Next Article in Journal
Relaxation to Fixed Points in the Logistic and Cubic Maps: Analytical and Numerical Investigation
Next Article in Special Issue
Efficient Algorithms for Electrostatic Interactions Including Dielectric Contrasts
Previous Article in Journal
Development of Metrics and a Complexity Scale for the Topology of Assembly Supply Chains
Previous Article in Special Issue
Molecular Dynamics at Constant Pressure: Allowing the System to Control Volume Fluctuations via a “Shell” Particle
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Elastic Properties of CaSiO3 Perovskite from ab initio Molecular Dynamics

1
Institute for Research on Earth Evolution, Japan Agency for Marine-Earth Science and Technology, 2-15 Natsushima-cho, Yokosuka-shi, Kanagawa 237-0061, Japan
2
Earthquake Research Institute, University of Tokyo, 1-1-1 Yayoi, Bunkyo-ku, Tokyo 113-0032, Japan 
Entropy 2013, 15(10), 4300-4309; https://doi.org/10.3390/e15104300
Submission received: 26 June 2013 / Accepted: 6 October 2013 / Published: 10 October 2013
(This article belongs to the Special Issue Molecular Dynamics Simulation)

Abstract

:
Ab initio molecular dynamics simulations were performed to investigate the elasticity of cubic CaSiO3 perovskite at high pressure and temperature. All three independent elastic constants for cubic CaSiO3 perovskite, C11, C12, and C44, were calculated from the computation of stress generated by small strains. The elastic constants were used to estimate the moduli and seismic wave velocities at the high pressure and high temperature characteristic of the Earth’s interior. The dependence of temperature for sound wave velocities decreased as the pressure increased. There was little difference between the estimated compressional sound wave velocity (VP) in cubic CaSiO3 perovskite and that in the Earth’s mantle, determined by seismological data. By contrast, a significant difference between the estimated shear sound wave velocity (VS) and that in the Earth’s mantle was confirmed. The elastic properties of cubic CaSiO3 perovskite cannot explain the properties of the Earth’s lower mantle, indicating that the cubic CaSiO3 perovskite phase is a minor mineral in the Earth’s lower mantle.
PACS Codes:
62.20.de; 91.60.Gf

1. Introduction

Mineral physics constraints on the composition of the Earth’s lower mantle rely on knowledge of the equations of state (EOSs) and sound wave velocities in candidate minerals. According to reliable estimates of the composition of the Earth, an MgO-FeO-SiO2-CaO-Al2O3 system could comprise about 99% of the mantle volume [1]. Three minerals have been proposed to be possible hosts of the MgO-FeO-SiO2-CaO-Al2O3 system in the Earth’s lower mantle. A recent phase equilibrium study using a more representative composition of the mantle shows that Mg, Fe, and Al are mostly accommodated in orthorhombic (Mg,Fe)SiO3 perovskite and ferropericlase, (Mg,Fe)O. On the other hand, a number of other experimental studies indicate that the most likely Ca-bearing phase is CaSiO3 perovskite [2,3]. Thus, the Earth’s lower mantle may be composed mainly of aluminous (Mg,Fe)SiO3 perovskite, CaSiO3 perovskite, and ferropericlase. To gain an understanding of the structure and dynamics of the Earth’s lower mantle, it is important to investigate the elastic properties of these minerals under the pressure and temperature conditions found in this region. It is easy to investigate the physical properties of orthorhombic (Mg,Fe)SiO3 perovskite and ferropericlase, because both minerals can be recovered under ambient conditions. By contrast, CaSiO3 perovskite is unstable under ambient conditions, and it readily transforms to glass on the release of pressure. Therefore, it is difficult to measure some of its physical properties.
The structure of CaSiO3 perovskite has tetragonal or orthorhombic symmetry at high pressures and room temperature, e.g., [4,5]. The structure of low-symmetry CaSiO3 perovskite is still an open question [6,7,8]. A phase transformation of CaSiO3 perovskite from this low symmetry into cubic symmetry with an increase in temperature was found in a previous study [5], indicating that the cubic structure of CaSiO3 perovskite is stable under the conditions of the Earth’s lower mantle, and that its physical properties are important for understanding the dynamics and evolution of the Earth’s interior. The elastic properties of cubic CaSiO3 perovskite were calculated at 0 K [9,10] and high temperatures [8]. These data relating to the elastic properties of cubic CaSiO3 perovskite are insufficient to discuss the composition of the Earth’s lower mantle, because an internally consistent data set describing the density-VP-VS relationship in cubic CaSiO3 perovskite is needed in order to compare the density-VP-VS relationship of a preliminary reference earth model (PREM) [11] with that estimated based on cubic CaSiO3 perovskite. Therefore, it is necessary to reevaluate the density-VP-VS relationship using ab initio calculations.
We employed the ab initio molecular dynamics (AIMD) method using density functional theory (DFT) to determine the density values and sound wave velocities for cubic CaSiO3 perovskite at pressures typical of the Earth’s lower mantle. We also used the experimental data to correct the calculated values of the density and the sound wave velocities for cubic CaSiO3 perovskite.

