14]. In addition, the lowermost mantle may exist in a partially molten state, which could facilitate the transfer of Mg to the outer core [15,16] to enrich the Mg content over geological time.
Meanwhile, Mg has received considerable attention as a potential light element in the Earth’s inner core [17,18]. Mg has never been seriously considered for inclusion in the inner core before, since under environmental conditions, Mg does not form an alloy with Fe at all. However, high pressure and high temperature can break this phenomenon. As long as the pressure increases to 45 GPa, Mg can be more siderophile with increasing temperature [19]. So the solid solubility of Mg in Fe alloy increases rapidly [20]. This miracle can be attributed to electronegativity differences, as Mg is dramatically changed under high pressure, which also led to the discovery of unexpected Fe-Mg compounds (FeMg4, FeMg3, FeMg2, Fe2Mg3, FeMg, and Fe2Mg) under inner core pressures [21]. A recent study shows that oxygen can reduce the enthalpy of Fe-Mg alloy and promote its stability under high pressure [22]. Given that the inner core appeared approximately one billion years ago and grew with the crystallization of the outer core [23], it is reasonable to deduce that the outer core contains the Mg element.
Although the amount of Mg in the outer core is small, it is likely to contribute an important light-element effect to liquid Fe alloy under the Earth’s outer-core conditions. Therefore, assuming Mg to be the sole light element in the outer core, we explored the effect of Mg on the physical properties of liquid Fe-Mg alloys (with Mg less than 10 wt.%) by first-principles molecular dynamics (FP-MD) simulations. This approach was adopted to disentangle the intrinsic effects of Mg concentration on thermodynamic properties and constrain Mg content in the outer core from a geophysical aspect. We firstly established an EoS of liquid Fe-Mg alloys and then investigated the effect of Mg on ρ, the coefficient of thermal expansivity (α), the Grüneisen parameter (γ), KT and KS, and VP under the Earth’s outer-core conditions. Consequently, compared with ρ and VP from the PREM, we constrained the maximum concentration of Mg in the Earth’s outer core.

2. Methods

2.1. Computational Method

We simulated the liquid Fe-Mg alloys by FP-MD based on the density functional theory (DFT) as implemented in the Vienna Ab initio Simulation Package (VASP) [24,25]. All FP-MD simulations are constrained by the canonical ensemble with a fixed number of atoms (N = 108 atoms), constant volume (V), and constant temperature (T), and constrained by a Nosé–Hoover thermostat [26,27]. The electronic configurations considered were fourteen valence electrons (3p63d74s1) for Fe and eight valence electrons (2p63s2) for Mg. The cutoff radii were 2.2 a.u. for Fe and 1.7 a.u. for Mg. To ensure the accuracy of the calculation results, we set the cutoff energy to 450 eV after the cutoff energy test (see Figure S1 in Supplementary Materials). The finite temperature effects in the electronic structure and force calculations were incorporated using the Fermi–Dirac smearing method [28]. Gamma point sampling was used.
The initial liquid Fe structure was obtained by melting face-centered, cubic phase Fe at 10,000 K for 5 ps with a time step of 0.5 fs. In order to explore the effect of Mg on molten Fe alloys, we studied five compositions: pure Fe, Fe103Mg5 with 2.06 wt.% Mg, Fe98Mg10 with 4.24 wt.% Mg, Fe93Mg15 with 6.96 wt.% Mg, and Fe88Mg20 with 8.98 wt.% Mg. The liquid Fe-Mg alloy structures were obtained by randomly replacing Fe atoms with Mg atoms to a fixed Mg component. Each Fe-Mg alloy was quenched to a target temperature of 4000, 5000, or 6000 K. The target pressures were realized by varying the simulation cell box. Here, the time step set was 1 fs. Each simulation run lasted for 10 ps. The simulation duration was long enough to ensure system stabilization (see Figure S2 in Supplementary Materials). The pressure was the statistical average of the latter 7 ps with the first 3 ps to reach equilibrium. We monitored the liquid state by mean square displacement (MSD) and radial distribution function (RDF). Finally, we collected the P-V-T data of the liquid Fe-Mg alloys to fit an EoS. The P-V-T data are listed in the Zenodo reservoir [29].

2.2. Pressure Correction

Considering the pressure underestimated from DFT calculations, we systematically corrected the raw simulated pressure according to experimental data using the DAC approach [30]. During the correction process, it was the high-pressure and high-temperature experimental data of pure Fe reported by Kuwayama et al. [30] that were used as a benchmark. The correction method employed was proposed by French et al. [31], with pressure deviation and volume satisfying the following relationship:
P ( V ) = P 0 V 0 c V χ + 1 e x p χ + 1 χ 1 V 0 c V χ
where Δ P 0 , V 0 c , and χ are parameters, which were obtained through least-squares fitting. We have presented the experimentally determined pressures (Pexp), the first-principles calculated pressures (Pcal), and the pressure deviations (∆P) for the pure Fe under various conditions in Table S1 of the Supplementary Materials. The parameters Δ P 0 , V 0 c , and χ were fitted using ∆P and V, yielding results of 19.13 GPa, 96.22 × 10−3cm3/g, and −0.5, respectively. Applying the correction function, we conducted a systematic correction of all theoretically pressures of liquid Fe-Mg alloys at different temperatures in this study. After that, the corrected pressure data (Pcorr) were used to construct an EoS. The necessity of pressure correction is explicitly stated in the Supplementary Materials.

