You are on page 1of 9

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115, B12205, doi:10.

1029/2010JB000843, 2010

Silicon self‐diffusion of MgSiO3 perovskite by molecular dynamics


and its implication for lower mantle rheology
Yosuke Ito1 and Mitsuhiro Toriumi1
Received 3 January 2010; revised 20 June 2010; accepted 2 July 2010; published 4 December 2010.
[1] The pressure effect of silicon self‐diffusion of MgSiO3 perovskite was investigated by
molecular dynamics (MD). The viscosity variation of MgSiO3 perovskite in the lower
mantle was derived by the Nabarro‐Herring (Herring‐Nabarro) model. For the MD
calculation, the spontaneous jumping of atoms by self‐diffusion was reproduced without
using artificial forces, and the consistency of migration enthalpy (Hm*) with
experimental data was improved. The results showed that migration enthalpy increases
monotonically with increasing pressure. The viscosity of MgSiO3 perovskite in the lower
mantle increases monotonically with increasing depth. The obtained depth profile is
distinguishable from that of MgO periclase and can be utilized to determine which
mineral dominates the lower mantle rheology. Depending upon the assumed shape of the
depth profile for lower mantle viscosity, we considered the dominant mineral as (1)
MgSiO3 perovskite for the monotonic shape case or (2) MgO periclase for the hill shape
case that has the highest‐viscosity zone in the middle of the lower mantle.
Citation: Ito, Y., and M. Toriumi (2010), Silicon self‐diffusion of MgSiO3 perovskite by molecular dynamics and its
implication for lower mantle rheology, J. Geophys. Res., 115, B12205, doi:10.1029/2010JB000843.

1. Introduction potential model used by Wright and Price [1993]. Karki


and Khanduja [2007] sought to examine the incon-
[2] Understanding the self‐diffusion of MgSiO3 perov-
sistencies by employing ab initio pseudopotential; how-
skite is important, largely because of the implications for
ever, the inconsistency still remained (657 kJ/mol at 30 GPa).
diffusion creep, which is considered to be the dominant
Ammann et al. [2008] attempted to reduce the inconsis-
mechanism of creep deformation in the lower mantle [Karato
tency by searching surrounding atoms’ configurations that
et al., 1995]. Diffusion creep is dominated by the diffusion
minimized the heights of energy barriers and reported an
of the slowest diffusion species. In MgSiO3 perovskite,
improved result (453 kJ/mol at 25 GPa), although there was
silicon is one of the most likely candidates for the slowest still a mismatch between experiments and simulations.
species since it has been confirmed as the slowest diffusion
[3] In order to resolve the inconsistency, we tried another
species in many silicates (reviewed by Béjina et al. [2003]).
approach. Atoms spontaneously jump into vacancies with
Reported experimental measurements for silicon diffusion
varying surrounding atoms’ configurations during natural
in MgSiO3 perovskite [Yamazaki et al., 2000; Dobson et al.,
self‐diffusion, though the previous simulations did not
2008] are limited to the uppermost pressure (25 GPa) area
reproduce the process. Instead, these simulations assumed
in the lower mantle. Because of the difficulty of performing
some atomic configurations of activated states (“saddle
experiments at high pressures, molecular simulations are
point configuration”) and tried to derive heights of energy
considered indispensable for investigating the rheology of
barriers by pushing atoms into vacancies based on the
the lower mantle. Wright and Price [1993] sought to
thermally activated process model. According to the model
determine the diffusive properties of MgSiO3 perovskite [e.g., Heitjans and Kaerger, 2004], the jump rate w with
using lattice dynamics. They derived activation enthalpy
which an atom jumps into an empty neighborhood site is
(Hm*) obtained from the height of energy barriers: the
described as w = v0 exp (−GM/RT), where GM is the height
increase in energy migrating atoms from their sites to saddle of energy barrier, R is the gas constant, and T is temperature.
points. However, there were inconsistencies that could not
The preexponential factor v0 denotes attempt frequency. The
be ignored between experiments (336 kJ/mol at 25 GPa
self‐diffusion coefficient D is described as D = (1/6)wl2,
[Yamazaki et al., 2000], 347 kJ/mol at 25 GPa [Dobson et al.,
where l is the jump distance. There are two difficulties in
2008]) and simulations (733 kJ/mol (0 GPa]) [Wright and
this approach. First, although definition of attempt fre-
Price, 1993]). These inconsistencies could be due to the
quency is given in the harmonic transition theory, the way to
apply the definition to more complex actual systems is
1
Department of Complexity Science and Engineering, Graduate School
unclear. Generally [e.g., Heitjans and Kaerger, 2004], the
of Frontier Science, University of Tokyo, Tokyo, Japan. attempt frequency is assumed to be in the order of the Debye
frequency. Ita and Cohen [1997] assumed that the attempt
Copyright 2010 by the American Geophysical Union. frequency is equal to the lowest frequency peak in the
0148‐0227/10/2010JB000843