2. Method

We performed the AIMD calculations based on DFT using the VASP code [12]. The interactions between the electrons and the ionic cores were described using the projector augmented wave (PAW) method [13] with generalized gradient approximations, known as PBE [14]. The advantage of this code is that the ab initio energy of the system can be combined with the molecular dynamics method to simulate the properties of cubic CaSiO3 perovskite at high pressure and high temperature simultaneously. The PAW potentials of Ca, Si, and O had core radii of 2.3, 1.5, and 1.1 a.u., respectively. Single particle orbitals were expanded in plane waves, with a plane-wave cut-off of 900 eV. The calculations were performed on the basis of a self-consistency convergence on the total energy of 10−4 eV per simulation cell. We used a 135-atom supercell, with Γ-point Brillouin zone sampling and a time step of 1 fs at a constant volume. The simulations were run in the constant NVT ensemble with a Nosé thermostat [15] for at least 5 ps after equilibration. The computation time required to reach equilibrium varied between configurations, and depended on the starting atomic position, velocity, temperature, and pressure. In previous studies, we have confirmed that useful data for the elastic properties of solids in high pressure and temperature conditions can be acquired using the previous AIMD calculations, e.g., [16,17]. In this study, AIMD calculations were performed under 27 selected pressure and temperature conditions up to 175 GPa and 4000 K. A comprehensive description of our method as applied to the modeling of condensed matter has been described previously [18].
The elastic constants can be determined from the computation of the second derivatives of the free energy as a function of small strains [19]. For a cubic crystal, the three elastic moduli, C11, C12, and C44, fully describe its elastic behavior. The values of C11 and C12 can be determined from the bulk modulus K and shear constant CS
K = ( C 11 + 2 C 12 ) / 3
C S = ( C 11 C 12 ) / 2
The following tetragonal strains were applied to obtain CS:
ε = ( e 0 0 0 e 0 0 0 ( 1 + e ) 2 1 )
where e is the strain magnitude. The change in the free energy of the strained structure, ΔE(e), is related to e as follows:
Δ E ( e ) = 3 V ( C 11 C 12 ) e 2 + O ( e 3 )
where V is the volume of the cell. C44 was calculated by applying the volume-conserving orthorhombic strain:
ε = ( 0 e 0 e 0 0 0 0 e 2 / ( 1 e 2 ) )
The energy associated with this strain is:
Δ E ( e ) = 2 C 44 V e 2 + O ( e 4 )
According to the calculations for unstrained and strained structures, the elasticity of cubic CaSiO3 perovskite at high pressure and high temperature can be determined.

