Vibrational and thermodynamic properties of forsterite at mantle conditions
Abstract
[1] We present a first-principle study of the vibrational and thermodynamic properties of Mg2SiO4 forsterite up to 20 GPa. The calculated local density approximation (LDA) frequencies and their pressure dependence are in good agreement with the available Raman and infrared spectroscopy data. We also predict the pressure dependence of the modes which are yet to be measured. Thermodynamic properties are obtained using the quasi-harmonic approximation (QHA) to the free energy in conjunction with these results. The calculated pressure-volume-temperature (P-V-T) relations and thermodynamic properties agree well with the reported experimental data within the regime of validity of the QHA. The only discrepancies with experimental data that point to intrinsic anharmonic effects are the mode Gruneisen parameters of two Raman and one infrared modes. However, their effect on thermodynamic properties appears to be negligible.
1. Introduction
[2] Mg2SiO4 forsterite is the end-member of olivine, which is the major constituent of the upper mantle. It has been widely studied; the results can be seen in both vibrational spectroscopy [Akaogi et al., 1984; Chopelas, 1990; Ghose et al., 1992; Gillet et al., 1991; Hofmeister, 1987; Iishi, 1978; Pilati et al., 1995; Rao et al., 1988; Wang et al., 1993] and elasticity measurements [Cynn et al., 1998; Downs et al., 1996; Duffy et al., 1995; Guyot et al., 1996; Isaak et al., 1989; Li et al., 1996; Meng et al., 1993; Zha et al., 1994]. Vibrational modes of forsterite measured at high pressure in the laboratory are limited in its number of detected modes (84 modes for forsterite in total) even though the stability of forsterite reaches 410 km depth of the Earth. Previous lattice dynamic calculations [Pavese, 1998; Price et al., 1987] on forsterite are limited to the usage of empirical potentials. The first-principles approach based on density functional theory [Hohenberg and Kohn, 1964] has been successfully applied to predict the structural and elastic properties of Mg2SiO4 forsterite at high pressures [da Silva et al., 1997; Wentzcovitch and Stixrude, 1997], while the vibrational properties are yet to be explored with this approach.
[3] In the past decade, first-principles lattice dynamics in conjunction with the quasi-harmonic approximation (QHA) has been applied to predict vibrational and thermodynamic properties of mantle minerals such as MgO pericalse [Karki et al., 2000c], MgSiO3 perovskite [Karki et al., 2000b], MgSiO3 ilmenite [Karki et al., 2000a], MgSiO3 postperovskite [Tsuchiy et al., 2004; Tsuchiya et al., 2005], and Mg2SiO4 ringwoodite [Yu and Wentzcovitch, 2006] at mantle conditions. The predicted vibrational and thermoelastic properties of these minerals and their aggregates have been reported to be in very good agreement with available experimental data and preliminary reference Earth model (PREM). Now we address the properties of Mg2SiO4 forsterite with the same methods. In contrast, with these other minerals, Mg2SiO4 forsterite is stable at lower pressures up to 14 GPa. Some studies have reported anharmonic characteristics in forsterite [Anderson, 1996; Cynn et al., 1996; Gillet et al., 1991] based on the observation of the temperature dependence of the Raman and acoustic frequencies, in addition to thermodynamic considerations by which CV was determined from measurements of CP. Here we will compare the calculated and measured vibrational properties of forsterite and investigate the validity of the QHA formulation to reproduce its thermodynamic properties.
2. Method
[4] Our computations use the local density approximation (LDA) [Celperley and Alder, 1990; Perdew and Zunger, 1981]. Calculation details are similar to those reported in previous works [ Karki et al., 2000b; Tsuchiy et al., 2004; Tsuchiya et al., 2005; Yu and Wentzcovitch, 2006]. The equilibrated structure of forsterite was calculated using the first-principles implementation of a variable cell-shape damped molecular dynamics (VCSMD) [Wentzcovitch and Price, 1996; Wentzcovitch et al., 1995]. Pseudopotentials of Mg were generated by the method of von Barth and Car [Karki et al., 2000b; Tsuchiya et al., 2005], while those of O and Si are by the method of Troullier and Martins [Troullier and Martins, 1991]. The plane wave energy cut-off used is 70 Ry, and the k point sampling of the charge density was performed on a 2 × 2 × 2 grid in Brillouin zone (BZ) shifted by The calculations with 2 × 2 × 2 and 4 × 2 × 4 grid of k points give a difference in energy by 0.1 meV/atom, in pressure by 0.03 GPa. Thus using 4 × 2 × 4 grid changes little on the precision for this study. The dynamical matrix was obtained using density functional perturbation theory (DFPT) [Baroni et al., 2001]. At each pressure, dynamical matrices were calculated on a 2 × 2 × 2 (without shift) q point mesh; force constants were extracted and used to produce matrices in a 12 × 12 × 12 q point grid. The corresponding normal modes were used in the calculation of the free energy.
