You are on page 1of 11

PHYSICS

OFTHE EARTH
AND PLAN ETA RY
INTERIORS

ELSEVIER

Physics of the Earth and Planetary Interiors 87(1994) 111121

Synthesis and re-investigation of the elastic properties


of single-crystal magnesium silicate perovskite
Amir Yeganeh-Haeri

Centerfor High Pressure Research and Department of Earth and Space Sciences, State University of New York at Stony Brook,
Stony Brook, NY11794, USA
Received 4 January 1994; revision accepted 18 April 1994

Abstract
Single crystals of MgSiO

3 in the perovskite structure have been grown at a peak pressure of 26 GPa and
temperature of
1600 K using a 2000 ton uniaxial split-sphere high-pressure apparatus (USSA-2000). The
specimens were subsequently utilized to re-investigate the single-crystal elastic properties of this phase at ambient
conditions using laser Brillouin spectroscopy. The nine adiabatic single-crystal elastic stiffness coefficients, in units of
GPa, are: C11 = 482, C22 = 537, C33 = 485, C44 = 204, C55 = 186, C66 = 147, C12 = 144, C13 = 147, C23 = 146. The
resulting estimated VoigtReussHill (VRH) aggregate isotropic elastic moduli are: K = 264.0 and ~ = 177.3 GPa,
respectively. The single-crystal elastic moduli of MgSiO3 perovskite display a pattern that is elastically somewhat
anisotropic. The maximum shear and compressional velocities are 18% and 7% greater than the minimum. The [010]
crystallographic direction contains both the fastest and the slowest shear wave velocities. If, under lower mantle
conditions, magnesium silicate perovskite grains were to become preferentially oriented, a shear wave propagating in
the Earths lower mantle could become polarized with two distinct velocities. The observed density and seismic
parameter of the lower mantle over the depth range of 10002700 km are compared with the calculated profiles for
a model mantle consisting of pure perovskite (Mg089,Fe011)Si03 and for a mixture composed of silicate perovskite
and magnesiowstite using our new elasticity results. At present, literature values of thermoelastic properties for
silicate perovskite, in particular, the coefficient of thermal expansion and the temperature derivative of the
isothermal bulk modulus, vary widely. Because of this disparity, we find that mantle models ranging from pure
perovskite to pyrolitic-type compositions provide acceptable fits to the seismically observed density and velocity
profiles of the Earths lower mantle.

1. Introduction
Clearly an accurate characterization of the
elastic properties of silicate perovskite is of fun-

damental importance in elucidating the composition and mineralogy of the earths lower mantle.
Over the past decade there have been a number
of experimental investigations aimed at characterizing the elastic properties and equation-of-state

1 Present address: Seismological Laboratory, 252-21 Callfornia Institute of Technology, Pasadena, CA 91125, USA.
Fax: (818) 564 0715

parameters of (Mg,Fe)Si03 perovskite (Yagi et


al., 1982; Knittle and Jeanloz, 1987; Kudoh et al.,
1989; Yeganeh-Haeri et al., 1989a,b; Ross and

0031-9201/94/$07.00 ~ 1994 Elsevier Science B.V. All rights reserved


SSDI 0031-9201(94)02954-A

112

A. Yeganeh-Haeri /Physics of theEarth and Planetary Interiors 87 (1994) 111121

Hazen, 1990; Mao et al., 1991; for a summary see


also Hemley and Cohen, 1992). The largest difference in the reported value of bulk modulus
among all of the above-mentioned experimental
studies has been only about 7%. The recent synthesis of good-quality single crystals of MgSiO3
perovskite allowed us to more accurately constrain the single-crystal elastic properties of this
important phase.
The main purpose of this paper is to present
the newly determined single-crystal elastic moduli
of MgSiO3 in the perovskite structure. We also
use our elasticity data in conjunction with the
latest measurements on the thermoelastic properties of silicate perovskite and magnesiowstite
conducted at elevated pressures and temperatures (Mao et al., 1991; Fei et al. 1992; Funamori
and Yagi, 1993; Wang et al. 1993) to construct
density and velocity profiles for two compositional models over the depth range of 10002700
km. We then compare these calculated profiles
with those of the PREM model (Dziewonski and
Anderson, 1981) and briefly discuss the implications of these data on geophysically plausible
models for the chemical constitution of Earths
lower mantle.