3. Results

The EOS of cubic CaSiO3 perovskite has been investigated in a previous experimental study [20]. Recent theoretical studies have investigated the physical properties of materials under high pressure and high temperature using first-principles calculations. We noticed that the scatter of the EOS determined by experimental study was smaller than that obtained from first-principles calculations [18], indicating that the EOS determined in experiments was more accurate than that determined from first-principles calculations. By contrast, the AIMD calculations present significant advantages for investigating the elastic properties of materials under high temperatures and high temperatures. Therefore, we used the experimental data to determine the EOS for cubic CaSiO3 perovskite. The pressures estimated by the AIMD calculations were corrected based on the EOS determined by the experimental data. The combination of first-principles molecular dynamics calculations and high pressure experimental data led us to determine reliable physical properties over a wide range of pressures and temperatures. The EOS for a solid can be described in a general form as a functional relationship between pressure, volume, and temperature:
P T o t a l ( V , T ) = P s t ( V , 300 ) + P t h ( V , T )
A fit of the volume-pressure data yielded volume and bulk modulus values of V0 = 45.58 Å3, KT0 = 236 GPa and K’T0 = 3.9 [20] for a third-order Birch-Murnaghan EOS [21]:
P s t = 3 2 K T 0 [ ( V 0 V ) 7 3 ( V 0 V ) 5 3 ] { 1 3 4 ( 4 K ' T 0 ) [ ( V 0 V ) 2 3 1 ] }
where V0 and KT0, and K’T0 are the volume, isothermal bulk modulus, and first pressure derivative of the isothermal bulk modulus, respectively. The thermal pressure, Pth, of the thermal pressure EOS can be written as follows:
P t h = [ α K T ( V 0 , T ) + ( K T T ) V ln ( V 0 V ) ] ( T T 0 )
A least squares fit of the high temperature data from the AIMD calculations yields α K T ( V 0 , T ) = 0.0083 and (δKT/δT)V = −0.0031. The value of Pth of cubic CaSiO3 perovskite was not sensitive to changes in volume at the pressures investigated in this study, because the values of (δKT/δT)V were very small. The fitting parameters of the third-order Birch-Murnaghan EOS combined with the thermal pressure EOS were V0, KT0, KT0, αKT(V0,T), and (δKT/δT)V. The results of the fit of our P-V-T data to the thermal pressure equation of state are summarized in Table 1.
Table 1. The thermoelastic parameters of cubic CaSiO3 perovskite. The third-order Birch–Murnaghan EOS was used to calculate the parameters of cubic CaSiO3 perovskite. Key: KT0, isothermal bulk modulus at 0 GPa and 300 K; K’T0, first pressure derivation of the bulk modulus; V0, volume at 0 GPa and 300 K. The terms αKT(V0,T) and (KT/T)V are parameters of the thermal pressure.
Table 1. The thermoelastic parameters of cubic CaSiO3 perovskite. The third-order Birch–Murnaghan EOS was used to calculate the parameters of cubic CaSiO3 perovskite. Key: KT0, isothermal bulk modulus at 0 GPa and 300 K; K’T0, first pressure derivation of the bulk modulus; V0, volume at 0 GPa and 300 K. The terms αKT(V0,T) and (KT/T)V are parameters of the thermal pressure.
ParameterValue
V03)45.58 a
KT0 (GPa)236 a
K’T03.9 a
αKT(V0,T) (GPa/K)0.0083(3)
(KT/T)V (GPa/K)−0.0031(31)
a The parameters are from Shim et al. [20].
We determined the elastic constant by computing the stress generated by small deformations of the equilibrium cell. Figure 1 shows three elastic constants of cubic CaSiO3 perovskite (C11, C12 and C44) at 2000 K as a function of pressure up to 160 GPa. The bulk modulus of an isotropic aggregate cubic crystal is well defined, whereas the shear modulus can be constrained by the Voigt-Reuss-Hill scheme [22]:
G V = 1 5 ( 2 C S + 3 C 44 )
G R = [ 1 5 ( 2 C S + 3 C 44 ) ) ] 1
G H = 1 2 ( G V + G R )