3. Results and Discussion
[6] The calculated phonon dispersion along several symmetry directions and the vibrational density of states at 0 and 20 GPa are shown in Figure 1. The diagonal components of the dielectric tensor are (2.85, 2.77, 2.79) at zero pressure and (2.87, 2.78, 2.79) at 20 GPa. The Born effective charges of Mg are close to 2 for Mg, while those for Si and O are significantly different from ideal values of 4 and −2 and are highly anisotropic. For Mg, they are Z*[Mg] = (2.17, 1.72, 1.94) at 0 Gpa and (2.11, 1.71, 1.89) at 20 GPa; Z*[Si] = (3.07, 2.62, 2.92) at 0 GPa and (3.10, 2.74, 2.99) at 20 GPa; Z*[O1] = (−2.33, −1.40, −1.47) and Z*[O2] = (−1.60, −1.98, −1.53) at 0 GPa, and Z*[O1] = (−2.27, −1.42, −1.49) and Z*[O2] = (−1.58, −1.61, −1.85) at 20 GPa.
[7] We list our calculated frequencies of the optical Raman modes at 0 GPa and 0 K in Table 1, infrared modes in Table 2, and inactive modes at zone center in Table 3 and compare these with previously reported results at 0 GPa pressure. Also, listed are the mode Grüneisen parameters (γi= −d(lnωi)/dlnV). A number of studies have reported the phonon frequencies of forsterite at zero pressure using different techniques such as Raman and infrared spectroscopy [e.g., Iishi, 1978; Lam et al., 1990] and inelastic neutron scattering [Rao et al., 1988]; only the results from one study [Iishi, 1978] are listed in Tables 1 and 2, since these studies agree well with each other. The frequency of an infrared active mode splits into two values depending on whether the mode is longitudinal (LO) or transverse (TO), which is due to the contribution of the macroscopic electric field to the LO mode in a polar crystal. The calculated eigenvectors, which contain information about the internal and external motions of the SiO4 tetrahedral and Mg ions, are used to deduce the symmetry labels of the modes following a method described elsewhere [Rao et al., 1988]. We find similar frequency-pressure relationships between Raman and infrared modes. The calculated frequencies at the zone center are all positive and increase with pressure for Raman (Figure 2a) and infrared modes alike (Figure 2b). The pressure dependences of mode frequencies are consistent with available experimentally fitted curves [Chopelas, 1990; Hofmeister, 1987; Wang et al., 1993]. Only two Raman modes, the lowest B3g ones and one B1u IR mode, have the calculated γi considerably different from measured ones. This might be signs of forsterite’s anharmonicity, not properly accounted for by the present approach. However, their effect on thermodynamics properties seems to be unnoticeable within the regime of validity of the QHA.