2. Experimental techniques
2.1. Sample synthesis and description
The starting material used for syntheses was
prepared by mixing appropriate proportions of
tetra-ethyl-orthosilicate Si(0C2H5)4 of 99.99%
purity and an aqueous solution of magnesium
nitrate Mg(N03) 6H~Oof 99.99% purity with an
Si/Mg ratio of 1. The precipitated gels were then
gently heated over a period of 24 h, they were
further heat treated at 1220 K for approximately
2 h. After removal from the furnace the starting
material was cooled to ambient temperature and
pulverized in an agate mortar.
High-pressure synthesis runs were performed
using the 2000 ton uniaxial split-sphere apparatus
(USSA-2000) installed at Stony Brooks HighPressure Laboratory. This equipment has previously been described in detail by Gasparik (1989).
.

The cell assembly used in our experiment is identical to that developed by Ito and Weidner (1986).
The single crystals were grown at a peak pressure
of 26 GPa and temperature of
1600 K during a
45 mm run. The recovered run product consisted
of transparent, colorless, euhedral single crystals
with dimensions ranging from 20 x 20 x 30 to 40
x 40 x 30 jim, although some crystals had dimensions in excess of 100 ~m. Initial petrographic examination of the samples revealed that,
within the resolution of the optical microscope,
most of the smaller specimens had uniform extinction and a lack of twin lamellae, whereas the
larger samples were always twinned. The nature
of these twins is similar to that previously described by White et al. (1985), Yeganeh-Haeri et
al. (1989b) and Wang et al. (1990). Electron probe
microanalysis (EPMA) of the run product yielded
a composition of (Mg0988Si0998O3012).
The crystals were examined using X-ray precession photography and were proven to be
MgSiO3 perovskite. The smaller specimens also
exhibited well-developed (110) and (001) type
growth faces, a habit that is ubiquitous in perovskites with orthorhombic symmetry. X-ray photography methods were initially used to select
only untwinned crystals for Brillouin spectroscopy
measurements. The selected crystals were further
oriented and examined with an automated fourcircle X-ray diffractometer and confirmed to be
twin free. The resulting lattice parameters, obtained from the centered positions of 24 high
angle reflections, are a
4.776(1)A, b
4.928(1)A and c 6.894(2)A. These unit cell
edges are in good agreement with the previous
determination of Ito and Matsui (1978). All together, four single crystals, two mounted along
the c-axis and two along the [110] crystallographic
direction, were prepared for acoustic velocity
measurements. The maximum dimensions of these
crystals did not exceed 40 j.tm.
In our previous elasticity study we encountered a great deal of difficulty in obtaining highquality Brillouin signals on the perovskite specimens synthesized by Ito and Weidner (1986) using either the 488.0 or 514.5 nm wavelengths of
the argon ion laser excitation source (YeganehHaeri et al., 1989a,b). These single crystals exhib=

A. Yeganeh-Haeri /Physics of the Earth and Planetary Interiors 87(1994)111121

1600

1200

800

0
495

545

595

845

695

Wavelength (nm)

113

source. The radiation damage was not severe


under this configuration. However, this modification to the experimental configuration inevitably
resulted in a dramatic decrease in the intensity of
the Brillouin signal because the intensity of Brilbum scattered light is proportional to the inverse
of the incident wavelength to the fourth power,
that is, B ~ 1/A4. Therefore, data were collected
over a limited region in crystallographic space.
Subsequently,
we were able to excite only nine
longitudinal waves.
The severe laser-induced fluorescence encoun-