B12205 1 of 9
B12205 ITO AND TORIUMI: SILICON SELF‐DIFFUSION OF PEROVSKITE B12205

spectrum of the trajectories of the migrating atom projected neutrality. As a result, the MD system contained 1997
onto the shortest path of the vacancy. Next, a specific saddle atoms. Periodic boundary conditions were applied for all
point configuration needs to be assumed. Ammann et al. (x, y, and z) directions to approximate the actual size of the
[2008] tried to search for a “better” saddle point configu- crystal. The potential that an atom feels was given by the
ration, though some overestimations of activation enthalpies infinite summation of potentials between one atom and all
still remained. On the other hand, our approach just re- the other atoms. The Coulomb term (the first term of equation
produces the spontaneous diffusion itself explicitly by (1)) was summed by the Ewald method [Ewald, 1921].
molecular dynamics (MD). The advantage of the approach Summations of short‐range terms (the second and third
is that what we see is what we get: it directly derives the terms) were cut off at 7.5 Å. The Verlet algorithm integrated
self‐diffusion coefficient without assuming attempt fre- Newton’s equations of motions with 2 fs intervals. Isothermal
quency and saddle point configuration. There is another and isobaric (NPT) ensembles were obtained by the scaling
complication in this approach. Since the running time of method [Woodcock, 1971]. The MXDTRICL program
MD simulations is short and the frequency of spontaneous (http://www.sccj.net/download/JCPE) was used for MD
jumping is low, the precision of results is usually limited. calculations.
However, Ito and Toriumi [2007] showed that sufficient [7] The method for obtaining self‐diffusion coefficients
precision could be obtained with long MD running time. was based on the work by Ito and Toriumi [2007]. The self‐
[4] In summary, we applied molecular dynamics, which diffusion coefficient (D) was derived from the Einstein
reproduces the spontaneous jumping of atoms explicitly, to equation, which has the following form:
the silicon self‐diffusion of MgSiO3 perovskite. We em-
ployed four pressure conditions (40, 60, 80, and 100 GPa) 1 dD E
D¼ jRn j2 ; ð2Þ
corresponding to the lower mantle and derived self‐diffu- 6 dt
sion coefficients. The depth profiles of the viscosities of
lower mantle minerals were estimated from self‐diffusion where Rn indicates the positional vector of the nth atom while
coefficients [Yamazaki and Karato, 2001; Ito and Toriumi, the angle brackets indicate the average of all atoms. Positional
2007]. We also estimated the depth profile of MgSiO3 vectors were given by the trajectories of atoms recorded
perovskite’s viscosity in the lower mantle from the derived during MD calculations. Initial relaxations were performed
self‐diffusion coefficients. for 2 ns (106 steps) to achieve thermal equilibrium. After the
initial relaxation, 2 ns (106 steps) MD calculations were
2. Methodology repeated 10 times and an average of 10 self‐diffusion coef-
ficients was adopted. During the MD calculation, the
2.1. Molecular Dynamics vacancy concentration was fixed since there was no external
[5] In the MD simulation, the Buckingham model was sink or source of vacancies. Note that the vacancy concen-
adopted to define interatomic potential. The form of the tration is artificial and the derived self‐diffusion coefficients
model is as follows: should be calibrated before applying to the Earth’s interior by
  equation (5). Because the vacancy concentration is fixed,
  Zi Zj Rij cij activation enthalpies of the self‐diffusion coefficients ob-
Uij Rij ¼ þ bij exp   6; ð1Þ
Rij ij Rij tained by MD simulations correspond to migration enthalpies
(Hm*). The Arrhenius relationship is expressed as
where Rij is the interatomic distance. The parameters (ZMg =  
+1.9104; ZSi = +2.9043; ZO = −1.6049; bMg‐O = 1041.435 eV; D ¼ D0 exp Hm* =RT ; ð3Þ
rMg‐O = 0.2866 Å, cMg‐O = 0, bSi‐O = 1137.028 eV,
rSi‐O = 0.2827 Å, cSi‐O = 0, bO‐O = 2023.8 eV, rO‐O =
0.2674 Å, cO‐O = 13.83 eVÅ6) are taken from Oganov et al. where D is the self‐diffusion coefficient derived by the MD
[2000]. The model’s reproduction of crystal structure, elastic simulation, D0 is the preexponential term, R is the gas con-
properties, and thermal expansion was examined by Oganov stant, and T is temperature.
et al. [2000]. We also examined compression, thermal
expansion, and orthorhombic distortions and confirmed that 2.2. Estimation of Viscosity
the model reproduced these properties. Details of the ex- [8] In a material that flows by creep dominated by inner-
aminations are given in the next section. Ab initio electronic crystal diffusion, viscosity and self‐diffusion coefficients are
state calculations (reviewed by Karki et al. [2001]), which related by the Nabarro‐Herring (Herring‐Nabarro) model
would provide more precise results, require more calcula- [e.g., Poirier, 2000]:
tions than the Buckingham model. Our study targeted  
spontaneous migrations of atoms in MD calculations. Ito and  ¼ Að RT =V Þ G2 =Deff ð4Þ
Toriumi [2007] reported that 107 steps (20 ns) were needed to
maintain the precision of self‐diffusion coefficient results. where h is viscosity, A is the geometrical constant, R is the
Therefore, we considered that ab initio electronic state cal- gas constant, T is temperature, V is molar volume, G is
culations were not appropriate for this study. grain size, and Deff is the effective self‐diffusion coeffi-
[6] The MD system was composed of 5 × 5 × 4 unit cells, cient, which is approximately equal to that of the slowest
which contained 4 Mg atoms, 4 Si atoms, and 12 O atoms diffusing species. To relate Deff to the self‐diffusion
per unit cell with Pbnm symmetry. To induce silicon’s coefficient D of the MD calculation, we first assumed the
diffusion, we introduced one silicon vacancy. Two oxygen slowest diffusing species was Si. Note that Mg would also
vacancies were also introduced to maintain the total charge be examined further because it was shown that Fe‐Mg