2.3. Equation of State of Liquid Fe-Mg Alloy

Based on the corrected P-V-T data of liquid Fe-Mg alloys, we constructed their thermal EoS. The EoS applied here is according to the Murnaghan [32,33] and Mie–Grüneisen–Debye (MGD) EoS [34]. The detailed formulation of the EoS refers to previous work [35]. The parameters in the EoS were fitted according to the P-V-T data of liquid Fe and liquid Fe-Mg alloys from FP-MD simulations by least-squares algorithm. The EoS parameters are listed in Table 1. Due to the content of Mg being lower than 10 wt.%, we dealt with Fe-Mg by ideal mixing as used by Xie et al [35]. The bulk modulus K T 0 , the derivative of bulk modulus K T 0 , and volume V0 at reference temperature T0 are associated with the concentrations of Fe (XFe) and Mg (XMg). They are expressed as
K T 0 = K T 0 F e X F e + K T 0 M g X M g
K T 0 = K T 0 F e X F e + K T 0 M g X M g
and
V 0 = V 0 F e X F e + V 0 M g X M g
To assess the accuracy of the established EoS of Fe-Mg alloys, we compared the computational isothermal ρ and VP of liquid Fe with previous studies [30,34,35,36], considering the scarcity of available references about Fe-Mg alloys under outer-core conditions. Further details are discussed in Section 3.1.

3. Results

3.1. Thermodynamic Properties of Liquid Fe with Pressure Correction

In order to verify the accuracy of the EoS, we derived the ρ, KS, and VP of liquid pure Fe from the EoS obtained by pressure correction and compared them with the DAC [30] and shock-wave compression approaches [36] (see Figure 1). The comparison of uncorrected results with experimental and theoretical data is displayed in the Supplementary Materials. The ρ at 4000 and 5000 K from the EoS is consistent with the values from the DAC approaches [30]. At 5000 K, the ρ difference between the DAC and DFT approaches at the ICB is approximately 0.2 g/cm3. Our calculated ρ differs from the experimental value of Kuwayama et al. by only 0.138 g/cm3 (1.15% relative error), and the VP discrepancy is 0.04 km/s (0.44%). With the temperature increase, the ρ decreases, which is also in line with DAC results. As for the KS, it shows excellent agreement with DAC results [30] as well as VP. In addition, the consistency with DAC data ensures the reliability of the EoS. The results also illustrate that pressure is a primary factor affecting ρ and VP, while the influence of temperature on VP is quite limited.

3.2. The Effects of Mg on Liquid Fe-Mg Alloys

After obtaining the EoS, we explored the impact of Mg concentration on the thermodynamic properties of liquid Fe-Mg alloys, including KT, α, γ, and KS, as shown in Figure 2. Evidently, its effect on KT and KS cannot be overlooked. The addition of Mg can decrease both the KT and KS. Taking Fe-8.98 wt.% Mg as an example, its shortages of KT and KS compared with liquid pure Fe are −7.31% (Figure 2a) and −7.22% (Figure 2d) at 330 GPa. This implies that it is necessary to consider the impact of light element concentrations on KT and KS and include them as constraints to study the composition of the outer core. However, the concentration of Mg has a minimal impact on α (Figure 2b) and γ (Figure 2c). With the increase of Mg content in liquid Fe-Mg alloys, the α and γ increase slightly. The 8.98 wt.% Mg only results in a +2.38% increase for α. The concentration of Mg exhibits negligible influence on γ along the pressure range of the outer core.
In the following, we calculated the ρ and VP under the isothermal conditions at 4000, 5000, and 6000 K to assess the approximate range of Mg concentration in the outer core. As shown in Figure 3, the addition of Mg decreases the ρ of liquid Fe-Mg alloys notably, as well as with temperature. Since the atomic mass of Mg is much smaller than Fe, the ρ of Fe-Mg alloys is inversely proportional to the concentration of Mg. As for the impact on the isothermal VP, the increasing Mg content can indeed increase the VP at temperatures of 4000 to 6000 K. By combining the constraints of ρ and VP, we predict that the outer core requires approximately 2.06~6.96 wt.% Mg to match the PREM. Above all, the impact of Mg on ρ and VP is apparent. It can decrease the ρ of liquid Fe alloys and increase the VP, which indicates that it is probably a candidate light element in the outer core.