Fig. 1. Emission spectra of a single crystal of MgSiO


3 perovskite synthesized by Ito and Weidner (1986). The excitation
source was provided by the 488.0 nm line of an Ar ion laser
operating at 10 mW output power. Laser spot size was enlarged to about 25 ~sm.

ited an intense fluorescence and would subsequently disintegrate into polycrystalline aggregates. Fig. 1 shows is the photoemission spectra
obtained from one of the single crystals grown by
Ito and Weidner (1986). Note the presence of the
large fluorescence peak in the blue-green region
of the visible spectrum. Obtaining a robust Brilbum signal under such efficient laser-induced
fluorescing conditions is indeed a difficult task.
As the mirrors installed in the FabryPerot interferometer are particularly coated for maximum
transmittance in the 488.0514.5 nm region, such
intense and broad fluorescence cannot be readily
filtered. This fluorescence signal easily entered
into the photodetection device and overwhelmed
the Brilbouin peak. More importantly, after only
several hours of exposure to the laser beam, with
laser powers as low as 5 mW, once transparent
single crystals would become frosty looking. X-ray
examination of these samples revealed the onset
of local twinning. Additional illumination by the
laser beam invariably resulted in the migration of
twin domains throughout the crystals, transforming single crystals into polycrystalline aggregates.
It is also clear from Fig. 1 that these crystals were
not as fluorescent in the red region of the visible
spectrum (e.g. 650.0700.0 nm). Therefore, in our
previous study the spectrometer was modified to
operate in this region of the visible spectrum and
a krypton ion laser was used as the excitation

tered in the crystals grown by Ito and Weidner


(1986) can probably be attributed to the presence
of transition metal impurities, such as V203,
Cr203 and FeO in these samples. For instance, it
is well known that radiation produces electronhole pairs in alkali halides. The hole becomes
self-trapped and recombination with an electron
produces the so-called self-trapped exciton (STE)
luminescence. Work is currently under progress
to elucidate better the nature of this luminescence in silicate perovskite.
When the new crystals were mounted on the
Brillouin spectrometer and excited with the 514.5
nm line of an Ar + laser no fluorescence was
detected and robust Brilbouin peaks appeared.
Unlike the previous perovskite specimens, these
single crystals were found to be stable under the
focused laser beam. The quality of the spectra
was also quite superior to that obtained in our
earlier investigation.
2.2. Acoustic velocities
Brilbouin spectra were gathered in 53 unique
crystallographic orientations on silicate perovskite. More than 110 separate spectra were
collected, each having either different polarization or different combinations of growth surfaces
serving as incident and scattered faces. For many
crystallographic directions several spectra were
collected; these redundant data were averaged
together. All data were explicitly corrected for
the effect of mismatch between the samples refractive index and that of the fluid (~5i45nm
1.631) as outlined by Vaughan and Bass (1983).
The robustness and precision of the Brillouin
=

114

A. Yeganeh-Haeri/Physics of the Earth and Planetary Interiors 87(1994) 111121

Table 1
Acoustic velocities for MgSiO

3 perovskite

Propagation direction
n2

Polarization direction
n3

1.00
0.04

0.04
1.00

0.01
0.04

0.56

0.83

0.03

0.27

0.96

0.01

0.98

0.04

0.22

0.90
0.04

0.04
0.97

0.44
0.25

0.04

0.85

0.53

1.00
0.04
0.00
0.04
0.95
0.79
0.04
0.04
0.01
0.01

0.04
1.00
0.02
1.00
0.04
0.03
0.94
0.94
0.64
1.00

0.01
0.02
1.00
0.03
0.30
0.61
0.33
0.33
0.77
0.00

0.33
0.60
1.00

0.95
0.80
0.02

0.01
0.03
0.02

0.94
0.78
1.00
0.95
0.81
0.01
0.00

0.34
0.63
0.01
0.01
0.00
0.94
0.79

0.01
0.00
0.01
0.30
0.58
0.33
0.61

0.00
0.01
0.86
1.00

0.79
1.00
0.00
0.03

0.61
0.02
0.51
0.01

0.97

0.04

0.26

0.88
0.75
0.13

0.05
0.05
0.10

0.48
0.66
0.99

0.03

0.00

1.00

0.71
0.51

0.70
0.86

0.06
0.02

Velocity (km s)

u1

u2

u3

Obs.