2 of 9
B12205 ITO AND TORIUMI: SILICON SELF‐DIFFUSION OF PEROVSKITE B12205

where Ceff and C are Si vacancy concentrations of MgSiO3


perovskite in the lower mantle and the MD simulation,
respectively. Variations in grain size certainly influence
viscosity. The Nabarro‐Herring (Herring‐Nabarro) form
(equation (4)) suggests a square linear relationship between
grain size (G) and viscosity (h). However, because of a lack
of information for grain size distribution in the lower mantle,
we assumed it to be uniform. Yamazaki et al. [1996] reported
a significant slow grain growth for MgO periclase and
MgSiO3 perovskite aggregates (G / t1/10.6 for MgSiO3
perovskite and G / t1/10.8 for MgO periclase, where G and t
are grain size and time, respectively) at the pressure and
temperature condition near that of the lower mantle (1700 K
and 25 GPa), implying that nonuniformities for grain sizes in
the lower mantle are small. Two unknown constants, Ceff and
A, remained. Because of the difficulty of estimating their
correct values, we introduced a proportional relationship.
From equations (4) and (5) and the assumption, we derived a
Figure 1. Comparison of isothermal (300 K) compressions
relationship between viscosity h and the self‐diffusion
of MgSiO3 perovskite. The solid squares indicate results of
coefficients of MD simulations D as follows:
this study. Experiments were performed by diamond anvil
cell (DAC) (Yagi et al. [1979], open squares; Mao et al.  / ðV =RT Þð1=DÞ: ð6Þ
[1991], plusses; and Fiquet et al. [2000], crosses) and shock
compression (natural enstatite [Gong et al., 2004] shown as Since equation (6) derives viscosity variations, results of
open triangles). analysis based on equation (6) were shown as depth varia-
tions (see Figure 10), not absolute values.

interdiffusion is as slow as Si diffusion at the uppermost 3. Results


temperature and pressure conditions of the lower mantle
[Holzapfel et al., 2005]. Next, we assumed that the Si [9] First, we performed MD calculations for (1) isother-
vacancy was formed by impurities (i.e., extrinsic), and the mal (300 K) compressions, (2) isobaric (25 GPa) thermal
vacancy concentration does not depend on temperature and expansions, and (3) orthorhombic distortions to examine the
pressure. Another possibility for the mechanism of vacancy potential model employed in this study. The results were
formation is a thermal one (i.e., intrinsic). Two (intrinsic compared with experimental results in Figures 1, 2, and 3,
and extrinsic) mechanisms can be characterized by acti- respectively. In Figure 1, we considered experiments on
vation enthalpies (H*). For the intrinsic case, the activation pure synthetic MgSiO3 perovskite [Yagi et al., 1979; Fiquet
enthalpy is given by the total sum of the migration (Hm*) et al., 2000] and (Mg,Fe)SiO3 perovskite of Fe/Mg = 0–0.2
and formation (Hf*) enthalpies as H* = Hf*+ Hm*,
whereas in the extrinsic case the activation enthalpy is
simply given by the migration enthalpy Hm* as H* = Hm*.
The reported formation enthalpy (Hf*) of SiO2 Pseudo‐
Schottky (600 kJ/mol at 25 GPa [Karki and Khanduja,
2006]) is much higher than the activation enthalpies ob-
tained in experiments [Yamazaki et al., 2000; Dobson et al.,
2008], so we considered that the intrinsic case is unlikely.
We also assumed that the enhancement of diffusion by grain
boundaries (i.e., grain boundary diffusion) could be ne-
glected. Grain boundary diffusion enhances the effective
diffusion, especially in cases where grain sizes are small. In
cases where grain boundary diffusion dominates in a mate-
rial, the other creep model (Coble creep) rather than the
Nabarro‐Herring (Herring‐Nabarro) model dominates the
flow. A reported comparison between lattice and grain
boundary diffusions [Yamazaki et al., 2000] showed that the
enhancements were not so significant; lattice diffusion is
considered to be dominant in cases when grain sizes are
larger than a few tens of micrometers. Using these assump-
tions, the Deff of MgSiO3 perovskite in the lower mantle and Figure 2. Comparison of isobaric (25 GPa) thermal expan-
the self‐diffusion coefficient derived by MD simulations in sions of MgSiO3 perovskite. The solid squares show data of
this study (D) are related by this study, and the crosses show data of Funamori et al.
[1996]. The data of Funamori et al. [1996] were picked
Deff ¼ ðCeff =C ÞD; ð5Þ
up at 25 ± 0.5 GPa.

