You are on page 1of 7

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112, B04206, doi:10.

1029/2005JB003685, 2007
Click
Here
for
Full
Article

Pressure effect of self-diffusion in periclase (MgO)


by molecular dynamics
Yosuke Ito1 and Mitsuhiro Toriumi2
Received 21 February 2005; revised 11 August 2006; accepted 27 October 2006; published 27 April 2007.

[1] Pressure effect of self-diffusion in periclase (MgO) was investigated by molecular


dynamics (MD), and viscosity variation of lower mantle periclase was predicted by
employing the Nabarro-Herring model. Self-diffusion coefficients were derived
directly from recorded trajectories of atoms. The MD calculations were performed at
2400–6000 K and 0–140 GPa. Migration energy (Hm*) and formation energy (Hf*) of
ordinary pressure (0 GPa) are consistent with experimental data. Observed pressure
dependence of self-diffusion coefficients was turned from negative with increasing
pressure. The behavior of self-diffusion coefficients predicts softening of lower mantle
periclase with increasing depth. We predict 1.3 times reduction of the viscosity
(0.3 K/km geothermal gradient) or 2.0 times (0.6 K/km geothermal gradient).
Citation: Ito, Y., and M. Toriumi (2007), Pressure effect of self-diffusion in periclase (MgO) by molecular dynamics, J. Geophys.
Res., 112, B04206, doi:10.1029/2005JB003685.

1. Introduction to reproduce the spontaneous vacancy migration by MD


(molecular dynamics) and to derive self-diffusion coeffi-
[2] Diffusion limits many kinetic phenomena, which are cients directly from recorded trajectories of atoms. Although
important in Earth and planetary interiors, such as creep of this approach does not need to treat the attempt frequency
solid phases and chemical mass transportation. Periclase explicitly, it is known that it is difficult to be precise about
(MgO), which is thought to be a minor component of the self-diffusion coefficient because frequencies of atom
Earth’s lower mantle (20%), may have a strong influence migrations are truncated due to the short (e.g., 100 ps) MD
on its rheological property because periclase is weaker than running time. Akamatsu and Kawamura [1999] derived
the major component perovskite (MgSiO3) [Yamazaki and self-diffusion coefficients of noble gas atoms in MgO, but
Karato, 2001]. Diffusion in periclase has been studied
they also reported that their results have large errors, up to 3
extensively in the materials science community, but the
orders of magnitude.
pressure dependence has hardly been considered. Recently,
[3] In this paper, we report on molecular dynamics
Van Orman et al. [2003] measured self-diffusion coeffi- (MD) calculations, which simulate self-diffusion in peri-
cients of periclase at 15– 25 GPa pressures, though the clase. We selected the latter approach (reproducing sponta-
pressures in the lower mantle exceed 25 GPa. Because of neous vacancy migrations) and found that one can achieve a
experimental difficulties theoretical approaches have been certain precision of the self-diffusion coefficient through
employed to derive self-diffusion coefficients, though those long (20 ns) running times. The Buckingham model was
lead to other difficulties. Since self-diffusion is an activated adopted as an interatomic potential. A pair of vacancies
mechanism, the self-diffusion coefficient can be separated (Mg and O atoms) was initially introduced into the MD
into two elements; the height of the energy barrier (i.e., system. MD calculations were performed at 2400 –6000 K
difference between the maximum and minimum free ener- temperatures and 0– 200 GPa pressures. Results were com-
gies during atom migration) and the attempt frequency of pared with past experimental and theoretical studies. The
atom migration. The difficulty is that no method to derive lack of seismic anisotropy in the lower mantle indicates
the attempt frequency is known. Ita and Cohen [1997] diffusion creep [Karato et al., 1995]. We derived the
measured the heights of energy barriers by free energy viscosity of lower mantle periclase by employing the
integrations, and they estimated the attempt frequency as Nabarro-Herring model of diffusion creep.
being equal to the lowest peak of the frequency spectrum,
which is obtained from the Fourier transform of trajectories
of diffusing atoms, though the relation between this peak 2. Methods
and the attempt frequency is not clear. Another approach is [4] In our MD simulations, the Buckingham model was
adopted to define the interatomic potential. Formulation of
1
Department of Earth and Planetary Science, Graduate School of the model is
Science, University of Tokyo, Tokyo, Japan.
2
Department of Complexity Science and Engineering, Graduate School  
  qq   ðai þ aj rij Þ CC
of Frontier Science, University of Tokyo, Tokyo, Japan. Uij rij ¼ riij j þ f bi þ bj exp b þ b  ri 6 j ð1Þ
ð i jÞ
Copyright 2007 by the American Geophysical Union.
0148-0227/07/2005JB003685$09.00