We also show the bulk modulus K and Hill’s average G at 2000 K as a function of pressure in Figure 1. Karki and Crain [9] calculated elastic constants and moduli of cubic CaSiO3 perovskite at 0 K up to 140 GPa. Our results for elastic parameters calculated at 2000 K were in general agreement with those at 0 K reported by previous studies.
Figure 1. Pressure dependence of three elastic constants, C11, C12, and C44, and the isotropic bulk (K) and shear (G) moduli of cubic CaSiO3 perovskite at 2000 K. The solid circles and diamonds represent the elastic constants and the elastic moduli, respectively. The solid and dashed lines are the fits of each parameter.
Figure 1. Pressure dependence of three elastic constants, C11, C12, and C44, and the isotropic bulk (K) and shear (G) moduli of cubic CaSiO3 perovskite at 2000 K. The solid circles and diamonds represent the elastic constants and the elastic moduli, respectively. The solid and dashed lines are the fits of each parameter.
Entropy 15 04300 g001
The three isotropically averaged aggregate sound velocities could be derived from the bulk modulus K and shear modulus G:
V P = [ ( K + 4 3 G ) / ρ ] 1 2
V B = ( K ρ ) 1 2
V S = ( G ρ ) 1 2
where VP, VB, and VS are the compressional, bulk, and shear sound wave velocities, respectively, and ρ is the density. The three sound wave velocities, VP, VB, and VS, increased with increasing pressure at 2000 K in Figure 2. Our results for sound wave velocities were in good agreement with those reported by Li et al. [8].
Figure 2. Sound wave velocities in cubic CaSiO3 perovskite at 2000 K calculated from the elastic constants. The solid circles, squares, and diamonds represent the compressional, bulk, and shear velocities, respectively.
Figure 2. Sound wave velocities in cubic CaSiO3 perovskite at 2000 K calculated from the elastic constants. The solid circles, squares, and diamonds represent the compressional, bulk, and shear velocities, respectively.
Entropy 15 04300 g002
The effect of temperature on the sound wave velocities was investigated at high temperatures corresponding to conditions in the Earth’s mantle. In Figure 3, the results of the AIMD simulations at high temperatures of 2000 ≤ T(K) ≤ 4000 showed that sound wave velocities decreased with increasing temperature. However, there were only small dependencies on temperature. As the pressure increased, these dependencies on temperature became small. The sound wave velocities were fitted to the following equation as functions of temperature and pressure:
v = a + ( b + c P ) T + d ln ( P )
where a, b, c and d are fitted parameters, and T and P are given in K and GPa, respectively. The results of the fitted parameters are summarized in Table 2.
Figure 3. Sound wave velocities in cubic CaSiO3 perovskite at high temperatures. The diamonds represent the compressional and shear wave velocities calculated by the AIMD simulations. The dashed lines represent the fitted velocities for 2000, 3000, and 4000 K.
Figure 3. Sound wave velocities in cubic CaSiO3 perovskite at high temperatures. The diamonds represent the compressional and shear wave velocities calculated by the AIMD simulations. The dashed lines represent the fitted velocities for 2000, 3000, and 4000 K.
Entropy 15 04300 g003
Table 2. Parameters of the compressional and shear sound velocities. The parameters are given by v = a + ( b + c P ) T + d ln ( P ) , where T and P are the temperature (K) and the pressure (GPa), respectively. The conditions for applying these parameters to Equation (16) are 15 GPa < P < 140 GPa and 1500 K < T < 4500 K.
Table 2. Parameters of the compressional and shear sound velocities. The parameters are given by v = a + ( b + c P ) T + d ln ( P ) , where T and P are the temperature (K) and the pressure (GPa), respectively. The conditions for applying these parameters to Equation (16) are 15 GPa < P < 140 GPa and 1500 K < T < 4500 K.
abcd
VP7.12(0.27)−6.96(0.33) × 10−44.07(0.30) × 10−71.56(0.06)
VS4.65(0.15)−3.20(0.02) × 10−41.75(0.17) × 10−60.74(0.04)