3 of 9
B12205 ITO AND TORIUMI: SILICON SELF‐DIFFUSION OF PEROVSKITE B12205

120 GPa [Gong et al., 2004]. Since the differences are


smaller than the difference between the experiments them-
selves, we believe that the model reproduced the compres-
sion very well. In Figure 2, results from this study are
compared with those of Funamori et al. [1996], who used a
multianvil apparatus. Better agreement (agreement within
3% of the values and their first derivatives (i.e., thermal
expansibility)) was confirmed. Orthorhombic distortions
were expressed by ratios of orthorhombic unit cell lengths
(b/a and c/a). Symmetry increases
pffiffiffi from orthorhombic to
cubic when b/a = 1 and c/a = 2 ffi 1.414; so differences in
values of b/a and c/a from these values correspond to the
amounts of orthorhombic distortions. In Figure 3, the
pressure dependencies of the ratios were compared with
those of experiments [Mao et al., 1991; Fiquet et al., 1998]
and other theoretical calculations [Marton et al., 2001;
Wentzcovitch et al., 1995] at various pressures (0–120 GPa).
The results of b/a in this study (1.02–1.015) are lower than
those of experiments (1.03). Results of the variational
induced breathing (VIB) model potential [Marton et al.,
2001] and zero‐temperature ab initio energy calculations
[Wentzcovitch et al., 1995] were also compared. The results
of the VIB model and this study are a little lower than those
of the experiments. The b/a distortion of ab initio calcula-
tions [Wentzcovitch et al., 1995] was more significant
(1.047–.081) than those of the experiments, and its c/a was
also larger than those of experiments. Despite the simplicity
of the potential model used in the study, it gave results
comparable to those of experiments.
[10] Silicon self‐diffusion coefficients for MgSiO3
perovskite were measured at four pressure conditions (40,
60, 80, and 100 GPa). At first, we examined the temperature
variations of self‐diffusion coefficients at 100 GPa, shown
in Figure 4. The temperature range is 4200–5900 K with a

Figure 3. Pressure dependences of orthorhombic distor-


tions (ratio of basic cell lengths b/a and c/a) of MgSiO3
perovskite at ordinary temperature (300 K). The crosses
and the plusses show experimental data [Mao et al., 1991;
Fiquet et al., 1998], and the circles and the squares show
theoretical data [Wentzcovitch et al., 1995; Marton et al.,
2001], respectively. Pressures of Fiquet et al. [1998] are
given by averages of multiple pressure scales (Pt and ruby
fluorescence). The 50 GPa data of Marton et al. [2001] are
given by 900 K data instead because of a lack of 300 K data.
Temperatures of energy calculations [Wentzcovitch et al.,
1995] were corresponding to 0 K because this calculation
does not treat finite temperatures.

[Mao et al., 1991], performed using diamond anvil cells


(DAC). Another experiment [Gong et al., 2004] consisted of
the shock compression of natural enstatite. Observed dif- Figure 4. Temperature dependences of silicon (Si) self‐
ferences of molar volumes between this study and the ex- diffusion coefficients at 100 GPa, 4200–5900 K. The ver-
periments [Yagi et al., 1979; Mao et al., 1991; Fiquet et al., tical line indicates the boundary of two (normal (lower),
2000] were under 3% at 25–70 GPa and under 5% at 40– abnormal (higher)) temperature regions.

4 of 9
B12205 ITO AND TORIUMI: SILICON SELF‐DIFFUSION OF PEROVSKITE B12205

Figure 5. Determination of the boundary of the lower and


higher regions at 100 GPa. (a) Original data of oxygen self‐
diffusion coefficients, (b) smoothed data, and (c) second de-
rivatives of the smoothed data. In Figure 5c, the solid arrow Figure 6. Temperature dependences of silicon (Si) self‐
shows the searching direction of the boundary, and the diffusion coefficients at 80 GPa (4200–5700 K), 60 GPa
dashed arrow shows the searched boundary. (3800—5100 K), 40 GPa (3800—4600 K). The vertical
lines indicate boundaries of the regions.