Calculation | Experiment | Modes Typea | ||||||
---|---|---|---|---|---|---|---|---|
υI | γi | υi | γi | |||||
a,b | c | c | d | e | ||||
Ag | 188 | 0.30 | 183 | T′(Mg2, SiO4:x) | ||||
222 | 0.75 | 227 | 227 | 0.70 | 0.64 | 0.67 | T′(SiO4, Mg2:y) | |
316 | 1.21 | 305 | 306 | 1.80 | 1.36 | 1.63 | T′(Mg2, SiO4:x) | |
333 | 1.21 | 329 | T′(Mg2, SiO4:x) | |||||
357 | 1.19 | 340 | 341 | 1.78 | 1.87 | R′(SiO4:z) | ||
436 | 1.36 | 424 | υ2 | |||||
529 | 0.77 | 546 | υ4 | |||||
596 | 0.69 | 609 | 609 | 0.70 | 0.68 | 0.70 | υ4 | |
818 | 0.46 | 826 | 826 | 0.50 | 0.48 | 0.48 | υ1 | |
850 | 0.42 | 856 | 856 | 0.50 | 0.43 | 0.49 | υ3 | |
965 | 0.59 | 965 | υ3 | |||||
B3g | 195 | 1.11 | 226 | 183 | 2.15 | 2.09 | T′(SiO4:z) | |
284 | 0.73 | 272 | 290 | 1.35 | T′(Mg2:z) | |||
320 | 1.07 | 318 | R′(SiO4:y) | |||||
383 | 0.94 | 376 | R′(SiO4:z) | |||||
418 | 1.14 | 412 | 424 | 0.97 | 0.99 | υ2 | ||
577 | 0.58 | 595 | 587 | 0.61 | υ4 | |||
914 | 0.36 | 922 | 922 | 0.40 | 0.38 | 0.38 | υ3 | |
B2g | 174 | 0.09 | 142 | T′(Mg2,z) | ||||
249 | 1.07 | 244 | 244 | 1.23 | 1.21 | T′(SiO4:z) | ||
329 | 1.24 | 324 | R′(SiO4:x) | |||||
370 | 1.04 | 368 | 376 | 1.40 | 1.26 | 1.25 | R′(SiO4:y) | |
450 | 1.27 | 441 | 441 | 1.80 | 1.59 | 1.60 | υ2 | |
568 | 0.57 | 588 | 545 | 0.60 | 0.42 | 0.53 | υ4 | |
877 | 0.40 | 884 | 884 | 0.50 | 0.44 | 0.44 | υ3 | |
B1g | 222 | 0.81 | 192 | 227 | 0.70 | 0.64 | 0.67 | T′(Mg2:SiO4:y) |
256 | 0.94 | 224 | T′(Mg2:SiO4:x) | |||||
327 | 1.20 | 260 | T′(Mg2:SiO4:x) | |||||
360 | 1.16 | 318 | 331 | 1.30 | 1.14 | 1.12 | T′(Mg2:y) | |
384 | 0.90 | 418 | R′(SiO4:z) | |||||
444 | 1.23 | 434 | 434 | 1.35 | 1.40 | υ2 | ||
569 | 0.61 | 583 | 584 | 0.60 | 0.48 | 0.66 | υ4 | |
618 | 0.68 | 632 | υ4 | |||||
829 | 0.48 | 839 | 826 | 0.50 | 0.48 | 0.48 | υ1 | |
858 | 0.40 | 866 | 856 | 0.50 | 0.43 | 0.49 | υ3 | |
975 | 0.59 | 976 | 966 | 0.70 | 0.66 | 0.66 | υ3 |
- a The assignments follow a previous study [Iishi, 1978]. R (rotational lattice mode) and T (translational lattice modes) are the two external modes. The four internal vibrational modes of SiO4 ion are υ1(A1 type), υ2(E type), υ3, and υ4 (F2 type). References: a, [Iishi, 1978]; b, [Lam et al., 1990]; c, [Gillet et al., 1991 ]; d, [Wang et al., 1993]; and e, [Chopelas, 1990]. vi is the vibrational frequency; γi is the Grüneisen parameter of mode i as defined in equation 5b.
Calculation | Experiment | Mode Typea | ||||||||
---|---|---|---|---|---|---|---|---|---|---|
υTO | υLO | γi | υTO | υLO | υi | γi | υi | γi | ||
a | a | b | b | c | c | |||||
B3u | 205 | 205 | 1.16 | 201 | T′(Mg1, SiO4:y) | |||||
277 | 277 | 0.57 | 224 | T′(Mg2:x,y) | ||||||
296 | 297 | 1.04 | 274 | 276 | T′(Mg1:x; Mg2:y) | |||||
321 | 323 | 1.12 | 293 | 298 | T′(Mg1:x,z) | |||||
398 | 403 | 1.13 | 320 | 323 | T′(Mg1:z,y) | |||||