B04206 1 of 7
B04206 ITO AND TORIUMI: SELF-DIFFUSION IN PERICLASE (MgO) B04206

Nose-Hoover method or Andersen method was used, the


equations of motions of heat baths (Nose-Hoover method)
or boundaries of basic cells (Anderson method) sometimes
diverged when atoms migrate into vacancies. We consider
that the divergence is due to the rapid increase of system
energy when atoms reach saddle points of their migration
path. In order to avoid the divergence, we selected the
scaling method. The scaling method is more stable because
it does not solve any additional equations of motion. The
MD program MXDTRICL (K. Kawamura, 1996, available
at http://www.sccj.net/download/JCPE/) was used for cal-
culations.
[6] The self-diffusion coefficient (D) was derived by
Einstein equation of which the formation is
Figure 1. Thermal expansions of molar volumes under  
ordinary pressure (0 GPa). D ¼ ðd=dtÞ < jRn j2 > =6 ð2Þ

where Uij is the interatomic potential, rij is the distance where Rn indicates the positional vector of nth atom, while
between ith and jth atoms, and q, a, b, c are parameters. The the angle brackets indicate the average of all atoms. The
parameters were those of Akamatsu and Kawamura [1999]. positional vectors were given by the trajectories of atoms
The Buckingham model is simple. It requires little recorded during MD calculations. Initial relaxation was
calculation and successfully reproduces the equation of performed for 2 ns (106 steps) to achieve thermal
state (EOS) of periclase [Akamatsu and Kawamura, 1999]. equilibrium. After the initial relaxation, 2 ns (106 steps)
Because details of the reproduction are not given by MD was repeated 10 times and the average of the 10 self-
Akamatsu and Kawamura [1999], we reexamined the EOS. diffusion coefficients (D) was adopted as result. To study
Three different sets of EOS are compared with those of the precision of the averaged result of D, we calculated
experiments in Figures 1, 2, and 3: (1) thermal expansions standard deviations of the 10 self-diffusion coefficients (D)
under ordinary pressure (0 GPa), (2) compression under at the lowest temperature conditions (2400 K, 3400 K) and
ordinary temperature (300 K), and (3) compression under several (0 GPa, 50 GPa, 100 GPa, and 140 GPa) pressure
high temperature (2000 K) of molar volumes, respectively. conditions. We confirmed that the differences are smaller
In the first two cases the molar volumes (Figures 1 and 2) than 25% of the averages. We also confirmed that the ratios
are in excellent agreement with those of the experiments. between the standard deviations and the averages mono-
The last case the molar volumes (Figure 3) are a little tonously decreased with increasing temperatures. Therefore
smaller than those of experiments, though the difference is the standard deviations of the 10 self-diffusion coefficients
small (1% at 50 GPa). Therefore we consider that the are considered to be always smaller than 25%, which is
reproduction of the EOS of the model is good and self- reliable enough.
diffusion coefficients calculated by the model are reliable [7] We also calculated the formation enthalpy (Hf*) of the
enough to provide base data for geophysics. It is known that vacancy pair. In this study, the Hf* were derived from
ab initio electronic state calculations (reviewed by Karki et differences of total enthalpies between perfect and vacancy
al. [2001]) successfully reproduce well the elasticity at high containing MD systems. Initial relaxation was performed
pressures, but they require much more calculation. As for 0.2 ns (105 steps). After the initial relaxation, MD
mentioned here below, the calculation of self-diffusion calculation were performed for 0.2 ns (105 steps) recording
coefficients needs an ensemble of many (107) steps, so it is total enthalpies of each step and the average was adopted as
considered that the electronic state calculations are not
appropriate for this study.
[5] The MD system prepared was composed of 5  5 
5 unit cells of periclase (rock salt structure), which contain
4 Mg atoms and 4 O atoms per unit cell, with a pair
vacancies (998 atoms) and three-dimensional periodic
boundary conditions. The potential, which an atom feels,
was given by infinite summation of potentials between the
atom and the other atoms. The Coulomb term (first term of
the equation of Buckingham model) was summed by the
Ewald method [Ewald, 1921]. Summations of short-range
(second and third) terms were cut off at 7.5 Å. Newton’s
equations of motions were integrated by the Verlet Algo-
rithm with 2 fs intervals. Isothermal and isobaric (NPT)
ensembles were realized by the scaling method [Woodcock,
1971]. Thermodynamically rigorous methods such as the
Nose-Hoover method [Nose, 1984] or the Andersen method
Figure 2. Compressions of molar volumes under ordinary
[Andersen, 1980] are more preferable. However, when the
temperature (300 K).