Cal.

0.01
0.03
0.00
0.49
0.00
0.21
0.00
0.97
0.24
0.47
0.03
0.01
0.03
0.03
1.00
0.03
7.02
0.00
0.32
0.33
0.01
0.03
0.05
0.01
0.00
0.00
0.00
1.00
0.02
0.01
0.00
0.01
0.32
0.60
0.00
0.00
0.00
0.00
0.01
0.54
1.00
0.01
0.96
0.28
0.50
0.63
0.14
0.09
0.03
0.00
0.74
0.90

0.00
1.00
0.04
0.87
0.04
0.98
0.01
0.03
0.01
0.04
0.97
0.26
0.85
0.52
0.03
1.00
7.05
0.03
0.01
0.83
0.34
0.94
0.74
1.00
0.00
0.01
0.04
0.02
0.00
0.00
0.00
0.00
0.00
0.01
0.34
0.80
0.59
0.59
1.00
0.00
0.03
0.00
0.04
0.01
0.06
0.30
0.13
0.99
0.00
1.00
0.67
0.43

1.00
0.04
1.00
0.03
1.00
0.01
1.00
0.24
0.97
0.88
0.26
0.97
0.52
0.85
0.01
0.02

6.73
11.45
7.09
11.11
7.08
11.39
7.10
10.82
6.62
6.46
11.58
6.98
11.51
6.76
10.79
11.39

6.73
11.43
7.04
10.94
6.95
11.30
7.02
10.87
6.67
6.52
11.45
6.95
11.47
6.73
10.83
11.43

1.00
0.95
0.45
0.94
0.33
0.67
0.00
1.00
1.00
1.00
0.02
1.00
1.00
1.00
1.00
0.95
0.80
0.94
0.59
0.80
0.80
0.02
0.84
0.01
1.00
0.28
0.96
0.86
0.72
0.98
0.11
1.00
0.00
0.03
0.01

6.95
6.52
6.38
6.89
11.48
6.66
11.27
7.02
6.96
6.88
10.87
6.74
6.75
6.84
6.75
6.52
6.44
6.82
11.51
6.56
6.59
11.35
6.44
10.90
6.82
11.02
6.67
6.46
6.36
11.07
6.99
10.86
7.10
6.60
6.42

7.04
6.62
6.40
6.90
11.46
6.65
11.43
7.05
7.01
6.93
10.83
6.73
6.77
6.86
6.73
6.62
6.43
6.89
11.45
6.67
6.67
11.43
6.47
10.83
6.73
10.88
6.65
6.49
6.40
10.90
7.01
10.87
7.05
6.66
6.47

A. Yeganeh-Haeri /Physics of the Earth and Planetary Interiors 87 (1994) 111121

115

Table 1 (continued)
Propagation direction
n

Polarization direction
n3

0.00

1.00

0.01

0.02

1.00

0.02

0.71
0.49

0.70
0.49

0.01
0.72

0.52

0.49

0.70

0.22

0.25

0.94

0.03

0.05

1.00

0.02

0.00

1.00

0.50
1.00
0.51
0.52
0.71
0.90
1.00

0.53
0.01
0.52
0.52
0.70
0.44
0.03

0.69
0.03
0.69
0.68
0.01
0.02
0.01

Velocity (km s I)

u1

u2

u3

Obs.

Cal.

