• No results found

Relative stability of phases of iron in the Earth's inner core

N/A
N/A
Protected

Academic year: 2022

Share "Relative stability of phases of iron in the Earth's inner core"

Copied!
5
0
0

Loading.... (view fulltext now)

Full text

(1)

Sergio M. Davis, 1 Anatoly B. Belonoshko, 1 Natalia V. Skorodumova, 2 and B¨orje Johansson 1, 2

1 Applied Materials Physics, Department of Materials Science and Engineering, KTH, SE-100 44 Stockholm, Sweden

2 Condensed Matter Theory Group, Department of Physics, Uppsala University, Uppsala Box 530, Sweden (Dated: August 10, 2009)

We have investigated the phase diagram of iron in a range close to the conditions inside the Earth’s inner core, using molecular dynamics with a semi–empirical, embedded atom interatomic potential.

Our calculated melting curves for the body centered cubic (bcc) and hexagonal close packed (hcp) phases of iron are in good agreement with experimental and first–principles data. Interestingly our model suggests evidence of a crossover in relative stability of these crystalline phases: the most stable phase being bcc below 315 GPa, and hcp above that point. The dependence of structural and dynamical properties of crystallized iron on the temperature, at the pressure of the inner core, were determined.

I. INTRODUCTION

The phase diagram of iron at the pressures and tem- peratures of the Earth’s core (between 4000 and 8000K, and between 330 and 360 GPa) is of considerable inter- est in geophysics, as the expected properties of the core depend strongly on both its composition and crystalline structure. A major point is to establish which of the known phases of iron is the one present at the Earth’s inner core (IC), but one for which there is little exper- imental information available. It is known that, at low temperatures and high pressures, iron is stable in the hexagonal close packed (hcp) phase, unlike the body cen- tered cubic (bcc) phase, which is dynamically unstable.

This led to consider hcp the most likely structure inside the IC, however recently it was discovered that the dy- namical instability of bcc vanishes at high temperatures, moreover, competing closely in energy with hcp. It is now known that there is a solid-solid phase transition prior to melting above 200 GPa.

Recent molecular dynamics (MD) simulations of iron at the conditions of the IC [1] have reproduced the ex- istence of the hcp-bcc phase transition, and established the stability of bcc over hcp at the conditions of the IC.

This was done by comparing the melting temperatures obtained for each solid phase, calculated from two–phase simulations.

In this paper we investigate the relative stability of bcc over hcp, using the same methodology but with a dif- ferent procedure for simulating liquid–solid coexistence.

We found that the relative stability of the two phases reverses above 315 GPa, hcp being the stable phase at higher pressures.

II. METHOD

For iron at high pressures, a highly successful semi–

empirical model of interaction, based on the embedded–

atom[2, 3] class of potentials, has been developed [4].

The particular form of the interaction energy for the whole system is

² (eV) a (˚ A) n m C

0.017306 3.471392 8.137381 4.7877 24.939

TABLE I: Parameters for the iron Sutton–Chen interatomic potential from Belonoshko et al [4].

U = ² X N

i=1

 1 2 X

j6=i

µ a r ij

n

− C ρ i

 (1)

where the first term is a pairwise interaction energy and ρ i plays the role of a local density “felt” by atom i due to all the other atoms in the system. This local density is given by

ρ i = X

j6=i

µ a r ij

m

(2)

and the values of the adjustable parameters are shown in table I. This potential, adjusted from first–principles full potential linear muffin-tin orbital (FP-LMTO) calcu- lations, reproduces many structural and dynamical prop- erties of solid [5] and liquid [6] iron at high–pressures and temperatures.

The calculation of the melting temperatures was done by freezing a liquid “matrix” containing a crystalline “nu- cleus” inside, in the microcanonical or NVE ensemble (total energy and volume of the system as a constant).