407 | 450 | 1.07 | 378 | 386 | T′(Mg2:x,y) | |||||
482 | 482 | 1.00 | 403 | 469 | R′(SiO4:x) | |||||
508 | 520 | 0.89 | 498 | 544 | υ2 | |||||
531 | 532 | 0.86 | 562 | 566 | υ4 | |||||
593 | 606 | 0.61 | 601 | 645 | υ4 | |||||
824 | 825 | 0.48 | 838 | 845 | 846 | 0.39 | 836 | 0.47 | υ1 | |
954 | 965 | 0.41 | 957 | 963 | 962 | 0.32 | 925 | 0.23 | υ3 | |
975 | 1024 | 0.61 | 980 | 1086 | υ3 | |||||
B2u | 146 | 146 | 1.06 | 144 | T1(Mg1:SiO4:x) | |||||
282 | 284 | 0.80 | 224 | T;(Mg2:x) | ||||||
295 | 307 | 0.74 | 280 | 283 | T′(Mg1:y) | |||||
363 | 376 | 1.27 | 294 | 313 | T′(Mg1:x,z; Mg2:y) | |||||
398 | 419 | 1.11 | 352 | 376 | T′(Mg2:x,y) | |||||
427 | 430 | 1.20 | 400 | 412 | T′(Mg1:z) | |||||
463 | 472 | 1.14 | 421 | 446 | R′(SiO4:z) | |||||
502 | 507 | 0.74 | 465 | 493 | υ2 | |||||
530 | 532 | 0.86 | 510 | 516 | υ4 | |||||
617 | 617 | 0.66 | 537 | 597 | 614 | 0.54 | 609 | 0.28 | υ4 | |
825 | 830 | 0.48 | 838 | 843 | υ1 | |||||
868 | 914 | 0.43 | 882 | 979 | υ3 | |||||
985 | 986 | 0.59 | 987 | 993 | 992 | 0.65 | 988 | 0.68 | υ3 | |
B1u | 194 | 194 | 0.63 | 201 | T′(Mg1:y,z) | |||||
278 | 279 | 0.57 | 224 | T′(Mg1:x) | ||||||
296 | 304 | 0.81 | 274 | 278 | T′(Mg1:y; Mg2:z) | |||||
316 | 317 | 1.11 | 296 | 318 | T′(Mg1:z) | |||||
426 | 438 | 1.24 | 365 | 371 | R′(SiO4:z) | |||||
428 | 443 | 1.38 | 423 | 459 | R′(SiO4:y) | |||||
475 | 490 | 0.96 | 483 | 489 | υ2 | |||||
504 | 518 | 0.87 | 502 | 585 | 517 | 0.50 | 483 | 0.38 | υ4 | |
870 | 931 | 0.40 | 885 | 994 | 887 | 0.39 | 876 | 0.32 | υ3 |
- a The assignments follow a previous study [Iishi, 1978]. R (rotational lattice mode) and T (translational lattice modes) are the two external modes. The four internal vibrational modes of SiO4 ion are υ1(A1 type), υ2(E type), υ3 and υ4 (F2 type). References: a, [Iishi, 1978]; b, [Wang et al., 1993]; c, [Hofmeister, 1987]. vi is the vibrational frequency; γi is the Grüneisen parameter of mode i as defined in equation 5b.
υi | υia | γi |
---|---|---|
104 | 171 | 0.98 |
180 | 161 | 0.16 |
246 | 225 | 0.69 |
299 | 286 | 1.17 |
348 | 347 | 1.08 |
390 | 376 | 1.17 |
439 | 437 | 1.31 |
480 | 469 | 0.95 |
509 | 511 | 0.86 |
905 | 885 | 0.37 |
- a [Lam et al., 1990].
4. Thermodynamic Properties
[9] The calculated pressure-volume-temperature (P-V-T) relations are plotted in Figure 4. These curves are fitted with isothermal third-order finite strain equations [Birch, 1986] yielding V0 = 289.5 Å3, K0 = 126.4 GPa, Ê0 = 4.2. These compression curves are in good agreement with the experimental data [Downs et al., 1996; Gillet et al., 1991, Meng et al., 1993, 106; Guyot et al., 1996]. A comparison of calculated and measured thermodynamic properties at room temperature and pressure is reported in Table 4. The agreement between calculated and measured values is excellent.