4. Discussions

It is well-known that the temperature in the outer core increases along a well-defined adiabatic gradient from the core mantle boundary (CMB) to the ICB. The adiabatic temperature profile, also known as the geotherm profile, is not well determined due to the challenges of extremely high pressure and temperature in the outer core. Here, the adiabatic temperature profile along the geotherm in the outer core is derived by T = T ICB ρ ρ ICB γ , where TICB is the temperature at the ICB, namely the anchor temperature, and ρICB is the ρ at the ICB. The ρ is taken from the PREM, and γ is calculated by the EoS.
There has been no work reported about the geotherm of the Earth’s outer core being predicted by Fe-Mg alloys. Hence, we first used liquid Fe to reproduce the geotherm profile and compared it with previous work to estimate the EoS used in this work (see Figure 4a). We took the two anchoring temperatures of 5000 and 5400 K at the ICB, as used in previous work [30,32,37,38]. With the anchoring temperature TICB = 5400 K, the TCMB predicted is 4229 K, in comparison with the reported 4005 K [30] and 3981 K [39]. If the temperature at the ICB is reduced to TICB = 5000 K, the temperature at the CMB is estimated to be 3916 K. It is also in agreement with previous TCMB estimates ranging from 3654 to 3829 K [30,34,40]. The coincidences here suggest that the geotherm of the Earth’s outer core predicted by the EoS is reliable.
Consequently, we considered a wider anchoring temperature range (4850–7100 K), which is widely accepted, to study the Mg effect on the prediction of the geotherm profile. The sampled four anchoring temperatures at the ICB were 7100 K [41,42], 6350 K [43,44], 5400 K [45,46], and 4850 K [47]. The temperature profiles are plotted in Figure 4b. Assuming the temperature at the ICB is 5400 K, it corresponds to TCMB = 4229 K and TCMB = 4268 K for Fe and Fe-8.98 wt.% Mg, respectively. The addition of 8.98 wt.% Mg raises the temperature at the CMB by 39 K. Therefore, if Mg is a possible light element in the outer core, the temperature profile in the outer core is higher than that predicted with pure Fe. It is evident that the impact of Mg on the temperature of the outer core is comparable to that of O [35], H [48], and C [49]. The temperature profile rises with the addition of Mg.
In order to constrain the Mg content in the outer core, the ρ and VP along the geotherm of the outer core were calculated. The ρ-P and VP-P profiles are displayed in Figure 5. Assuming the temperature at the ICB is 5400 K, it is Fe-4.24 wt.% Mg that generally reproduces the PREM density profile (see Figure 5a). In comparison with the PREM, liquid Fe–4.24 wt.% Mg alloy yields absolute deviations of 0.119 g/cm3 in ρ and 0.238 km/s in VP, corresponding to discrepancies of 1.06% and 2.53%, respectively. However, the slope of ρ-P is higher than that of the PREM. In terms of the VP, the addition of Mg indeed increases the VP. Similarly, the slope of the VP-P curve for the Fe-Mg alloys is less than that of the PREM. To match the VP profiles of the PREM, the outer core would need to contain more than 10 wt.% Mg. However, this is inconsistent with geochemical constraints (see Figure 5b). So, we conclude that a single Mg concentration cannot match the entire ρ-P or VP-P of the PREM along the geotherm. It implies that Mg is not the sole light element in the outer core. Hence, a multiple content model is required to align the ρ and VP curve of the PREM. In addition, when establishing such a model, consideration should be given to those with a lower ρ and higher VP gradient than that of the PREM. Suggested light elements such as S [50] and C [49] are the possible light elements alloyed with Fe in the outer core, assuming that Mg is a candidate light element.
We further explored the Mg concentration at the boundaries of the outer core, namely the CMB and ICB. We found a linear relationship for both the ρ and VP of the Fe-Mg liquid alloy, such that ρ X T < 0 and V p X T > 0 (X = wt.% of Mg) at the CMB and ICB (see Figure 6). The magnitude of the ρ X T contributed by Mg is 5.2–7.7 times larger than the V p X T (see Table S2 in Supplementary Materials), a disparity that markedly exceeds the signatures of other candidate light elements. For C, N, and Si, the ρ X T is 2.5–3.0 times the corresponding V p X T , whereas for H, O, and S, the ratio rises to 4.5–5.0 [40]. The results indicate that it is difficult to simultaneously satisfy the ρ at the CMB and ICB with liquid Fe-Mg alloy (see Figure 6a). Considering the uncertainty of the temperature at the ICB, the ρ constraints suggest that the concentration of Mg at the CMB is approximately 2.9–5.4 wt.%, while at the ICB, it is around 3.8~6.1 wt.%. And the Mg concentration uncertainty is 0.6 wt.% at 4850 K and 0.8 wt.% at 7100 K. Additionally, the Mg content at the ICB is higher than that at the CMB. This aligns with reports that light elements are stratified within the outer core and that their concentration increases with depth [51]. However, matching the VP of the PREM at both the top and the bottom of the outer core would demand Mg contents exceeding 10 wt.%, a value already above the total light elements content inferred for the outer core and thus implausible. Even so, Figure 6b indicates that the ICB requires a higher Mg concentration than the CMB. This discrepancy can serve as one piece of evidence supporting the notion that Mg is not the sole light element in the outer core.
However, in a multi-light-element system, the interactions among these elements cannot be neglected. For example, increasing the O concentration alters the partition coefficient of H, driving additional H into the inner core [8]. Therefore, it is necessary to discuss the composition of the outer core by considering a multi-component system composed of Mg and other candidate light elements (such as an Fe-Mg-S system or Fe-Mg-O [22,49]) to match the ρ-P or VP-P profile of the PREM. Here, we propose that the presence of Mg in the outer core is likely due to its retention after the solidification of the magma ocean. Additionally, the growth and crystallization of the inner core may have resulted in a higher Mg content at the boundary of the inner core, contributing to the higher Mg content at the ICB than that at the CMB observed here.