2 of 7
B04206 ITO AND TORIUMI: SELF-DIFFUSION IN PERICLASE (MgO) B04206

proportionally on the vacancy concentrations C. Intrinsic


vacancy concentrations are given by equation (4). Therefore, by
replacing Cint,O (or Cint,Mg) for C in equation (3), the following
relation is derived:
   
D / exp  Hm* þ H*f =2 =RT ð5Þ

Equation (5) means that the activation energies of the intrinsic


regime are interpreted as Hm* + Hf*/2. Up to now, the
introduction of the equation (3) or (5) is the same as in the
work by Ando [1989].
[9] In this simulation, the natural system is approximated
by three-dimensional infinite repetitions of the small sys-
tems (5  5  5 unit cells with a pair vacancies (998 atoms)
Figure 3. Compressions of molar volumes under high for self-diffusion or 10  10  10 unit cells with a pair
temperature (2000 K). vacancies or without any vacancy (7998 or 8000 atoms)).
The vacancy concentrations of the small systems (1/500 for
result. We are able to use larger system than the calculations 5  5  5 unit cells or 1/4000 for 10  10  10 unit cells)
of the self-diffusion coefficient, because the calculation for are much higher than that of natural systems. This high
Hf* needs shorter ensemble time (0.2 ns) than that for the vacancy concentrations may affect the results of self-
self-diffusion coefficient (2 ns). The size of the MD systems diffusion coefficients D or vacancy formation energy Hf*
is 10  10  10 unit cells containing 8000 and 7998 atoms because of interactions between vacancies. When there is no
for the perfect system and for the one with a pair of size effect, the self-diffusion coefficient D is inversely pro-
vacancies, respectively. portional to the vacancy concentration as of equation (3).
[8] In ionic crystals such as MgO, the dominant mecha- In order to examine whether D is inversely proportional
nism of the self-diffusion is vacancy hopping. The vacancy with the vacancy concentration or not, we calculated D at
hopping is classified in two types; extrinsic and intrinsic. the small size (5  5  5 unit cells with a pair vacancies;
The ‘‘extrinsic’’ indicates that vacancies are formed by 998 atoms) and the large size (8  8  8 unit cells with a
external sources and the ‘‘intrinsic’’ means that vacancies pair vacancies; 4094 atoms) at several P-T conditions and
are thermally formed. Typical external sources are impurity compared D normalized by vacancy concentrations. The
atoms, which have different charges than host atoms. For ratio of the normalized D998/D4094 (where D998 indicates the
keeping charge neutrality, in MgO two trivalent impurity usual size (998 atoms) and D4094 indicates the large size
atoms produce one Mg vacancy or two monovalent impu- (4094 atoms)) is distributed from 77% to 144%. The wide
rity atoms produce one O vacancy. If there is no external distribution seems to come from the error of D (25%) and it
sink or source of the impurities, the extrinsic vacancy suggests that the size effect of the D is below the margin.
concentrations do not depend on temperature or pressure. We also studied the size effect for the vacancy forma-
In this case, the self-diffusion coefficient is proportional to tion energy Hf* by comparing Hf* of the normal system
the self-diffusion coefficient of the MD calculations because (10  10  10 unit cells with a vacancy pair, 7998 atoms)
in the both cases the vacancy concentrations are constant. with that of the large system (16  16  16 unit cells with a
The proportional relation can be written as D = (C/C0)*DC0 vacancy pair, 32,766 atoms). Note that the Hf* does not
where D and C are the self-diffusion coefficient and depend on the system size because we adopt kJ/mol as the
vacancy concentration of the extrinsic system and DC0 unit. The observed Hf* of the large system is 2% smaller
and C0 is that of MD calculations, respectively. In addition, than that of the normal system at 300 K and 0 GPa, which
this proportional relation postulates that the activation
energies of the extrinsic system and the MD calculations
can be considered to be equal. This relation is expressed as

D ¼ ðC=C0 Þ*DC0;0 expðHm*=RT Þ ð3Þ

where DC0,0 is the prefactor, Hm* are the activation energies


of the MD self-diffusion coefficients. The Hm* are often called
the migration energies. When intrinsic, the O and Mg vacancy
concentrations (Cint,O, Cint,Mg) depend on Hf* as Cint,O Cint, Mg
/ exp(Hf*/RT ). The vacancy concentration of O and Mg is
considered to be the same because of charge neutrality.
Therefore this relation can be written as
   
Cint;0 ¼ Cint;Mg / exp  H*f =2 =RT ð4Þ

Equation (3) can be interpreted as (1) the diffusion coefficients


D depend exponentially on Hm* and (2) the D also depend Figure 4. Trajectories of atoms observed by MD simula-
tions under 2500 K and 0 GPa.