300 K, 0 GPa Calculation, This Study | 300 K, 0 GPa Experiment | References | |
---|---|---|---|
V(Å3) | 289.5 | 291.9 | [Guyot et al., 1996] |
290.1 | [Downs et al., 1996] | ||
289.2 | [Gillet et al., 1991] | ||
KT(GPa) | 126.4 | 125(2) | [Downs et al., 1996] |
127.7(2) | [Gillet et al., 1991] | ||
KT′ | 4.2 | 4.0(4) | [Downs et al., 1996] |
α(× 10−5 K−1) | 2.64 | 2.83 | [Bouhifd et al.,1996] |
2.77 | [Gillet et al., 1991] | ||
2.48 | [Suzuki et al., 1984] | ||
2.56 | [Hazen, 1976] | ||
2.72 | [Kajiyoshi, 1986] | ||
CP(J mol−1 K−1) | 119.3 | 117.9 | [Gillet et al., 1991] |
S(J mol−1 K−1) | 95.9 | 93.14 | [Chopelas, 1990] |
γth | 1.23 | 1.28 | [Gillet et al., 1991] |
1.29 | [Isaak et al., 1989] | ||
1.29 | [Chopelas, 1990] | ||
KS(GPa) | 127.6 | 128.8(5) | [Zha et al., 1994] |
∂KT/∂T(× 10−2 GPa K−1) | −2.1 | −2.1(2) | [Meng et al., 1993] |
−2.0(2) | [Gillet et al., 1991] |
[10] In Figures 5a–5g, the calculated thermodynamical properties are plotted and compared with reported data as a function of pressure and temperature. The isothermal (KT) and adiabatic (KS) bulk modulus are plotted in Figures 5a and b. Our KTs are in excellent agreement with experimental values at 0 GPa [ Gillet et al., 1991] and so is KS = KT(1 + αγT) within the regime of validity of the QHA [ Gillet et al., 1991 ; Isaak et al., 1989]. Figure 5c shows the thermal expansivity, , which is determined from the volume dependence of temperature at each pressure. The scattering of experimental data [Bouhifd et al., 1996; Chopelas, 1990; Fei and Saxena, 1987; Gillet et al., 1991; Hazen, 1976; Kajiyoshi, 1986; Suzuki et al., 1984] is substantial at high temperatures even at 0 GPa. Nevertheless, our results are in very good agreement with experimental data within the range of validity of the QHA. At higher pressure, the effect of anharmonicity caused by temperature is expected to decrease. The estimated thermal expansivity at 10 GPa by Guyot et al. [1996] is in excellent agreement with our predictions even beyond the QHA permitted region. Below 1000 K, the differences at 10 GPa are probably due to usage of a less than suitable functional form for α(T) to fit experimental data.
[12] A solid is said to be anharmonic in behavior when CV is larger than that predicted by the Dulong and Petit limit (3nR) [Cynn et al., 1996]. Forsterite demonstrates this behavior [Anderson, 1996; Cynn et al., 1996; Gillet et al., 1991] at zero pressure and high temperatures. The deviation of our calculated CV from the experimental results increases with temperature, particularly in the P-T regime where the QHA is no longer expected to be valid. This is the sign of anharmonic effects and is shown in Figure 5e. The difference between calculated and measured CP, CP = CV(1 + αγT), is relatively small (Figure 5f). The temperature and pressure dependences of entropy are plotted in Figures 5g–5h. The negative pressure dependences of the entropy at all temperatures are consistent with the experimental results [Chopelas, 1990; Guyot et al., 1996]. Our calculated entropy agrees very nicely with the experimental estimates considering the limited number of phonon modes used in the experimental data.
5. Summary
[13] We have reported first-principles phonon dispersion and vibrational density of states for Mg2SiO4 forsterite up to 20 GPa using density functional theory. Our calculated Raman and infrared frequencies and their pressure dependences are in excellent agreement with available experimental data. We also predict the pressure dependence of some modes which are yet to be measured. The thermoelastic properties were derived from calculated vibrational density of states (VDoS) in conjunction with the QHA. The calculated compression curves, the isothermal bulk modulus, and the constant pressure-specific heat and entropy agree extremely well with the reported experimental data, regardless of claimed anharmonic effects in this solid. The computed mode Grüneisen parameters, γi at zero pressure for three modes, two B3g and one B1u, differ considerably from measured values, pointing perhaps to the origin of anharmonic effects in this solid. However, the QHA still seems to be very effective in describing thermodynamic, including CV, properties within its regime of validity. Deviations beyond this limit support the conclusion that forsterite is a solid that is both anharmonic in CV and quasi-harmonic in the thermal pressure in high-temperature region [Anderson, 1996; Cynn et al., 1996 ; Gillet et al., 1991]. The properties of this material at relevant mantle conditions are well predicted by the QHA.
Acknowledgments
[14] This research is supported by NSF/EAR 01355, 0230319, NSF/ITR 0428774 (VLAB), and Minnesota Supercomputing Institute. Calculations are performed with Quantum-ESPRESSO package from the Web site http://www.pwscf.org. DJW and LL acknowledge NSF EAR-9909266, EAR0135551, and EAR00135550.