5. Conclusions

We employed FP-MD simulations to model the Fe-Mg alloys under outer-core conditions. Based on the corrected P-V-T data, we fit an EoS to calculate the thermodynamic properties of liquid Fe-Mg alloys. As a result, the addition of Mg will reduce ρ, KT, and KS, and increase the VP of liquid Fe-Mg alloys. Further, ρ and VP along the geotherm of the outer core are calculated to constrain the Mg content in the outer core. The addition of Mg can indeed reduce the ρ and increase the VP. Therefore, Mg is probably a candidate light element in the outer core. With TICB = 5400 K, it is Fe-4.24 wt.% Mg that could match the ρ of the PREM along the geotherm. Due to its inability to simultaneously satisfy ρ and VP, it cannot be a dominant light element. Under the assumption that Mg is the sole light element, we derived an upper limit of 2.9–6.1 wt.% Mg that could satisfy the ρ of the PREM at the ICB and CMB. The uncertainty of temperature at the CMB will result in a deviation of approximately 2.5 wt.% Mg and 2.3 wt.% Mg. But cosmochemical and geochemical constraints suggest that the actual Mg content is likely significantly lower, approximately 0.5 wt.%, which further narrows the Mg concentration in the outer core.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/app15169065/s1: Figure S1: ENCUT test results; Figure S2: Simulation time test; Figure S3: The comparison of density before and after pressure correction at 4000 K; Figure S4: Uncorrected isothermal density and sound velocity curves of liquid pure Fe; Figure S5: Uncorrected density and sound velocity of liquid Fe-Mg alloys as a function of Mg concentration at CMB and ICB; Table S1: Experimental and calculated data for pressure correction; Table S2: The slopes of density and sound velocity with respect to Mg concentration.

Author Contributions

Conceptualization, H.X., M.X. and J.F.; methodology, H.X., M.X. and J.F.; validation, H.X.; formal analysis, H.X.; investigation, H.X.; data curation, H.X.; writing—original draft preparation, H.X.; writing—review and editing, H.X. and J.F.; visualization, H.X. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by the National Natural Science Foundation of China (NSFC) (Grant No. 11804175), the Program for Science and Technology Innovation Team in Zhejiang (Grant No. 2021R01004), the Natural Science Foundation of Ningbo (Grant Nos. 2021J099 and 2024J093), and the K.C. Wong Magna Foundation at Ningbo University.

Data Availability Statement

The pressure–volume–temperature data of liquid Fe-Mg alloys used to build the equation of state are available at the Denodo online repository [29].