3 of 7
B04206 ITO AND TORIUMI: SELF-DIFFUSION IN PERICLASE (MgO) B04206

destroyed because the system does not have any external


sources of vacancies. There are also no dislocations or grain
boundaries in the system because the system consists of 5 
5  5 MgO unit cells with a pair of vacancies and their
infinite three-dimensional repetitions.
[12] Table 1 compares Hm* with the activation enthalpies
of three experimental [Sempolinsky and Kingery, 1980;
Ando et al., 1983; Yang and Flynn, 1994] and one theoret-
ical study [Ita and Cohen, 1997]. Table 1 also compares
formation enthalpies (Hf*) given by the total energy differ-
ences of perfect and vacancy containing crystals. In Table 1,
we interpret the activation enthalpies derived from experi-
ments of Mg and O self-diffusions as Hm* and Hm* + Hf*/2,
respectively, because the Mg and O self-diffusions are
thought to be extrinsic and intrinsic, respectively. We will
discuss the interpretation in the next section. Observed Hm*
Figure 5. Self-diffusion coefficients (D) under 2400 – and Hf* were also consistent with those of experimental
3000 K temperatures at ordinary pressure (0 GPa). studies. Hm* of this study and Ita and Cohen [1997] are
indistinguishable, though Hf* of our study is larger than that
of Ita and Cohen [1997]. We suppose that the difference
seems to be small. Finally, we examined the effect of the
between the Hf* comes from the differences in the inter-
integration interval by comparing the self-diffusion coeffi-
atomic potentials because the two studies gave formation
cient D for the normal integration interval (2 fs) with that
enthalpies by the same methods (differences of total ener-
for the short integration interval (1 fs). The ratio of the D2fs/
gies between perfect and vacancy containing systems).
D1fs (where D2fs and D1fs indicate normal integration
[13] We observed self-diffusion coefficients under various
interval (2 fs) and shorter integration interval (1 fs), respec-
temperatures (2400 – 6000 K) and pressures (0 –140 GPa).
tively) is distributed from 86% to 125%, which seems to be
Like under ordinary pressure (0 GPa), vacancies did not
within the margin of error (25%).
migrate when temperatures were too low and the systems
melted when temperatures were too high. At each pressure,
3. Results self-diffusion coefficients were fitted to the Arrhenius
[10] First, we performed MD calculations under various equations. Table 2 shows the fitted results of preexponential
(2000 – 3400 K) temperatures at ordinary pressure (0 GPa) factors (D0) and Hm* over the temperature range at each
with 100 K intervals. Figure 4 shows trajectories of atoms pressure. We calculated squared multiple correlation coef-
of MD calculation at 2500 K. In Figure 4, an O atom and an ficients (s2) to estimate how well they fit. The correlation
Mg atom are migrating to O and Mg vacancies, respectively. coefficients s of the two arrays x = {x1, x2, . . . , xn} and y =
Such migrations were observed at 2400 –3000 K. Above {y , y2, . . . , yn} are given by s = (nSxy  SxSy)/
r1ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
h P
3100 K, the MD systems melted. The melting above 3100 K P ih P P i
n x2  ð xÞ2 n y2  ð yÞ2 . In this case, x
is consistent with the melting point of periclase (3100 K).
Under 2300 K, no vacancy migrations were observed. This and y correspond to temperature and ln(D), respectively.
is because the probability of vacancy migration was too Results of s2 are always close to 1 (over 0.99), indicating
small within the time span of the MD simulation. Note that that the self-diffusion coefficients at high pressures were
the running time of our MD simulations (2 ns) was much also linear in the Arrhenius plot. Pressure dependences
smaller than the typical experimental timescale of self- of self-diffusion coefficients under several temperatures
diffusion measurements (over 1 hour). Self-diffusion coef- (3000 K, 3400 K, 4400 K, and 5400 K) were drawn at
ficients were calculated for 2400– 3000 K. Figure 6. We found that the self-diffusion coefficients
[11] Figure 5 shows calculated self-diffusion coefficients decreased at low pressure, though the decrease slowed
at temperatures of 2400 – 3000 K at ordinary (0 GPa) down and rose with increasing pressure. The behavior of
pressure. The self-diffusion coefficients D lie on a straight self-diffusion coefficients was different from that of the
line in the Arrhenius (log(D)  1/T ) plot, where T is previous study [Ita and Cohen, 1997]. Their self-diffusion
temperature. This feature is consistent with experimental coefficients did not increase with increasing pressure. In
self-diffusion coefficients. Activation enthalpies (H *) were order to examine the difference of these two results, we
given by the Arrhenius equation D = D0exp(H */RT ), compared two components (preexponential factors D0 and
where D are the self-diffusion coefficients, D0 is the migration energies Hm*) of self-diffusion coefficients in
preexponential factor, R is the gas constant, and T is the
temperature, respectively. Derived activation enthalpies
were similar; 202 kJ/mol for Mg and 215 kJ/mol for O. Table 1. Comparison of H *m and H *m + Hf*/2
The similar activation enthalpies suggest that the diffusion H *m, kJ/mol H *m + H f*/2, kJ/mol
of Mg and O occurs as a pair. We compare the derived This study 202 586
activation enthalpies to migration energies (Hm*) because the Ita and Cohen [1997] 190 427
vacancy concentration does not vary during the simulation. Sempolinsky and Kingery [1980] 220 –
Ando et al. [1983] – 536
In the MD simulation, vacancies are not created or Yang and Flynn [1994] 243 667