4. Discussion

It is important to assess the uncertainties of the ab initio calculations to understand the implications of the calculated results. In general, different types of approximation have led to different values in previous ab initio studies, e.g., [18]. The difference between Local Density Approximation (LDA) and Generalized Gradient Approximation (GGA) leads to a change in cell volume of a few percent. This uncertainty is non-negligible in the context of discussing the behavior of the Earth’s mantle. Although GGA was used in the present study, we corrected the calculated values according to the experimental EOS to minimize the uncertainties related to approximations used in ab initio simulations. Therefore, our discussion of the comparison between estimated elastic properties and PREM values is more reliable than those of previous studies.
It is believed that the lower mantle contains three minerals; (Fe,Al)-bearing Mg-perovskite, ferropericlase, and Ca-perovskite [2,3]. We calculated the density, and the compressional and shear sound wave velocities for cubic CaSiO3 perovskite in order to compare them with the values from the PREM [11]. The values for cubic CaSiO3 perovskite were calculated using the EOS defined in Table 1 and Equation (16) defined in Table 2. The estimated values for cubic CaSiO3 perovskite and the PREM are compared in Figure 4. The adiabatic temperature profile (geotherm) was used as the temperature profile in the Earth’s lower mantle [3]. The calculated density of cubic CaSiO3 perovskite was in good agreement with that estimated by seismological data (PREM). Li et al. [8] estimated the density of cubic CaSiO3 perovskite, and the estimated density was higher than that in the PREM data. Although our AIMD method was similar to that used by Li et al. [8], the discrepancy between this and the previous study was confirmed. As the pressure was corrected to estimate density accurately in this study, the difference between our estimated density of cubic CaSiO3 perovskite and that from the PREM data should be small. For the compressional sound wave velocity, the discrepancy between the calculated values for cubic CaSiO3 perovskite and the observed data was small. The difference increased as the depth increased. By contrast, the shear sound wave velocity in cubic CaSiO3 perovskite was much higher than that from the PREM data. If cubic CaSiO3 perovskite is a major mineral in the Earth’s lower mantle, the sound wave velocity profiles cannot be explained. Therefore, our study implies that cubic CaSiO3 perovskite is a minor mineral in the Earth’s lower mantle. In a previous study, the shear sound wave velocity in orthorhombic (Mg,Fe)SiO3 perovskite was reported to be lower than that in cubic CaSiO3 perovskite at 0 K [23]. As the shear sound wave velocity in ferropericlase, (Mg,Fe)O, is much lower than that of orthorhombic (Mg,Fe)SiO3 perovskite, sound wave velocities in orthorhombic (Mg,Fe)SiO3 perovskite at higher temperatures are therefore in general agreement with those from PREM, indicating that orthorhombic (Mg,Fe)SiO3 perovskite might be a major mineral in the Earth’s lower mantle.
Figure 4. Density and sound wave velocities for cubic CaSiO3 perovskite compared with PREM data. The solid circles represent the values from PREM [11]. The solid lines represent the calculated values under the conditions of the Earth’s lower mantle. The values for cubic CaSiO3 perovskite were calculated using the equations defined in Table 1 and Table 2 and the adiabatic temperature profile (geotherm) in the Earth’s interior [3]; a: compressional sound velocity; b: shear sound velocity; c: density.
Figure 4. Density and sound wave velocities for cubic CaSiO3 perovskite compared with PREM data. The solid circles represent the values from PREM [11]. The solid lines represent the calculated values under the conditions of the Earth’s lower mantle. The values for cubic CaSiO3 perovskite were calculated using the equations defined in Table 1 and Table 2 and the adiabatic temperature profile (geotherm) in the Earth’s interior [3]; a: compressional sound velocity; b: shear sound velocity; c: density.
Entropy 15 04300 g004
In general, a certain quantity of most elements dissolves in three minerals, namely (Fe,Al)-bearing Mg-perovskite, ferropericlase, and Ca-perovskite, and the partition coefficients of minor elements between the three minerals change with temperature and pressure. In this study, the physical properties of a pure cubic CaSiO3 perovskite host were calculated, because the complicated chemical composition would need a very large simulation time, and it would have been difficult to perform a reliable AIMD study. The effects of the minor elements on the density and the sound wave velocities therefore require further investigation.

5. Conclusions

We have predicted from the first-principles theory the sound velocities of cubic CaSiO3 perovskite at high pressures and temperatures corresponding to the Earth’s lower mantle. Comparison of the elastic properties of cubic CaSiO3 perovskite with the lower mantle properties estimated from the seismic observations supports the prevailing hypothesis that the lower mantle consists primarily of (Mg,Fe)SiO3 perovskite.