Acknowledgments

We gratefully acknowledge the financial support from various funding agencies. We also thank our colleagues for their technical assistance, as well as the journal editors and reviewers for their constructive comments.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Rubie, D.C.; Melosh, H.J.; Reid, J.E.; Liebske, C.; Righter, K. Mechanisms of Metal–Silicate Equilibration in the Terrestrial Magma Ocean. Earth Planet. Sci. Lett. 2003, 205, 239–255. [Google Scholar] [CrossRef]
  2. Birch, F. Elasticity and Constitution of the Earth’s Interior. J. Geophys. Res. 1952, 57, 227–286. [Google Scholar] [CrossRef]
  3. Dziewonski, A.M.; Anderson, D.L. Preliminary Reference Earth Model. Phys. Earth Planet. Inter. 1981, 25, 297–356. [Google Scholar] [CrossRef]
  4. Birch, F. Density and Composition of Mantle and Core. J. Geophys. Res. (1896–1977) 1964, 69, 4377–4388. [Google Scholar] [CrossRef]
  5. Badro, J.; Côté, A.S.; Brodholt, J.P. A Seismologically Consistent Compositional Model of Earth’s Core. Proc. Natl. Acad. Sci. USA 2014, 111, 7542–7545. [Google Scholar] [CrossRef]
  6. Umemoto, K.; Hirose, K. Chemical Compositions of the Outer Core Examined by First Principles Calculations. Earth Planet. Sci. Lett. 2020, 531, 116009. [Google Scholar] [CrossRef]
  7. Pease, A.; Liu, J.; Lv, M.; Piper, J.; Kono, Y.; Dorfman, S.M. Liquid Structure of Iron and Iron–Nitrogen–Carbon Alloys Within the Cores of Small Terrestrial Bodies. JGR Planets 2025, 130, e2024JE008599. [Google Scholar] [CrossRef]
  8. Zhang, Z.; Wang, W.; Liu, J.; Zhang, Y.; Mitchell, R.N.; Zhang, Z. Oxygen Driving Hydrogen Into the Inner Core: Implications for the Earth’s Core Composition. Geophys. Res. Lett. 2025, 52, e2024GL110315. [Google Scholar] [CrossRef]
  9. Satyal, S.; Wang, J. Structure and Non-Ideal Mixing of Fe-Ni-S Liquid at High Temperature and Pressure and Its Implication for the Earth’s Outer Core Composition. JGR Solid Earth 2024, 129, e2024JB029436. [Google Scholar] [CrossRef]
  10. O’Rourke, J.G.; Stevenson, D.J. Powering Earth’s Dynamo with Magnesium Precipitation from the Core. Nature 2016, 529, 387–389. [Google Scholar] [CrossRef]
  11. Liu, W.; Zhang, Y.; Yin, Q.-Z.; Zhao, Y.; Zhang, Z. Magnesium Partitioning between Silicate Melt and Liquid Iron Using First-Principles Molecular Dynamics: Implications for the Early Thermal History of the Earth’s Core. Earth Planet. Sci. Lett. 2020, 531, 115934. [Google Scholar] [CrossRef]
  12. Murakami, M.; Khan, A.; Sossi, P.A.; Ballmer, M.D.; Saha, P. The Composition of Earth’s Lower Mantle. Annu. Rev. Earth Planet. Sci. 2024, 52, 605–638. [Google Scholar] [CrossRef]
  13. Li, Y.; Vočadlo, L.; Alfè, D.; Brodholt, J. Mg Partitioning between Solid and Liquid Iron under the Earth’s Core Conditions. Phys. Earth Planet. Inter. 2018, 274, 218–221. [Google Scholar] [CrossRef]
  14. Pu, C.; Gao, X.; Wu, Z.; Du, Z.; Jing, Z. Metal-Silicate Partitioning of Si, O, and Mg at High Pressures and High Temperatures: Implications to the Compositional Evolution of Core-Forming Metallic Melts. Geochem. Geophys. Geosyst. 2025, 26, e2024GC011940. [Google Scholar] [CrossRef]
  15. Takafuji, N.; Hirose, K.; Mitome, M.; Bando, Y. Solubilities of O and Si in Liquid Iron in Equilibrium with (Mg,Fe)SiO3 Perovskite and the Light Elements in the Core. Geophys. Res. Lett. 2005, 32, 2005GL022773. [Google Scholar] [CrossRef]
  16. Deng, J.; Miyazaki, Y.; Lee, K.K.M. Implications for the Melting Phase Relations in the MgO-FeO System at Core-Mantle Boundary Conditions. JGR Solid Earth 2019, 124, 1294–1304. [Google Scholar] [CrossRef]
  17. Kádas, K.; Vitos, L.; Ahuja, R. Elastic Properties of Iron-Rich Hcp Fe–Mg Alloys up to Earth’s Core Pressures. Earth Planet. Sci. Lett. 2008, 271, 221–225. [Google Scholar] [CrossRef]
  18. Kádas, K.; Vitos, L.; Johansson, B.; Ahuja, R. Stability of Body-Centered Cubic Iron–Magnesium Alloys in the Earth’s Inner Core. Proc. Natl. Acad. Sci. USA 2009, 106, 15560–15562. [Google Scholar] [CrossRef]
  19. Rahm, M.; Cammi, R.; Ashcroft, N.W.; Hoffmann, R. Squeezing All Elements in the Periodic Table: Electron Configuration and Electronegativity of the Atoms under Compression. J. Am. Chem. Soc. 2019, 141, 10253–10271. [Google Scholar] [CrossRef]