4 of 7
B04206 ITO AND TORIUMI: SELF-DIFFUSION IN PERICLASE (MgO) B04206

Figure 6. Pressure variations (0 – 140 GPa) of self- Figure 8. Comparisons of migration enthalpies (Hm*) of
diffusion coefficients (D) under several (3000, 3400, Mg.
4400, 5400 K) temperatures.
given by the total enthalpy difference of perfect and defect
Figures 7 and 8, respectively. We also drew pressure containing systems.
derivatives of Hm* (that is, activation volumes of migrations
(Vm*), Vm* = dHm*/dP) in Figure 9 because Vm* can be com- 4. Discussion
pared with experimental data [Van Orman et al., 2003]).
Although Ita and Cohen’s [1997] D0 is constant due to [15] Our calculations predict that pressure dependencies
the constant attempt frequency which they assumed, the of self-diffusion coefficients (Figure 6) turn from negative
observed D0 decreases considerably with increasing pres- to positive with increasing pressure. This behavior may
sure (1.5 times from upper (25 GPa) to lower (135 GPa) seem strange because it is not consistent with previous
mantle pressures). This observation indicates that the defect analyses [Karato, 1981] or theoretical calculations
dependence of attempt frequency should not be ignored [Ita and Cohen, 1997]. However, there are several studies
during calculations of self-diffusion coefficients. Regarding that are consistent with this behavior. For example,
Hm* and Vm*, our Vm* went from negative to positive values at Gregoryanz et al. [2005] report a negative pressure depen-
60 GPa, indicating a rise of self-diffusion coefficients with dence of the sodium melting point at high pressure. A
increasing pressure. On the other hand, Vm* of Ita and Cohen negative pressure dependence of melting points would
[1997] did not turn negative. The difference corresponds to imply a positive pressure dependence of the self-diffusion
the difference between the self-diffusion coefficients shown coefficient because self-diffusion coefficients D and the
by Figure 6. At 20 GPa, Vm* of our study and of Ita and melting point Tm are related through the homogeneous
Cohen [1997] were indistinguishable and both Vm* were relation D = D0 exp (g Tm/T ), where D0 is preexponential
consistent with Vm* of Van Orman et al. [2003]. factor, g is a coefficient, and T is temperature. Although data
[14] Figure 10 compares the formation enthalpies (H*f ) of MgO melting experiments are available up to 35 GPa
with that of Ita and Cohen [1997]. Both Hf* increased with [Zerr and Boehler, 1994], MD calculations of MgO melt-
pressure, but our increase rate was larger than that of Ita and ing are performed by employing the two-phase method
Cohen [1997]. The cause of the difference is considered to [Belonoshko and Dubrovinsky, 1996] under whole lower
come from the interatomic potentials because both H*f are mantle pressures (up to 140 GPa). Increasing of their
melting points dramatically slows with increasing pressure
(increment is only 300 K between 60 and 140 GPa, though
3000 K between 0 and 60GP), which is a trend similar to
that of small positive activation volumes of self-diffusion

Figure 7. Comparisons of preexponential terms (D0) of Figure 9. Comparison of activation volumes of migration
Mg. (Vm*) of Mg.