0.00
1.00
0.01
1.00
0.74
0.46
0.61
0.49
0.61
0.23
0.33
0.04
0.01
0.02
0.00
0.62
0.01
0.61
0.62
0.74
0.42
0.03

1.00
0.00
1.00
0.01
0.68
0.52
0.78
0.51
0.78
0.29
0.92
0.07
1.00
0.00
1.00
0.76
1.00
0.77
0.77
0.68
0.91
1.00

0.02
0.00
0.02
0.00
0.00
0.72
0.17
0.71
0.14
0.93
0.21
1.00
0.07
1.00
0.00
0.19
0.00
0.17
0.16
0.01
0.01
0.00

11.39
6.03
11.47
5.99
6.61
10.85
6.75
10.97
6.80
11.00
6.97
10.94
7.03
10.97
7.12
6.78
5.95
6.85
6.73
6.56
6.49
6.08

11.43
5.99
11.43
5.99
6.66
11.13
6.80
11.12
6.79
10.99
6.93
10.87
7.04
10.87
7.05
6.78
5.99
6.78
6.78
6.66
6.41
5.99

Polarizations of the shear waves are known; number of data 73; rms error 0.085 km s~.
Refractive index: n~= 1.760, n~= 1.770, n7 = 1.785.

spectra were deduced from the internal consistency and reproducibility of the results from specimen to specimen with different crystallographic
orientations, and by the redundant determinations of acoustic velocities. The final set of acoustic velocities consisting of 73 independent cornpressional and shear modes is presented in
Table 1.

bocities is 0.085 km s~ (see also Table 1). It is


important to point out that our newly determined
single-crystal elastic coefficients for MgSiO3 perovskite are, far superior to our previous measurements, and as such, the earlier results should no
longer be used.
3. Discussion and conclusions

2.3. Single crystal elastic moduli


3.1. Comparison with previous studies
The inversion method developed by Weidner
and Carleton (1977) was used to calculate singlecrystal elastic moduli from the measured veboci3
ties. adopted
In the inversion,
a density
of 4.108The
g cm
was
for the silicate
perovskite.
nine
adiabatic single-crystal elastic coefficients of
MgSiO
3 perovskite along with the estimated
isotropic aggregate properties are given in Table
2. The velocity surface of MgSiO3 perovskite is
illustrated in Fig. 2. The root mean square deviation of the best-fit model from the observed ye-

Pressurevolume data for silicate perovskite


have been
by either
severaladifferent
tors,
each reported
employing
differentinvestigatype of
sample or experimental technique. A comparison
of all the literature values for the bulk modulus
and linear compressibilities of (Mg,Fe)Si03 perovskite is given in Table 3. Mao et al. (1991)
discuss in some detail the differences in bulk
modulus and linear compressibibities among the
various static compression experiments. For ex-

116

A. Yeganeh-Haeri /Physics of the Earth and Planetary Interiors 87 (1994) 111121

Table2
Single-crystal elastic coefficients for MgSiO

14

3 perovskite deter-

mined at iO~GPa and 22CC


~

11
22
33
44
55
66
12
13
23

482(4)
537(3)
485(5)
204(2)
186(2)
147(3)
144(6)
147(6)
146(7)

0.00240
0.00213
0.00239
0.00490
0.00538
0.00678
0.00048
0.00058
0.00049

10

4
I

___________

Voigt

Reuss

Hill

264.2
178.5
11.06
6.59
8.03

263.7
176.2
11.02
6.55
8.01

264.0(5.0)
177.3(3.5)
11.04
6.57
8.02

Isotropic aggregate properties

V~
V,1.

I~-u

__________________________________________

K
is

12

Crystal Orientation
Fig. 2. The velocity surface of MgSiO3 perovskite projected
on to ab, bc, and ac planes. The solid points are
experimentally determined velocities. The solid line is the best
fit model. The rms deviation of the model from the measured
velocities is 0.085 km s~.
.