  20. Wahl, S.M.; Militzer, B. High-Temperature Miscibility of Iron and Rock during Terrestrial Planet Formation. Earth Planet. Sci. Lett. 2015, 410, 25–33. [Google Scholar] [CrossRef]
  21. Gao, P.; Su, C.; Shao, S.; Wang, S.; Liu, P.; Liu, S.; Lv, J. Iron–Magnesium Compounds under High Pressure. New J. Chem. 2019, 43, 17403–17407. [Google Scholar] [CrossRef]
  22. Fang, Y.; Sun, Y.; Wang, R.; Zheng, F.; Zhang, F.; Wu, S.; Wang, C.-Z.; Wentzcovitch, R.M.; Ho, K.-M. Structural Prediction of Fe-Mg-O Compounds at Super-Earth’s Pressures. Phys. Rev. Mater. 2023, 7, 113602. [Google Scholar] [CrossRef]
  23. Gomi, H.; Ohta, K.; Hirose, K.; Labrosse, S.; Caracas, R.; Verstraete, M.J.; Hernlund, J.W. The High Conductivity of Iron and Thermal Evolution of the Earth’s Core. Phys. Earth Planet. Inter. 2013, 224, 88–103. [Google Scholar] [CrossRef]
  24. Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  25. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  26. Nosé, S. A Molecular Dynamics Method for Simulations in the Canonical Ensemble. Mol. Phys. 1984, 52, 255–268. [Google Scholar] [CrossRef]
  27. Hoover, W.G. Canonical Dynamics: Equilibrium Phase-Space Distributions. Phys. Rev. A 1985, 31, 1695–1697. [Google Scholar] [CrossRef]
  28. Mermin, N.D. Thermal Properties of the Inhomogeneous Electron Gas. Phys. Rev. 1965, 137, A1441–A1443. [Google Scholar] [CrossRef]
  29. Xu, H.; Xie, M.; Fu, J. T-P-V Data of Liquid Fe-Mg Alloys under the Outer Core Conditions by FP-MD Simulations. 2025. [Google Scholar] [CrossRef]
  30. Kuwayama, Y.; Morard, G.; Nakajima, Y.; Hirose, K.; Baron, A.Q.R.; Kawaguchi, S.I.; Tsuchiya, T.; Ishikawa, D.; Hirao, N.; Ohishi, Y. Equation of State of Liquid Iron under Extreme Conditions. Phys. Rev. Lett. 2020, 124, 165701. [Google Scholar] [CrossRef]
  31. French, M.; Mattsson, T.R. Thermodynamically Constrained Correction to Ab Initio Equations of State. J. Appl. Phys. 2014, 116, 013510. [Google Scholar] [CrossRef]
  32. Murnaghan, F.D. The Compressibility of Media under Extreme Pressures. Proc. Natl. Acad. Sci. 1944, 30, 244–247. [Google Scholar] [CrossRef]
  33. Birch, F. Finite Strain Isotherm and Velocities for Single-crystal and Polycrystalline NaCl at High Pressures and 300°K. J. Geophys. Res. 1978, 83, 1257–1268. [Google Scholar] [CrossRef]
  34. Ichikawa, H.; Tsuchiya, T.; Tange, Y. The P-V-T Equation of State and Thermodynamic Properties of Liquid Iron. JGR Solid Earth 2014, 119, 240–252. [Google Scholar] [CrossRef]
  35. Xie, M.; Fu, J.; Belonoshko, A.B. Equation of State and Thermodynamic Properties of Liquid Fe-O in the Earth’s Outer Core. Geosci. Front. 2024, 16, 101847. [Google Scholar] [CrossRef]
  36. Brown, J.M.; McQueen, R.G. Phase Transitions, Grüneisen Parameter, and Elasticity for Shocked Iron between 77 GPa and 400 GPa. J. Geophys. Res. 1986, 91, 7485–7494. [Google Scholar] [CrossRef]
  37. Alfè, D.; Kresse, G.; Gillan, M.J. Structure and Dynamics of Liquid Iron under Earth’s Core Conditions. Phys. Rev. B 2000, 61, 132–142. [Google Scholar] [CrossRef]
  38. Komabayashi, T.; Fei, Y. Internally Consistent Thermodynamic Database for Iron to the Earth’s Core Conditions. J. Geophys. Res. 2010, 115, 2009JB006442. [Google Scholar] [CrossRef]
  39. Bajgain, S.K.; Mookherjee, M.; Dasgupta, R.; Ghosh, D.B.; Karki, B.B. Nitrogen Content in the Earth’s Outer Core. Geophys. Res. Lett. 2019, 46, 89–98. [Google Scholar] [CrossRef]
  40. Anderson, W.W.; Ahrens, T.J. An Equation of State for Liquid Iron and Implications for the Earth’s Core. J. Geophys. Res. 1994, 99, 4273–4284. [Google Scholar] [CrossRef]
  41. Belonoshko, A.B.; Lukinov, T.; Fu, J.; Zhao, J.; Davis, S.; Simak, S.I. Stabilization of Body-Centred Cubic Iron under Inner-Core Conditions. Nat. Geosci. 2017, 10, 312–316. [Google Scholar] [CrossRef]
  42. Belonoshko, A.B.; Fu, J.; Bryk, T.; Simak, S.I.; Mattesini, M. Low Viscosity of the Earth’s Inner Core. Nat. Commun. 2019, 10, 2483. [Google Scholar] [CrossRef]
  43. Alfè, D. Temperature of the Inner-Core Boundary of the Earth: Melting of Iron at High Pressure from First-Principles Coexistence Simulations. Phys. Rev. B 2009, 79, 060101. [Google Scholar] [CrossRef]