5 of 9
B12205 ITO AND TORIUMI: SILICON SELF‐DIFFUSION OF PEROVSKITE B12205

the same procedure, the temperatures of the boundaries were


derived as 4950 K (80 GPa) and 4550 K (60 GPa). At the
lower‐temperature areas (4200–4950 K and 3800–4550 K),
activation enthalpies (Hm*) were calculated as 376 kJ/mol
and 364 kJ/mol, respectively. Unlike the higher pressures,
the number of 40 GPa data was limited and the boundary
was difficult to derive by a deterministic procedure. We
hence placed the boundary at 4450 K by eye since there was
a small jump. The nondeterministic procedure contains
some ambiguity, and as a result we expected larger error in
the activation enthalpy of 40 GPa. Scattering at lower
temperatures was observed at 40 GPa. Although relative
absolute differences between neighborhood data points (R =
∣log10(D1) − log10(D0)∣/∣log10(D0)∣) were small (−0.25) at
4100–4450 K, the differences suddenly became large (−0.5)
at 4000–4050 K and lower temperatures (Figure 7). The
Figure 7. Relative absolute differences between neighbor- large differences at the lower temperatures exceeded 25%,
hood data points (R = ∣log10(D1) − log10(D0)∣/∣log10(D0)∣) of the acceptable upper limit of error [Ito and Toriumi, 2007].
40 GPa data. Hence, to avoid a large error, we did not adopt the data of
the scattered lower area. This treatment derives migration
enthalpy Hm* as 346 kJ/mol for 40 GPa at 4050–4450 K.
50 K step interval. The observed result had the following [12] From the available data (at 40, 60, 80, and 100 GPa),
features: (1) a linear increase in the lower‐temperature area pressure dependencies of the two coefficients (D0, Hm) in
and (2) a discontinuous rapid increase in the higher‐tem- equation (3) were expressed by functions of pressures.
perature area. Since melting occurred at a higher temper- Figure 8 plots observed migration enthalpies (Hm*) at the
ature of the higher area, we consider that the behavior of four pressures. Although the linearity of pressure variations
the higher‐temperature area was that of premelting and not was difficult to examine since the number of available data
solid‐state diffusion, which this study investigated. We points was only four, the data increased monotonically with
hence adopted the activation enthalpy of the lower area as increasing pressure. We then derived a linear relationship by
the value of migration enthalpy (Hm*). Because of the fitting the equation as follows:
margin of error of the data, the boundary of the two regions is
indistinct and hard to determine. We adopted the following Hm *ð PÞ½kJ=mol ¼ 0:875P½GPa þ 310: ð7Þ
procedure to give the boundary. First, we chose the data of
oxygen self‐diffusion coefficients to give the boundary.
Oxygen diffusion is two orders of magnitude faster than Si [13] As a result, the activation volume of migration Vm* =
self‐diffusion [Dobson et al., 2008]. Note that activation dHm*/dP was estimated to be 0.875 cm3/mol. The pre-
enthalpies of the oxygen diffusion data cannot be compared exponential term D0 varied greatly with an adopted Hm*, so
with those of experiments since this study does not treat the we used the linearly fitted Hm* (equation (7)) instead of the
most energetically favorable [Karki and Khanduja, 2006]
Mg‐O vacancy combinations. The error is reduced with
faster self‐diffusion coefficients [Ito and Toriumi, 2007].
The faster data gives a more distinct boundary. The routine to
derive the boundary is shown in Figure 5. The original data
of oxygen self‐diffusion coefficients (Figure 5a) were
smoothed by the filter (fnew(x) = f(x + Dx)/4 + f(x)/2 + f(x −
Dx)/4), where Dx represents finite increment of x (Figure
5b), and then second derivatives of log10D were derived
(Figure 5c). In Figure 5c, the boundary was searched from
the maximum point to the lower temperatures (shown by the
solid arrow), and given by the first negative point (shown
by the dashed arrow). On the basis of the routine, the
temperature of the boundary was given as 5200 K, and the
activation enthalpy of the lower‐temperature area (4200–
5200 K) was calculated as 400 kJ/mol. Belonoshko [1994]
calculated the melting temperature of MgSiO3 perovskite
as 5300 K at 110 GPa. The temperature of the boundary is
consistent with the melting temperature, supporting the idea
that the origin of the boundary is premelting. Figure 8. Temperature dependences of migration enthal-
[11] The temperature dependences of Si self‐diffusion pies (Hm*) of silicon (Si) self‐diffusion coefficients. Solid
coefficients were also examined at other pressures (40, 60, diamonds indicate this study. The open circle and cross
and 80 GPa). The results are shown in Figure 6. The features show experimental data [Yamazaki et al., 2000; Dobson
observed at 100 GPa were also seen at 80 GPa and 60 GPa. By et al., 2008] at 25 GPa, respectively.