As the dynamics of the atoms in a purely microcanon-

ical MD simulation is only driven by the forces dictated

by the interatomic potential (without any artificial ex-

ternal forces introduced by thermostat algorithms), this

methodology allows the system to reach a state of co-

existence between liquid and the crystalline phase used

as a nucleus. The melting temperatures obtained in this

way, corresponding to the equilibrium temperature on

each simulation, are more reliable [7, 8] than the ones

obtained by other methods of simulating coexistence of

phases, including the two–phase method.

(2)

0 100 200 300 400 Time (ps)

5500 6000 6500 7000

T (K)

HCP+liquid BCC+liquid

0 2e+05 4e+05

U (eV)

330 335 340 345

P (GPa)

FIG. 1: Instantaneous temperature as a function of time, for bcc embedded into liquid (black line) and hcp embedded into liquid (red line). The insets show instantaneous pres- sure (above) and potential energy (below), the latter using the final value for bcc as a zero level. Both systems reach a pressure P =331.8 GPa and the potential energy difference

∆u = u bcc − u hcp = 1.55 meV/atom.

We prepared each sample by “embedding” a crystalline structure (bcc or hcp in each case) into a bigger liquid structure. This was done by placing the crystal (100 000 atoms) at the center of the liquid cell (originally having 2 million atoms) and removing from the liquid all atoms closer than a minimum distance r min to any atoms in the crystal. This minimum distance r min was chosen as 1.55

˚ A, corresponding to the highest value of r such that the radial distribution function g(r) of the liquid is still zero.

For each case we performed microcanonical MD simu- lations, using the DLPOLY [9] package, starting from a temperature sufficiently below the melting temperature at that pressure, so as to ensure that the solid nucleus grows inside the liquid instead of simply melting.

As the crystalline nucleus starts to grow inside the liq- uid “matrix”, potential energy is gradually converted into kinetic energy, thus increasing the temperature of the system until equilibrium is reached. This we observed after 400 ps of simulation, both for hcp and bcc nuclei, as shown in figure 1.

The final temperatures should correspond to the melt- ing temperature of each crystalline phase, assuming we have reached a state in which the chosen crystalline phase coexists with the liquid phase.

III. DETERMINATION OF MELTING POINTS

In order to verify this assumption (liquid–solid coexis- tence) we performed a simple “local” structure analysis.

For this we divided the simulation cell in space into 27 (3×3×3) subcells, computing a radial distribution func- tion (RDF) g i (r) for each of them. We see a mixture of clearly liquid and clearly solid RDFs (conserving the characteristic structure of either bcc or hcp intermedi-

0 1 2 3 4 5 6 7

r (Å) 0

0.5 1 1.5 2 2.5 3

g (r)

FIG. 2: Radial distribution functions (RDF) calculated lo- cally for different regions of an equilibrium configuration of bcc and liquid. There are spatial regions with clear signatures of bcc structure (note the “shoulder” at r=3.7 ˚ A), together with regions with no solid structure.

0 1 2 3 4 5 6 7

r (Å) 0

1 2 3

g (r)

FIG. 3: Radial distribution functions (RDF) calculated lo- cally for different regions of an equilibrium configuration of hcp and liquid. There are spatial regions with clear signatures of hcp structure (note the little peak at r=3.0 ˚ A), together with regions with no solid structure.

ate range order), together with mixed subcells. Some of these RDF curves are shown (superimposed, for clearer effect) in figures 2 and 3.

Interestingly, figure 3 reveals a slight amount of a different crystalline structure, face–centered cubic (fcc), which only differs from hcp in the stacking order. This is clear from the presence of the small peak around r=4.7

˚ A.

Repeating this process for different pressures, we ob- tained three points (P , T ) precisely on the melting curve for each of the crystalline phases. The bcc and hcp melt- ing curves obtained from these points are shown in figure 5, together with the Z-method results.