  44. Anzellini, S.; Dewaele, A.; Mezouar, M.; Loubeyre, P.; Morard, G. Melting of Iron at Earth’s Inner Core Boundary Based on Fast X-Ray Diffraction. Science 2013, 340, 464–466. [Google Scholar] [CrossRef] [PubMed]
  45. Laio, A.; Bernard, S.; Chiarotti, G.L.; Scandolo, S.; Tosatti, E. Physics of Iron at Earth’s Core Conditions. Science 2000, 287, 1027–1030. [Google Scholar] [CrossRef] [PubMed]
  46. Balugani, S.; Hernandez, J.A.; Sévelin-Radiguet, N.; Mathon, O.; Recoules, V.; Kas, J.J.; Eakins, D.E.; Doyle, H.; Ravasio, A.; Torchio, R. New Constraints on the Melting Temperature and Phase Stability of Shocked Iron up to 270 GPa Probed by Ultrafast X-Ray Absorption Spectroscopy. Phys. Rev. Lett. 2024, 133, 254101. [Google Scholar] [CrossRef]
  47. Boehler, R. Temperatures in the Earth’s Core from Melting-Point Measurements of Iron at High Static Pressures. Nature 1993, 363, 534–536. [Google Scholar] [CrossRef]
  48. Umemoto, K.; Hirose, K. Liquid Iron-hydrogen Alloys at Outer Core Conditions by First-principles Calculations. Geophys. Res. Lett. 2015, 42, 7513–7520. [Google Scholar] [CrossRef]
  49. Bajgain, S.K.; Mookherjee, M.; Dasgupta, R. Earth’s Core Could Be the Largest Terrestrial Carbon Reservoir. Commun. Earth Environ. 2021, 2, 165. [Google Scholar] [CrossRef]
  50. Umemoto, K.; Hirose, K.; Imada, S.; Nakajima, Y.; Komabayashi, T.; Tsutsui, S.; Baron, A.Q.R. Liquid Iron-sulfur Alloys at Outer Core Conditions by First-principles Calculations. Geophys. Res. Lett. 2014, 41, 6712–6717. [Google Scholar] [CrossRef]
  51. Ganguly, J. Thermodynamics of Light Elements Stratification in the Earth’s Outer Core and Implications. Earth Planet. Sci. Lett. 2025, 659, 119333. [Google Scholar] [CrossRef]
Figure 1. Isothermal density, adiabatic bulk modulus, and sound velocity curves of liquid pure Fe. (a) ρ, (b) KS, and (c) VP as a function of pressure at 4000 K (blue), 5000 K (orange), 6000 K (green) with reported data [30,36], and PREM [3].
Figure 1. Isothermal density, adiabatic bulk modulus, and sound velocity curves of liquid pure Fe. (a) ρ, (b) KS, and (c) VP as a function of pressure at 4000 K (blue), 5000 K (orange), 6000 K (green) with reported data [30,36], and PREM [3].
Applsci 15 09065 g001
Figure 2. The thermodynamic properties of liquid Fe-Mg alloys as a function of pressure at 4000 K. (a) Isothermal bulk modulus KT, (b) the coefficient of thermal expansion α, (c) Grüneisen parameter γ, and (d) adiabatic bulk modulus KS. The results for different concentrations are indicated by solid lines in green (liquid pure Fe), yellow (Fe-2.06 wt.% Mg), purple (Fe-4.24 wt.% Mg), orange (Fe-6.96 wt.% Mg), and blue (Fe-8.98 wt.% Mg). Black empty circles represent the PREM [3]. The red crosses and spheres represent the results of solid Fe-2.23 wt.% Mg (Fe0.9Mg0.1) at 7000 K [18] and 0 K [17], respectively. The shaded area highlights the range of influence of Mg content on thermodynamic properties.
Figure 2. The thermodynamic properties of liquid Fe-Mg alloys as a function of pressure at 4000 K. (a) Isothermal bulk modulus KT, (b) the coefficient of thermal expansion α, (c) Grüneisen parameter γ, and (d) adiabatic bulk modulus KS. The results for different concentrations are indicated by solid lines in green (liquid pure Fe), yellow (Fe-2.06 wt.% Mg), purple (Fe-4.24 wt.% Mg), orange (Fe-6.96 wt.% Mg), and blue (Fe-8.98 wt.% Mg). Black empty circles represent the PREM [3]. The red crosses and spheres represent the results of solid Fe-2.23 wt.% Mg (Fe0.9Mg0.1) at 7000 K [18] and 0 K [17], respectively. The shaded area highlights the range of influence of Mg content on thermodynamic properties.
Applsci 15 09065 g002
Figure 3. Isothermal density and sound velocity as a function of pressure for liquid Fe-Mg alloys from the EoS. (af) Separately show the ρ and VP at 4000~6000 K. Compared with the PREM values marked with cyan circles [3], the colored lines are the results for this work with different Mg concentrations.