6 of 9
B12205 ITO AND TORIUMI: SILICON SELF‐DIFFUSION OF PEROVSKITE B12205

equation (7). By employing this equation, Hm* of MgSiO3


perovskite at the pressure at the top of the lower mantle
(25 GPa) was derived to be 332 kJ/mol. This result can be
compared with the activation enthalpies of the experi-
mental studies [Yamazaki et al., 2000; Dobson et al., 2008]
directly. In the experiments, we consider that Si diffusion
was extrinsic as stated in section 2.2. We also consider that
Si vacancy concentrations can be assumed to be constant
with temperature and pressure variations since elements
with variable oxidation states (e.g., Fe) were not included
in the experimental samples. This result was consistent
with the results of experimental studies (336 kJ/mol at
25 GPa [Yamazaki et al., 2000], 347 kJ/mol at 25 GPa
[Dobson et al., 2008]). Hence, we first consider that the
reproduction of spontaneous jumping, which this study em-
Figure 9. Pressure dependence of preexponential terms ployed, is a valid approach for investigating self‐diffusion.
(D0) estimated by minimizing the least squares difference On the other hand, the previous approaches (733 kJ/mol at
(equation (8)). 0 GPa in the work by Wright and Price [1993]; 657kJ/mol
at 30 GPa in the work by Karki and Khanduja [2007];
453 kJ/mol at 25 GPa in the work by Ammann et al.
observed Hm* to avoid the effect of the error of the observed [2008]) somewhat overestimated the experimental results.
Hm* on D0. D0 was estimated using the following proce- According to the thermally activated process model,
dure. For each pressure condition, (1) Hm* was obtained by activation enthalpies are related with saddle point con-
equation (7); (2) D0,min was derived by minimizing fol- figuration. In this study, atoms spontaneously jump into
lowing the square difference as follows: vacancies, and the saddle points are naturally configured
Xh  i 2 without any assumptions. Hence, although detail of the
Di  D0;min exp Hm* =RTi ; ð8Þ difference of the saddle point configurations is still unclear,
i the authors suppose the assumption‐free saddle point con-
figurations as a cause of the consistency. The potential model
where Di and Ti are the observed self‐diffusion coefficients and parameters employed in this study were optimized for
and the corresponding temperature, respectively; and (3) equilibrium properties of experiments [Oganov et al., 2000].
the derived D0,min was adopted as D0. With this procedure, There is a weak point of the potential model. Although
D0 was derived as 2.92 × 10−9 [m2/s] at 40 GPa, 4.17 × coordination numbers and bond angles of atoms vary during
10−9 [m2/s] at 60 GPa, 6.83 × 10−9 [m2/s] at 80 GPa, and their migrations, the variations were not taken into account
9.60 × 10−9 [m2/s] at 100 GPa. The parameters increased during the optimization. However, from the agreement
log linearly with increasing pressure as seen in Figure 9. A between activation enthalpies of this study and those of the
log linear fitting of the data derived the pressure (P) experiments, it is implied that the weak point does not exert
dependence of D0 as much effect.

3
[15] The migration enthalpies (Hm*) increase monotoni-
D0 m2 =s ¼ 108:82510 P½GPa8:892 : ð9Þ cally with increasing pressure. This behavior is different
from those of O and Mg of MgO periclase [Ito and Toriumi,
Although the temperatures of the MD simulations were
much higher than those of the lower mantle, the results
could be extrapolated using equation (3). We derived the
depth profile of the viscosity of MgSiO3 perovskite in the
lower mantle by extrapolation. Regarding the geotherm, we
assumed an adiabatic temperature gradient (0.3 K/km) and
the temperature of decomposition of ringwoodite to
perovskite and periclase at the top of the lower mantle
(1900 K at 660 km [Ito and Katsura, 1989]). Figure 10
compares the derived viscosity profile with that of MgO
periclase [Ito and Toriumi, 2007]. The viscosity of
MgSiO3 perovskite increased log linearly with increasing
depth up to 16 times from the top of the lower mantle. On
the other hand, MgO periclase increased slightly (2.3
times) in the upper area of the lower mantle and began to
decrease with increasing depth.

4. Discussion
[14] We obtained an equation for migration enthalpies Figure 10. Depth variations of relative viscosities (h/
Hm* of silicon self‐diffusion of MgSiO3 perovskite as h660km) of the lower mantle minerals.