5 of 7
B04206 ITO AND TORIUMI: SELF-DIFFUSION IN PERICLASE (MgO) B04206

Table 2. Fitted Results of Self-Diffusion Coefficients to Arrhenius Equation D = D0exp(H */RT )a


D0, m2/s H *m , kJ/mol s2
Pressure, GPa Temperature Range, K O Mg O Mg O Mg
0 2400 – 3000 9.05E-08 6.90E-08 215 202 0.9948 0.9961
10 2700 – 3400 3.44E-08 2.80E-08 230 215 0.9957 0.9930
20 2800 – 4200 3.52E-08 3.57E-08 257 250 0.9981 0.9978
30 3400 – 5000 2.77E-08 1.96E-08 269 252 0.9962 0.9983
40 3400 – 5000 2.17E-08 1.94E-08 277 266 0.9991 0.9987
50 3400 – 5200 2.33E-08 2.36E-08 292 287 0.9991 0.9990
60 3400 – 6000 2.04E-08 1.84E-08 297 288 0.9984 0.9968
70 3400 – 6000 1.55E-08 1.21E-08 294 279 0.9978 0.9987
80 3400 – 6000 1.24E-08 9.93E-09 290 276 0.9996 0.9983
90 3400 – 6000 7.34E-09 7.07E-09 272 265 0.9984 0.9976
100 3400 – 6000 5.39E-09 4.16E-09 261 245 0.9969 0.9989
110 3400 – 6000 4.82E-09 3.36E-09 257 236 0.9990 0.9977
120 3400 – 6000 3.46E-09 2.56E-09 243 226 0.9984 0.9978
130 3400 – 6000 2.35E-09 2.09E-09 226 215 0.9981 0.9979
140 3400 – 6000 2.19E-09 1.68E-09 223 205 0.9987 0.9994
a
D is self-diffusion coefficients, D0 is the preexponential factor, R is the gas constant, T is temperature; s2 is squared multiple correlation coefficient of
the least squares fitting. Temperature range is also shown at each pressure. Intervals of temperatures are 100 K (0 – 10 GPa) or 200 K (20 – 140 GPa). Read
9.05E-08 as 9.05  108.

coefficients reported in this manuscript. In addition, Karato [1998] speculated that O self-diffusion is extrinsic because
[1978] reports local minima of defect formation energies O vacancies are introduced through monovalent impurities.
with increasing pressures, supporting the observed pressure We considered that O diffusion of lower mantle periclase is
dependence of self-diffusion coefficients. not intrinsic but extrinsic, because Hf* rapidly increases
[16] We discuss a mechanism of MgO self-diffusion. In with increasing pressure (Figure 10) [Ita and Cohen, 1997].
ordinary pressure, it is considered that the Mg diffusion is Equation (4) says that the intrinsic vacancy concentration
extrinsic, because Mg vacancies are formed by trivalent exponentially decreases with increasing Hf*. Monovalent
impurities, while O diffusion is intrinsic [Ando, 1989]. It is impurities may be possible as a source of O vacancies
known that the Mg self-diffusion coefficient increases with because the pyrolite model [Ringwood, 1975] postulates
increasing trivalent impurity concentrations [Sempolinsky some (1%) sodium.
and Kingery, 1980] and that the O self-diffusion coefficient [17] The lack of seismic anisotropy indicates that the
does not depend on the trivalent impurity concentration lower mantle flows by diffusion creep [Karato et al.,
[Ando et al., 1983]. At high pressure, the mechanism of 1995]. In a material flowing by diffusion creep, viscosity
MgO self-diffusion has hardly been known. Van Orman and self-diffusion coefficients are related according to the
et al. [2003] suggested that Mg diffusion is extrinsic at 15– Nabarro-Herring model
25 GPa because the self-diffusion coefficients of Mg depend
 
on Al concentrations. It is considered that Mg diffusion in h ¼ A*ð RT=V Þ* G2 =Deff ð6Þ
lower mantle periclase is extrinsic because the pyrolite
model [Ringwood, 1975] indicates that the lower mantle where h is viscosity, A is a constant, R is the gas constant, T
has a certain (5%) Al content. O self-diffusion coefficients is temperature, V is the molar volume of the material, G is
at 15– 25 GPa [Van Orman et al., 2003] are larger than that the grain size, and Deff is the effective self-diffusion
at ordinary pressure [Ando et al., 1983; Yang and Flynn, coefficient, which is that of the most slowly diffusing
1994]. Van Orman et al. [2003] indicated that the O self- species in the mineral. Here we try to estimate the viscosity
diffusion is extrinsic where vacancies are created by dis-
locations generated during compression. Ita and Cohen