Figure 3. Isothermal density and sound velocity as a function of pressure for liquid Fe-Mg alloys from the EoS. (af) Separately show the ρ and VP at 4000~6000 K. Compared with the PREM values marked with cyan circles [3], the colored lines are the results for this work with different Mg concentrations.
Applsci 15 09065 g003
Figure 4. Adiabatic temperature profile in the outer core. The results from previous research are represented with green symbols [30,32,37,38]. (a) Two anchoring temperatures at the inner core boundary of 5000 K and 5400 K. (b) Four anchoring temperatures at the ICB, TICB = 4850, 5400, 6350, and 7100 K. The adiabatic temperature profiles of liquid pure Fe (green line), Fe-2.06 wt.% Mg (yellow line), Fe-4.24 wt.% Mg (purple line), Fe-6.96 wt.% Mg (orange line), and Fe-8.98 wt.% Mg (blue line) are presented.
Figure 4. Adiabatic temperature profile in the outer core. The results from previous research are represented with green symbols [30,32,37,38]. (a) Two anchoring temperatures at the inner core boundary of 5000 K and 5400 K. (b) Four anchoring temperatures at the ICB, TICB = 4850, 5400, 6350, and 7100 K. The adiabatic temperature profiles of liquid pure Fe (green line), Fe-2.06 wt.% Mg (yellow line), Fe-4.24 wt.% Mg (purple line), Fe-6.96 wt.% Mg (orange line), and Fe-8.98 wt.% Mg (blue line) are presented.
Applsci 15 09065 g004
Figure 5. Density and sound velocity as a function of pressure along the geotherm in the outer core. (a) ρ and (b) VP. The ρ and VP distributions of the Fe-4.24 wt.% Mg (purple line), Fe-6.96 wt.% Mg (orange line), and Fe-8.98 wt.% Mg (blue line) are plotted along the geotherm profile. PREM values are marked with cyan circles [3]. The red circles and squares indicate the ρ calculated for solid Fe-2.23 wt.% Mg (Fe0.9Mg0.1) and Fe-4.60 wt.% Mg (Fe0.95Mg0.05) [17,18].
Figure 5. Density and sound velocity as a function of pressure along the geotherm in the outer core. (a) ρ and (b) VP. The ρ and VP distributions of the Fe-4.24 wt.% Mg (purple line), Fe-6.96 wt.% Mg (orange line), and Fe-8.98 wt.% Mg (blue line) are plotted along the geotherm profile. PREM values are marked with cyan circles [3]. The red circles and squares indicate the ρ calculated for solid Fe-2.23 wt.% Mg (Fe0.9Mg0.1) and Fe-4.60 wt.% Mg (Fe0.95Mg0.05) [17,18].
Applsci 15 09065 g005
Figure 6. Density and sound velocity of liquid Fe-Mg alloys as a function of Mg concentration at the CMB and ICB. (a) ρ and (b) VP. Purple solid circles represent calculated results (TICB is set to 4850, 5400, 6350, and 7100 K). Green solid balls mark the matching Mg content at the CMB or ICB.
Figure 6. Density and sound velocity of liquid Fe-Mg alloys as a function of Mg concentration at the CMB and ICB. (a) ρ and (b) VP. Purple solid circles represent calculated results (TICB is set to 4850, 5400, 6350, and 7100 K). Green solid balls mark the matching Mg content at the CMB or ICB.
Applsci 15 09065 g006
Table 1. EoS parameters of liquid Fe-Mg.
Table 1. EoS parameters of liquid Fe-Mg.
Parameter aValue
K T 0 F e (GPa)80.24
K T 0 M g (GPa)222.39
K T 0 F e 3.38
K T 0 M g 0.21
V 0 F e  (10−3 cm3/g)172.31
V 0 M g  (10−3 cm3/g)392.95
e0 (10−7/K)2.19
g 4.76
γ01.43
a Subscript zero means the values at 0 GPa and 6000 K.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xu, H.; Xie, M.; Fu, J. Thermodynamic Properties of Liquid Fe-Mg Alloys Under Outer-Core Conditions Using First-Principles Molecular Dynamics. Appl. Sci. 2025, 15, 9065. https://doi.org/10.3390/app15169065

AMA Style

Xu H, Xie M, Fu J. Thermodynamic Properties of Liquid Fe-Mg Alloys Under Outer-Core Conditions Using First-Principles Molecular Dynamics. Applied Sciences. 2025; 15(16):9065. https://doi.org/10.3390/app15169065

Chicago/Turabian Style

Xu, Hangli, Miaoxu Xie, and Jie Fu. 2025. "Thermodynamic Properties of Liquid Fe-Mg Alloys Under Outer-Core Conditions Using First-Principles Molecular Dynamics" Applied Sciences 15, no. 16: 9065. https://doi.org/10.3390/app15169065

APA Style

Xu, H., Xie, M., & Fu, J. (2025). Thermodynamic Properties of Liquid Fe-Mg Alloys Under Outer-Core Conditions Using First-Principles Molecular Dynamics. Applied Sciences, 15(16), 9065. https://doi.org/10.3390/app15169065

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