7 of 9
B12205 ITO AND TORIUMI: SILICON SELF‐DIFFUSION OF PEROVSKITE B12205

2007], which show negative activation volumes at higher Dobson, D. P., R. Dohmen, and M. Wiedenbeck (2008), Self‐diffusion of
pressures. Karato [1978, 2008] indicated the possibility of oxygen and silicon in MgSiO3 perovskite, Earth Planet. Sci. Lett., 270,
125–129, doi:10.1016/j.epsl.2008.03.029.
transitions of diffusion mechanisms from vacancies to Ewald, P. P. (1921), Die Berechnung optischer und electrostatischer Gitter-
interstitial atoms with increasing pressure, causing negative potentiale, Ann. Phys., 369, 253–287, doi:10.1002/andp.19213690304.
activation volumes at high pressures. Karato [1978] also Fiquet, G., D. Andrault, A. Daweale, T. Charpin, M. Kunz, and D.
Haussermann (1998), P‐V‐T equation of state of MgSiO3‐perovskite,
indicated “soft” interatomic potential favors causing such Phys. Earth Planet. Inter., 105, 21–31, doi:10.1016/S0031-9201(97)
transitions. Since Hm* of silicon of MgSiO3 perovskite 00077-0.
(400 kJ/mol at 100 GPa) is higher than those of O and Mg Fiquet, G., A. Dewaele, D. Andrault, M. Kunz, and T. Bihan (2000), Ther-
of MgO periclase (261 and 245 kJ/mol at 100 GPa, moelastic properties and crystal structure of MgSiO3 perovskite at lower
mantle pressure and temperature conditions, Geophys. Res. Lett., 27(1),
respectively), the interatomic potential of MgO periclase is 21–24, doi:10.1029/1999GL008397.
considered to be softer than that of MgSiO3 perovskite. So Funamori, N., T. Yagi, W. Utsumi, T. Kondo, T. Uchida, and M. Funamori
we consider that our results [this study; Ito and Toriumi, (1996), Thermoelastic properties of MgSiO3 perovskite determined by in
situ X ray observations up to 30 GPa and 2000 K, J. Geophys. Res., 101,
2007] support the suggestions made by Karato [1978]. 8257–8269, doi:10.1029/95JB03732.
Although these studies did not confirm the hypothesis Gong, Z., Y. Fei, Y. Dai, L. Zhang, and F. Jing (2004), Equation of state
directly, further direct observation of the transition by MD and phase stability of mantle perovskite up to 140 GPa shock pressure
simulation might be possible. and its geophysical implications, Geophys. Res. Lett., 31, L04614,
doi:10.1029/2003GL019132.
[16] The obtained viscosity profile of MgSiO3 perov- Heitjans, P., and J. Kaerger (2004), Diffusion in Condensed Matter:
skite shows a monotonic increase with increasing depth Methods, Materials, Models, Springer, Berlin.
(Figure 10). On the other hand, the viscosity profile of Holzapfel, C., D. Rubie, D. Frost, and F. Langenhorst (2005), Fe‐Mg inter-
diffusion in (Mg,Fe)SiO3 perovskite and lower mantle reequilibration,
MgO periclase is hill shaped; it increases in the upper area of Science, 309(5741), 1707–1710, doi:10.1126/science.1111895.
the lower mantle (660–1160 km) and decreases in the deeper Ita, J., and R. E. Cohen (1997), Effects of pressure on diffusion and
area (1160–2700 km) with increasing depth. Yamazaki and vacancy formation in MgO from nonempirical free‐energy integrations,
Phys. Rev. Lett., 79, 3198–3201, doi:10.1103/PhysRevLett.79.3198.
Karato [2001] approximated the lower mantle by a two‐ Ito, E., and T. Katsura (1989), A temperature profile of the mantle transi-
phase system composed of MgSiO3 perovskite and MgO tion zone, Geophys. Res. Lett., 25, 1095–1098.
periclase. The model rheology for the entire lower mantle is Ito, Y., and M. Toriumi (2007), Pressure effect of self‐diffusion in periclase
based on two microstructures: (1) a load‐bearing framework (MgO) by molecular dynamics, J. Geophys. Res., 112, B04206,
doi:10.1029/2005JB003685.
(LBF) of a strong phase containing isolated pockets of a Karato, S. (1978), The concentration minimum of point defects in solid at
weak phase and (2) an interconnected layer of a weak phase high‐pressures and the plasticity of the lower mantle (in Japanese), paper
(IWL), separating boudins and clasts of a strong phase. presented at Annual Meeting, Seismol. Soc. Jpn., Tokyo.
Karato, S. (2008), Deformation of Earth Materials: An Introduction to the
MgSiO3 perovskite is the dominant mineral in the LFB case, Rheology of Solid Earth, Cambridge Univ. Press, London.
and MgO periclase is the dominant mineral in the IWL case. Karato, S., S. Zhang, and H. Wenk (1995), Superplasticity in Earth’s lower
The depth profiles of the two minerals are clearly different mantle: Evidence from seismic anisotropy and rock physics, Science,
and well distinguished as shown in Figure 10. On the basis of 270, 458–461, doi:10.1126/science.270.5235.458.
Karki, B. B., and G. Khanduja (2006), Computer simulation and visualiza-
the difference, we could use the profiles to determine which tion of vacancy defects in MgSiO3 perovskite, Model. Simul. Mater. Sci.
mineral dominates the lower mantle rheology by assuming Eng., 14, 1041–1052, doi:10.1088/0965-0393/14/6/011.
the depth profile of lower mantle viscosity. Monotonic Karki, B. B., and G. Khanduja (2007), A computational study of ionic
vacancies and diffusion in MgSiO3 perovskite and post‐perovskite, Earth
nearly constant depth profiles were considered by studies of Planet. Sci. Lett., 260, 201–211, doi:10.1016/j.epsl.2007.05.031.
geodynamic modeling [e.g., Mitrovica and Forte, 1997]. The Karki, B., L. Stixrude, and R. Wentzcovitch (2001), High‐pressure elastic
shapes of the profiles are similar to those of MgSiO3 properties of major materials of Earth’s mantle from first principles, Rev.
Geophys., 39(4), 507–534, doi:10.1029/2000RG000088.
perovskite, and we consider MgSiO3 perovskite as dominant Mao, H. K., R. J. Hemley, Y. Fei, J. F. Shu, L. C. Chen, A. P. Jephcoat, and
in that case. The possibility of a hill shape, in which the Y. Wu (1991), Effect of pressure, temperature, and composition on lattice
highest‐viscosity zone is in the middle of the lower mantle, parameters and density of (Fe,Mg)SiO‐perovskites to 30 GPa, J. Geo-
has previously been considered [Morra et al., 2009]. Since phys. Res., 96, 8069–8079, doi:10.1029/91JB00176.
Marton, F. C., J. Ita, and R. E. Cohen (2001), Pressure‐volume‐temperature
the profile corresponds to that of MgO periclase, we consider equation of state of MgSiO3 perovskite from molecular dynamics and con-
that the lower mantle rheology is dominated by MgO peri- straints on lower mantle composition, J. Geophys. Res., 106, 8615–8627,
clase in that case. doi:10.1029/2000JB900457.
Mitrovica, J. X., and A. M. Forte (1997), The radial profile of mantle vis-
cosity: Results from the Joint Inversion of convection and postglacial
rebound observables, J. Geophys. Res., 102, 2751–2769, doi:10.1029/
[17] Acknowledgment. Some computations in this study were per- 96JB03175.
formed on Earth Simulator at Earth Simulator Center, Japan Agency for Morra, G., D. A. Yuen, C. Philippe, P. Koumoutsakos, L. Boschi, and P.J.
Marine‐Earth Science and Technology (JAMSTEC). Tackley (2009), The role of a high viscosity hill in the lower mantle on
the fate of subducting slabs, Geophys. Res. Abstr., 11, Abstract
EGU2009‐13553.
Oganov, A. R., J. P. Brodholt, and G. D. Price (2000), Comparative study
References of quasiharmonic lattice dynamics, molecular dynamics and Debye
Ammann, M. W., J. P. Brodholt, and D. P. Dobson (2008), DFT study of model applied to MgSiO3 perovskite, Phys. Earth Planet. Inter., 122,
migration enthalpies in MgSiO 3 perovskite, Phys. Chem. Miner., 36, 277–288, doi:10.1016/S0031-9201(00)00197-7.
151–158, doi:10.1007/s00269-008-0265-z. Poirier, J. P. (2000), Introduction to the Physics of the Earth’s Interior,
Bejina, F., O. Jaoul, and R. C. Liebermann (2003), Diffusion in minerals at Cambridge Univ. Press, Cambridge, U. K.
high pressure: A review, Phys. Earth Planet. Inter., 139, 3–20, Wentzcovitch, R. M., N. L. Ross, and G. D. Price (1995), Ab initio study of
doi:10.1016/S0031-9201(03)00140-7. MgSiO3 and CaSiO3 perovskites at lower‐mantle pressures, Phys. Earth
Belonoshko, A. B. (1994), Molecular‐dynamics of MgSiO3 perovskite at Planet. Inter., 90, 101–112, doi:10.1016/0031-9201(94)03001-Y.
high‐pressures: Equation of state, structure, and melting transition, Geo- Woodcock, L. V. (1971), Isothermal molecular dynamics calculations for
chim. Cosmochim. Acta, 58, 4039–4047, doi:10.1016/0016-7037(94) liquid salts, Chem. Phys. Lett., 10, 257–261, doi:10.1016/0009-2614
90265-8. (71)80281-6.