Figure 10. Comparison of vacancy formation energies Figure 11. Relative viscosity (h/h0) of the lower mantle
(Hf*). periclase stimated from Nabarro-Herring equation.

6 of 7
B04206 ITO AND TORIUMI: SELF-DIFFUSION IN PERICLASE (MgO) B04206

of the lower mantle periclase by employing the Nabarro- Ando, K., Y. Kurokawa, and Y. Oishi (1983), Oxygen self-diffusion in Fe-
doped MgO single crystals, J. Chem. Phys., 78, 6890 – 6892.
Herring model. We assume that (1) the most slowly Belonoshko, A., and L. Dubrovinsky (1996), Molecular dynamics of NaCl
diffusing species in periclase is O, (2) the O diffusion of (B1 and B2) and MgO (B1) melting: Two-phase simulation, Am. Mineral.,
lower mantle periclase is extrinsic, and (3) the vacancy 81, 303 – 316.
Dewaele, A., G. Fiquet, D. Andrault, and D. Hausermann (2000), P-V-T
concentration is always uniform. At ordinary pressure, the equation of state of periclase from synchrotron radiation measurements,
Mg diffusion is 2– 3 times faster than the O diffusion [Ando, J. Geophys. Res., 105, 2869 – 2877.
1989]. If the sources of the O vacancies are monovalent Dubrovinsky, L., and S. Saxena (1997), Thermal expansion of periclase
cations, the uniform assumption may be appropriate when (MgO) and tungsten (W) to melting temperatures, Phys. Chem. Mineral.,
24, 547 – 550.
neglecting partitioning of the monovalent cations between Ewald, P. P. (1921), Die Berechnung Optischer und Electrostatischer Git-
periclase and other lower mantle minerals such as terpotentiale, Ann. Phys., 64, 253 – 287.
perovskite. We can apply self-diffusion coefficients of the Fei, Y. (1999), Effects of temperature and composition on the bulk modulus
of (Mg, Fe)O, Am. Mineral., 84, 272 – 276.
MD calculations for the Deff of equation (6) by introducing Gregoryanz, E., O. Degtyareva, M. Somayaulu, R. J. Hemley, and H. K.
these three assumptions. This gives us three unknown Mao (1997), Melting of dense sodium, Phys. Rev. Lett., 94, 185502.
parameters (A, G in equation (6) and C in equation (3)). The Ita, J., and R. E. Cohen (1997), Effects of pressure on diffusion and vacancy
formation in MgO from non-empirical free-energy integrations, Phys.
coefficient A in the Nabarro-Herring equation is difficult to Rev. Lett., 79, 3198 – 3201.
estimate, so we eliminate A by using a proportional relation Ita, J., and R. E. Cohen (1998), Diffusion in MgO at high pressure: Im-
and the variable C is eliminated by introducing a plications for lower mantle rheology, Geophys. Res. Lett., 25, 1095 –
proportional relation. We already assume that the vacancy 1098.
Ito, E., and T. Katsura (1989), A temperature profile of the mantle transition
concentration is uniform. Assuming that the grain size G is zone, Geophys. Res. Lett., 16, 425 – 428.
constant in the lower mantle, we can derive the relationship Karato, S. (1978), The concentration minimum of point defects in solid at
between viscosities of lower mantle periclase h and self- high-pressures and the plasticity of the lower mantle (in Japanese), Seismol.
Soc. Jpn. Programme Abstr., 1, D31.
diffusion coefficients derived by the MD simulations D as Karato, S. (1981), Pressure dependence of diffusion in ionic solids, Phys.
Earth Planet. Inter., 25, 38 – 51.
h / ðV =RT Þ*ð1=DÞ ð7Þ Karato, S., S. Zhang, and H. Wenk (1995), Superplasticity in Earth’s lower
mantle: Evidence from seismic anisotropy and rock physics, Science,
270, 458 – 461.
where T is lower mantle geotherm, V is molar volume. The Karki, B., and G. Khanduja (2006), Vacancy defects in MgO at high pres-
grain growth rate of periclase is very small under high pres- sure, Am. Mineral., 91, 511 – 516.
Karki, B., L. Stixrude, and R. Wentzcovitch (2001), High-pressure elastic