Acknowledgments

This work made use of the super computer system of JAMSTEC and of the computer systems of the Earthquake Information Center of the Earthquake Research Institute. This work was supported by Grant-in-Aid for Scientific Research from JSPS and the Earthquake Research Institute cooperative research program, Japan.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Anderson, D.L. Composition of the Earth. Science 1989, 243, 367–370. [Google Scholar] [CrossRef] [PubMed]
  2. Irifune, T.; Ringwood, A.E. Phase Transitions in Primitive MORB and Pyrolite Compositions to 25 GPa and Some Geophysical Implications. In High-Pressure Research in Mineral Physics; Manghnanii, M.H., Syono, Y., Eds.; Terra Scientific Publishing Company, Tokyo/American Geophysical Union: Washington, DC, USA, 1987; pp. 231–242. [Google Scholar]
  3. Ono, S. Experimental constraints on the temperature profile in the lower mantle. Phys. Earth Planet. Inter. 2008, 170, 267–273. [Google Scholar] [CrossRef]
  4. Shim, S.H.; Jeanloz, R.; Duffy, T.S. Tetragonal structure of CaSiO3 perovskite above 20 GPa. Geophys. Res. Lett. 2002, 29. [Google Scholar] [CrossRef]
  5. Ono, S.; Ohishi, Y.; Mibe, K. Phase transition of Ca-perovskite and stability of Al-bearing Mg-perovskite in the lower mantle. Am. Mineral. 2004, 89, 1480–1485. [Google Scholar]
  6. Caracas, R.; Wentzcovitch, R.; Price, G.D.; Brodholt, J. CaSiO3 perovskite at lower mantle pressures. Geophys. Res. Lett. 2005, 32, L06306. [Google Scholar] [CrossRef]
  7. Adams, D.J.; Oganov, A.R. Ab inito molecular dynamics study CaSiO3 perovskite at P-T conditions of Earth’s lower mantle. Phys. Rev. B 2006, 73, 184106. [Google Scholar] [CrossRef]
  8. Li, L.; Weidner, D.J.; Brodholt, J.; Alfè, D.; Price, G.D.; Caracas, R.; Wentzcovitch, R. Phase stability of CaSiO3 perovskite at high pressure and temperature: Insights from ab initio molecular dynamics. Phys. Earth Planet. Inter. 2006, 155, 260–268. [Google Scholar] [CrossRef]
  9. Karki, B.B.; Crain, J. First-principles determination of elastic properties of CaSiO3 perovskite at lower mantle pressures. Geophys. Res. Lett. 1998, 25, 2741–2744. [Google Scholar] [CrossRef]
  10. Stixrude, L.; Lithgow-Bertelloni, C.; Kiefer, B.; Fumagalli, P. Phase stability and shear softening in CaSiO3 perovskite at high pressure. Phys. Rev. B 2007, 75, 024108. [Google Scholar] [CrossRef]
  11. Dziewonski, A.M.; Anderson, D.L. Preliminary reference earth model. Phys. Earth Planet. Inter. 1981, 25, 297–356. [Google Scholar] [CrossRef]
  12. Kresse, G.; Furthmuller, 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]
  13. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  14. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  15. Nosé, S. A molecular dynamics method for simulations in the canonical ensemble. Mol. Phys. 1984, 52, 255–268. [Google Scholar] [CrossRef]
  16. Ono, S.; Brodholt, J.P.; Alfè, D.; Alfredsson, M.; Price, G.D. Ab initio molecular dynamics simulations for thermal equation of state of B2-type NaCl. J. Appl. Phys. 2008, 86, 5801–5808. [Google Scholar] [CrossRef]
  17. Ono, S. First-principles molecular dynamics calculations of the equation of state for tantalum. Int. J. Mol. Sci. 2009, 10, 4342–4351. [Google Scholar] [CrossRef] [PubMed]
  18. Ono, S. Synergy Between First-Principles Computation and Experiment in Study of Earth Science. In Some Applications of Quantum Mechanics; Pahlavani, M.R., Ed.; InTech: Vienna, Austria, 2012; pp. 91–108. [Google Scholar]
  19. Mehl, M.J.; Osburn, J.E.; Papaconstantopoulos, D.A.; Klein, B.M. Structural properties of ordered high-melting-temperature intermetallic alloys from first-principles total-energy calculations. Phys. Rev. B 1990, 41, 10311–10323. [Google Scholar] [CrossRef]
  20. Shim, S.H.; Duffy, T.S. The stability and P-V-T equation of state of CaSiO3 perovskite in the Earth’s lower mantle. J. Geophys. Res. 2000, 105, 25955–25968. [Google Scholar] [CrossRef]
  21. Poirier, J.P. Introduction to the Physics of the Earth’s Interior; Cambridge University Press: Cambridge, UK, 2000. [Google Scholar]
  22. Anderson, O.L. Equations of State of Solids for Geophysics and Ceramic Science; Oxford University Press: New York, NY, USA, 1995. [Google Scholar]
  23. Karki, B.B.; Stixrude, L. Seismic velocities of major silicate and oxide phases of the lower mantle. J. Geophys. Res. 1999, 104, 13025–13033. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Ono, S. Elastic Properties of CaSiO3 Perovskite from ab initio Molecular Dynamics. Entropy 2013, 15, 4300-4309. https://doi.org/10.3390/e15104300

AMA Style

Ono S. Elastic Properties of CaSiO3 Perovskite from ab initio Molecular Dynamics. Entropy. 2013; 15(10):4300-4309. https://doi.org/10.3390/e15104300

Chicago/Turabian Style

Ono, Shigeaki. 2013. "Elastic Properties of CaSiO3 Perovskite from ab initio Molecular Dynamics" Entropy 15, no. 10: 4300-4309. https://doi.org/10.3390/e15104300

Article Metrics

Back to TopTop