8 of 9
B12205 ITO AND TORIUMI: SILICON SELF‐DIFFUSION OF PEROVSKITE B12205

Wright, K., and G. D. Price (1993), Computer simulations of defects and Yamazaki, D., T. Kato, and H. Yurimoto (2000), Silicon self‐diffusion in
diffusion in perovskites, J. Geophys. Res., 98, 22,245–22,253, MgSiO3 perovskite at 25 GPa, Phys. Earth Planet. Inter., 119, 299–309,
doi:10.1029/93JB01098. doi:10.1016/S0031-9201(00)00135-7.
Yagi, T., H. K. Mao, and P. M. Bell (1979), Lattice parameters and specific
volume for perovskite phase of orthopyroxene composition (Mg,Fe)
Y. Ito and M. Toriumi, Department of Complexity Science and
SiO3, Year Book Carnegie Inst. Washington, 78, 612–613. Engineering, Graduate School of Frontier Science, University of
Yamazaki, D., and S. Karato (2001), Some mineral physics constraints Tokyo, 5‐1‐5 Kashiwanoha, Kashiwa, Chiba, 277‐0882, Tokyo, Japan.
on the rheology and geothermal structure of Earth’s lower mantle, (ito@gaea.k.u‐tokyo.ac.jp)
Am. Mineral., 86(4), 385–391.
Yamazaki, D., T. Kato, E. Ohtani, and M. Toriumi (1996), Grain growth
rates of MgSiO3 perovskite and periclase under lower mantle conditions,
Science, 274, 2052–2054, doi:10.1126/science.274.5295.2052.

9 of 9

You might also like