sure [Yamazaki et al., 1996], indicating uniform grain size properties of major materials of Earth’s mantle from first principles, Rev.
of periclase in the lower mantle. The temperature at top of Geophys., 39, 507 – 534.
the lower mantle is 1900 K [Ito and Katsura, 1989]. We con- Mao, H., and P. Bell (1979), Equations of state of MgO and epsilon-Fe
sider adiabatic (0.3 K/km) and superadiabatic (0.6 K/km) under static pressure conditions, J. Geophys. Res., 84, 4533 – 4536.
Nose, S. (1984), A unified formulation of the constant temperature mole-
temperature gradients. We derive the relative viscosity of cular-dynamics methods, J. Chem. Phys., 81, 511 – 519.
lower mantle periclase by using equation (7). Figure 11 Ringwood, A. E. (1975), Composition and Petrology of the Earth’s Mantle,
shows the relative viscosity. We found that the viscosity of McGraw-Hill, New York.
Sempolinsky, D. R., and W. D. Kingery (1980), Ionic conductivity and
lower mantle periclase decreases with increasing depth. The magnesium vacancy mobility in magnesium oxide, J. Am. Ceram. Soc.,
rate of the softening is 1.3 times (0.3 K/km) or 2.0 times 63, 664 – 669.
(0.6 K/km). The softening of the lower mantle periclase Van Orman, J. A., Y. Fei, H. Hauri, and J. Wang (2003), Diffusion in MgO
at high pressures: Constraints on deformation mechanisms and chemical
comes from the self-diffusion coefficients decreasing at low transport at the core-mantle boundary, Geophys. Res. Lett., 30(2), 1056,
pressure, though the decrease slows down and turns into an doi:10.1029/2002GL016343.
increase with increasing pressure (Figure 6). Yamazaki and Woodcock, L. V. (1971), Isothermal molecular dynamics calculations for
Karato [2001] suggested that the viscosity of the lower liquid salts, Chem. Phys. Lett., 10, 257 – 261.
Yamazaki, D., and S. Karato (2001), Some mineral physics constraints on
mantle is near that of the weaker phase, periclase, in the the rheology and geothermal structure of Earth’s lower mantle, Am.
place most highly strained. The softening may be one of the Mineral., 86(4), 385 – 391.
causes of dynamic geological activities of the lower mantle, Yamazaki, D., T. Kato, E. Ohtani, and M. Toriumi (1996), Grain growth
rates of MgSiO3 perovskite and periclase under lower mantle conditions,
such as plume generation. Science, 274, 2052 – 2054.
Yang, M. H., and C. P. Flynn (1994), Intrinsic diffusion properties of an
[18] Acknowledgments. We thank K. Kawamura and T. Tsuchiya for oxide: MgO, Phys. Rev. Lett., 73, 1809 – 1812.
their helpful discussion. K. Kawamura kindly offered the MD program Zerr, A., and R. Boehler (1994), Constraints on the melting temperature of
MXDTRICL. Some computations were performed by Earth Simulator, at the lower mantle from high-pressure, experiments on MgO and magne-
Earth Simulator Center, Japan Agency for Marine-Earth Science and siowustite, Nature, 371, 506 – 508.
Technology (JAMSTEC).

References Y. Ito, Department of Earth and Planetary Science, Graduate School of
Science, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033,
Akamatsu, T., and K. Kawamura (1999), Molecular dynamics of solid Japan. (ito@gaea.k.u-tokyo.ac.jp)
solution and coexisting liquid, Mol. Simul., 21, 387 – 399. M. Toriumi, Department of Complexity Science and Engineering, Graduate
Andersen, H. C. (1980), Molecular dynamics simulations at constant pres- School of Frontier Science, University of Tokyo, 5-1-5 Kashiwanoha,
sure and/or temperature, J. Chem. Phys., 72, 2384 – 2393. Kashiwa, Chiba, Tokyo 277-8581, Japan. (tori@k.u-tokyo.ac.jp)
Ando, K. (1989), Self-diffusion in oxides, in Rheology of Solids and of the
Earth, pp. 57 – 82, Oxford Univ. Press, New York.

7 of 7

You might also like