Although the melting temperatures are quite close to

each other, our result clearly supports the possibility of

a crossover between the bcc and hcp melting curves close

to 315 GPa, leading to the slightly superior stability of

(3)

280 300 320 340 360 P (GPa)

5000 6000 7000 8000

T (K)

BCC HCP

FIG. 4: Melting points for bcc (circles) and hcp (diamonds) computed using the Z-method.

280 300 320 340 360

P (GPa) 6200

6400 6600 6800 7000 7200 7400 7600

T (K)

BCC, embedded crystal HCP, embedded crystal BCC, Two-phase method (Ref. 4) BCC, Z-method

HCP, Z-method

FIG. 5: Melting curves for bcc (circles) and hcp (diamonds).

Error bars are much smaller than the size of the symbols. The solid (green) line corresponds to the hcp melting curve from Belonoshko et al [4], calculated using the two-phase method.

the high pressure hcp phase over bcc. It is possible that this increase in stability could be due to the formation of hcp-fcc interfaces near the solid–liquid boundaries.

IV. EFFECT OF QUENCHING RATES ON

CRYSTALLIZATION OF LIQUID AROUND A CRYSTAL SEED

Next we studied the dependency of structural and dy- namic properties of Fe crystallized from a bcc seed. For this, we used the same scheme as in the coexistence sim- ulations, this time starting from different temperatures below the melting point and letting the system crystal- lize fully (when possible). When attempting to crystallize from too high temperatures (close to the melting point from below) the system indeed never crystallizes. This

happened in our case at T =6700 K.

To assess the quality of the solid obtained, some struc- tural and dynamical properties were calculated. Figure 6 shows the Common Neighbor Analysis [10] (CNA) re- sults for the final sample at each of the temperatures considered.

6000 6100 6200 6300 6400 6500

Temperature (K) 0

10 20 30 40

Percentage (%)

CNA, 6-6-6 pairs CNA, 4-4-4 pairs CNA, 5-4-4 pairs

FIG. 6: Common neighbor analysis (CNA) statistics for seeded crystallization of iron starting from different temper- atures.

As a point of comparison, if the result were an ideal bcc crystal the expected percentages would be 42.857%

(3/7) and 57.143% (4/7) for 4-4-4 pairs and 6-6-6 pairs, respectively. Even when there is an overall decrease with temperature, the proportion of 6-6-6 pairs with respect to the sum of 6-6-6 and 4-4-4 is very closely maintained (between 56.24% and 55.78%). The increasing propor- tion of 5-4-4 pairs, usually associated with amorphous structures, is probably due to formation of defects in the growth of the structure.

To get a perspective from the point of view of diffusiv- ity of the atoms due to vacancies or collective displace- ment of grain boundaries, we calculated the mean square displacement (MSD) in an interval of 200 ps for each of the temperatures considered. The results are shown in figure 7.

There is a slight increase with temperature, but no qualitative change, which suggests there is no formation of diffusive boundaries or other sources of diffusive atoms, just the expected effect of increased thermal vibrations.

Using the proportions obtained from the CNA statis- tics (figure 6) it is possible to estimate the fraction of va- cancies on each sample. For this, two calibration curves were constructed, for 6-6-6 pairs and 4-4-4 pairs respec- tively, by introducing a controlled, small amount of va- cancies (removing atoms at random) in an ideal bcc sam- ple, equilibrating at the desired temperature and per- forming the same CNA statistics. These curves are shown in figure 8.

Following these calibration curves, the fraction of va-

cancies was determined to be always below 0.3%, as

shown in figure 9.

(4)

6000 6100 6200 6300 6400 6500 Temperature (K)

1.5 2 2.5 3 3.5

Mean Square Displacement (Angstrom2)

FIG. 7: Mean square displacement in a time interval of 200 ps, for seeded crystallization of iron starting from different temperatures.

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

Percentage of vacancies (%) 30

32 34 36 38 40

Proportion of 6-6-6 pairs (%)

T=6000 K T=6250 K T=6500 K

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

Percentage of vacancies (%) 24

26 28 30 32

Proportion of 4-4-4 pairs (%)

T=6000 K T=6250 K T=6500 K

FIG. 8: Calibration curves for 6-6-6 and 4-4-4 pairs as a func- tion of the fraction of vacancies, for different temperatures.