C.~(GPa); S~,(GPa 1); velocities (km s i)~


= 0.18,
= a
Linear compressibilities: j3~3~Ap
= 1.33, = 0.07
p6 = As
1.14,
~ =A1.31
(TPa~)~
density
g cmexpressed in terms of acoustic
measure of
elastic= 4.108
anisotropy
velocities: Ap =(~max 1,min)/ V, aggregate; A~=(~cm,,,,
~min)/
v, aggregate.

well
to the lower
pressures
attained
the
staticascompression
studies
of Ross
and in
Hazen
(1990) and Kudoh et al. (1989). We also note that
the primary purpose of conducting a static compression experiment is to measure the pressure
derivative of the bulk modulus, yet in almost all
cases a value of 4 has been assumed for K
(Table 3). Taken at face value and notwithstanding the differences noted above, and neglecting
the difference between adiabatic and isothermal
conditions, there is now a fair amount of agree-

ample, the disparity may be attributed to the


occurrence of non-hydrostatic pressure in the
measurements of Yagi et al. (1982) and Knittle
and Jeanboz (1987), and to uncertainties in volume measurement, pressure determination, as

Table 3
Bulk modulus and linear compressibility data for silicate perovskite
K

j3~

P
6

p~

Method

Sample

Pre.media

Max.press.

258(20)
266 (6)
247(14)
247 (5)
254 (5)
273 (4)
261 (4)
264 (5)

1.58

1.19

1.10

1.41
1.31
1.31
1.29
1.34
1.33

1.07
1.20
1.05
1.03
1.08
1.14

1.57
1.56
1.24
1.31
1.43
1.31

4(assuming)
3.9(0.4)
4(assuming)

DAC
DAC
DAC
B.S.
DAC
DAC
DAC
B.S.

PolyX
PolyX
SingX
SingX
SingX
PolyX
PolyX
SingX

me
None
mew

7
111
10

4(assuming)
4(assuming)
4(assuming)

me
Ne
Ne

12
30
30

1
2
3
4
5
6
7
8

K (GPa); linear compressibilities (TPa i); DAC, diamond anvil cell; B.S., Brillouin scattering; PolyX, polycrystalline sample;
SingX, single-crystal sample; Pre.media, pressure medium; mew, methanolethanolwater; me, methanolethanol; Ne, neon;
max.press., max.pressure attained in the experiment (GPa).
(1) Yagi et al. (1982); (2) Knittle and Jeanloz (1987); (3) Kudoh et al. (1989); (4) Yeganeh-Haeri et al. (1989a,b); (5) Ross and Hazen
(1990); (6) and (7) Mao et al. (1989, 1991) data reported here was obtained from a least-squares fit of three different samples of
(Fe~,Mg1_~)SiO3
where x = 0.0, 0.1 and 0.2; (8) this study.

A. Yeganeh-Haeri /Physics of the Earth and Planetary Interiors 87 (1994) 111121

ment concerning the bulk modulus of silicate


perovskite.

16.00

3.2. Crystal structure and elasticity of MgSiO3


On the basis of single-crystal structural refinements, Horiuchi et al. (1987) suggested that the
structure of MgSiO3 perovskite is essentially
composed of chains of rigid and relatively regular
Si06 octahedra that extend in three dimensions.
Such a topology inevitably leads to an elastically
stiff structure because it is hard to deform chains
load-bearing framework. The large
td of
the elastic stiffness coefficients of MgSiO3, especially moduli C11, C22 and C33, indeed demonstrate the strength of this material (Table 2).
The single-crystal elastic moduli of MgSiO3
perovskite also display a pattern that is elastically
somewhat anisotropic. The maximum shear and
compressional velocities are 18% and 7% greater
than the minimum (Table 2). On a comparative
basis, the differences between the maximum and
minimum shear and compressional velocities in
olivine, which is considered to be extremely
anisotropic, are about 25% and 24%, respectively. The [010] crystallographic direction in perovskite contains both the fastest and the slowest
shear wave velocity (Table 2). The fastest wave is
polarized parallel to the [001] direction (i.e. C~)
and the slowest is polarized along the [100] crystalbographic direction (i.e. C66). If under lower
mantle conditions magnesium silicate perovskite
grains were to become preferentially oriented, as
opposed to being randomly oriented, a shear wave
propagating in the Earths lower mantle could
become polarized with two distinct velocities,
It is useful to look at the effects of pressure
and temperature on the degree of structural distortion by examining the rotation angles ~ and w.
The linear compressibility data (Table 2) allow us
to calculate directly the pressure dependence of
~ and w, assuming that the Si06 octahedra behave as rigid and rather regular units. As seen
from Fig. 3(a). increases slightly, while w does
not change significantly with increasing pressure.
This is consistent with earlier observations of
Knittle and Jeanboz (1987). For instance, at ambi-

117

14.00

12.00

__._._..._._._._._._._.

~
~
.~

10.00

20

25

30

(GPa)

13.480

12.570

11.680

___________________________________

10.750290

389

488

587

686

785

864

983

1082

Temperature (K)
Fig. 3. (a) Variation of octahedral rotation angles ~ (solid
line) and w (dashed-dotted line) as a function of pressure. (b)
Variation of the same angles (open circles = ~ closed circles
w) with increasing temperature. Data are from Wang et al.
(1993).

ent conditions the and w octahedral rotation


angles for MgSiO3 are 14.35 and 11.67, whereas
at 30 GPa they are 15.6 and 11.89, respectively.
Interestingly, analysis of the high-temperature
unit cell data of Wang et al. (1993) indicates that
both ~ and w decrease slightly with increasing
temperature (Fig. 3b). It appears that the net
effect of these two offsetting and competing
mechanisms is such that the overall distortion of
the perovskite structure, at least up to 3036 GPa
and 12001500 K, is comparable in magnitude
with that observed at 1 bar and 300 K.
3.3. Elasticity and composition of the lower mantle
Compositional models deemed appropriate for
the Earths deep mantle range from pyrolite
(silica-poor) to pyroxene (silica-rich) stoichiometries. Under lower mantle conditions, this compo-

118

A. Yeganeh-Haeri /Physics of the Earth and Planetary Interiors 87 (1994) 111121

sitional variation corresponds to petrobogical assemblages of


7080% (by volume) perovskite
with 3020% magnesiowustite or to pure per
ovskite Tests of these compositional models can
be made by comparing the elastic properties im
plied by a given assemblage with those inferred
from seismology
The current data base for thermoelastic properties of silicate perovskite and magnesiowstite
is presented in Table 4. The following analysis
rests on the variation of density and seismic parameter over a depth range of 10002700 km
This depth range was selected because both at
shallower depths in the vicinity of the 670 km
discontinuity and in the deeper regions of the
Earth s mantle near the D layer, the density
and velocity structures are quite complex Corn
puted density and velocity trajectories for a pure
perovskite, (Mg089,Fe011)Si03, modeled as a
~

function of depth are shown in Figs. 4(a) and (b).


All calculations were performed along a 2000 K
adiabat initiated at zero pressure (Anderson,
1989). It is clear from these figures that a pure
perovskite model can provide a satisfactory fit to
the density and velocities observed seismically
throughout the entire lower mantle provided that
the coefficient of thermal expansion and temperature derivative of the isothermal bulk modulus,
of silicate perovskite are1, respectively.
3.61 x iO~ This
K and
is
5.8
x
10~
GPa
K
consistent with earlier suggestions of Jackson
(1983), Bukowinski and Wolf (1990), and with the
Table 4
..
.
Thermoelastic properties of silicate perovskite and magnesiowstite used in this study
3Property
(Mg, Fe)Si0
3
(Mg, Fe)O
K
(iKT / ~P)
(aK/aT)~x 10-2
(aK/aFe)
is

264
3.9
(2.3, 8.0)
O.OXFe
177.3

(~is/~) x10-2

(a,s/aFe)
a x i0~

162
4.0
9.OXFe
2.7
~
132
24

(1.65, 4.0)

~79.0XF
3.8

Numbers in parentheses indicate the range of a given parameter as reported by several different investigators. K and ~
(GPa); temperature derivatives (GPa K 1); a (K 1)

5280
(A)
.~

5020
4760

4.500

120.0
~
~
,~

109 0
~

(B)