Note that this estimate does not include the fraction of thermal vacancies, as the calibration procedure already includes the temperature effects in the CNA statistics.

6000 6100 6200 6300 6400 6500

Temperature (K) 0.1

0.15 0.2 0.25

Percentage of vacancies (%)

FIG. 9: Estimated fraction of vacancies of the crystallized sample as a function of the starting temperature.

V. CONCLUDING REMARKS

We have determined the melting point of the hcp and bcc phases of iron by coexistence simulations, for large samples, consisting of 2 million atoms. The results show agreement with previous two-phase simulations and Z- method simulations for smaller system sizes, and suggest the possibility of a crossover of stability between the hcp and bcc phases at the conditions of the Earth’s inner core. Estimations of the amount of vacancies remaining when producing samples by crystallizing liquid using a bcc seed were obtained, indicating the expected trend of increase as the final equilibration temperature of the crystallization procedure increased.

VI. ACKNOWLEDGMENTS

Computations were performed at the National Super- computing Center in Link¨oping and the Parallel Com- puter Center in Stockholm. The study was supported by the Swedish Research Council (VR) and the Founda- tion for Strategic Research (SSF). SD gratefully acknowl- edges the facilitation of the CNA software implemented by Zhuo Bao.

[1] A. B. Belonoshko, R. Ahuja, and B. Johansson, Nature 424, 1032 (2003).

[2] M. W. Finnis and J. E. Sinclair, Philos. Mag. A 50, 45

(1984).

[3] A. P. Sutton and J. Chen, Philos. Mag. Lett. 61, 139

(1990).

(5)

[4] A. B. Belonoshko, R. Ahuja, and B. Johansson, Phys.

Rev. Lett. 84, 3638 (2000).

[5] A. B. Belonoshko, N. V. Skorodumova, S. Davis, A. N.

Osiptsov, A. Rosengren, and B. Johansson, Science 316, 1603 (2007).

[6] L. Koci, A. B. Belonoshko, and R. Ahuja, Phys. Rev. B 73, 224113 (2006).

[7] A. B. Belonoshko, N. V. Skorodumova, A. Rosengren,

and B. Johansson, Phys. Rev. B 73, 012201 (2006).

[8] A. B. Belonoshko, S. Davis, N. V. Skorodumova, P. H.

Lundow, A. Rosengren, and B. Johansson, Phys. Rev. B 76, 064121 (2007).

[9] W. Smith, dLPOLY 3, release 3.06 (2006).

[10] J. D. Honeycutt and H. C. Andersen, J. Phys. Chem. 91,

4950 (1987).

References

Related documents

Stöden omfattar statliga lån och kreditgarantier; anstånd med skatter och avgifter; tillfälligt sänkta arbetsgivaravgifter under pandemins första fas; ökat statligt ansvar

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

Generally, a transition from primary raw materials to recycled materials, along with a change to renewable energy, are the most important actions to reduce greenhouse gas emissions

För att uppskatta den totala effekten av reformerna måste dock hänsyn tas till såväl samt- liga priseffekter som sammansättningseffekter, till följd av ökad försäljningsandel

40 Så kallad gold- plating, att gå längre än vad EU-lagstiftningen egentligen kräver, förkommer i viss utsträckning enligt underökningen Regelindikator som genomförts

The increasing availability of data and attention to services has increased the understanding of the contribution of services to innovation and productivity in

Av tabellen framgår att det behövs utförlig information om de projekt som genomförs vid instituten. Då Tillväxtanalys ska föreslå en metod som kan visa hur institutens verksamhet

Regioner med en omfattande varuproduktion hade också en tydlig tendens att ha den starkaste nedgången i bruttoregionproduktionen (BRP) under krisåret 2009. De