76 0

13~

16

1900

2200

2500

Depth (kin)
Fig. 4. Comparison of the observed density (a) and seismic
parameter (b) vs. the calculated profiles for a pure (Mg089,
Feoii)Si03 perovskite model. The filled circles are from the
PREM model. All necessary elasticity data are given in Table
4. The shaded area represents the uncertainty in the mineralogical model. Dashed-dotted line as explained in text.

more
analysis
of Stixrude
al. (1992), and
impliesrecent
that the
Earths
mantle isetcompositionally
stratified (e.g. the lower mantle is silica-enriched
relative to the upper mantle). It is useful to note
that density perturbations drive mantle convection. And because density perturbations are essentially susceptible to changes in Mg/Fe ratio
.

rather than to changes in silicon content, a lower


mantle composed of ferromagnesium perovskite
with an Fe/(Mg + Fe) ratio of 0.11 would, therefore, not impose a complete blockade on whole
It is convection.
important to realize that to date there
mantle
have only been three separate measurements of
thermoelastic properties of silicate perovskite
conducted at simultaneous high pressures and
temperatures (Mao et al., 1991; Funamori and
Yagi 1993 Wang et al., 1993). The experiments
of Mao et al. (1991) yielded a rather large coefficient of thermal expansion, that is, a ~ 3.5(0.5) X

A. Yeganeh-Haeri /Physics of the Earth and Planetary Interiors 87 (1994) 111121

10 ~ K1 and a very large negative temperature


derivative of the isothermal bulk modulus,
(aK/8T)p ~ 63(05) x 102 GPa K1 On the
other hand, the measurements of Wang et al
(1993) yielded significantly lower values of a
1 65(0 19) x iO~K1 and (8K/aT)p
23(11)
x 102 GPa K More recently Funamori and
Yagi (1993) measured thermal expansivity of silicate perovskite from 300 to 1900 K at a constant
pressure of 36 GPa. These authors determined
the thermal expansivity of MgSiO
3 perovskite to
be 1 7(02) x iO~ K~ The measurement of Fu
namori and Yagi (1993) is in reasonable accord
with those of Wang et al (1993), but somewhat at
odds with the results of Mao et al (1991) The
thermal expansivity reported by Funamori and
Yagi (1993) at 36 GPa is also in fair agreement
with the zero-pressure results of Parise et al.
(1990) and Ross and Hazen (1989), indicating
that the pressure dependence of a is rather small.
More measurements over a wider range of conditions are needed to confirm this observation.
A pivotal question, then is whether a pure
perovskite model can still be plausible given that
the thermal expansivity and the temperature
derivative of bulk modulus have been reduced by
factors of about 2 and 3, respectively, from the
values appropriate to Figs. 4(a) and (b). If the
coefficient of thermal expansion and the temperature derivative of bulk modulus of silicate perovskite are as low as those reported by Wang et
al. (1993) and Funamori and Yagi (1993) then a
pure perovskite model mantle would be too fast
by about several percent in t, to provide an
acceptable fit to the seismic data. This is shown
by the dash-dotted line in Fig. 4(b). On the other
hand, by increasing the magnesiowstite, or in
other words the olivine content of the system,
and by performing the calculations along a 1750
K adiabat, the quality of the fit improved substantially. A model mixture composed of 75% (by
volume) silicate perovskite and 25% magnesiowiistite, consistent with a pyrolitic-type composition, agrees reasonably well with the density and
seismic parameter throughout the lower mantle,
The results of this calculation are shown in Figs.
5(a) and (b). This observation is also in accord
with the earlier analyses of Jackson (1983) and

119

5280

5020

4780

4500

120.0
Lu

~Z

1090
(B)
980

870
78.0
~

is~o

1600

1900

2200

2500

Depth (kin)
Fig. 5. Comparison of the observed density (a) and seismic

parameter (b) vs. the calculated profiles for a pyrolite-type


.
..
.
model. The iron partition coefficient between perovskite and
magnesiowstite was set at 3.3. Symbols are the same as in
Fig. 4.

Bukowinski and Wolf (1990), but in sharp contrast to that of Stixrude et al. (1992).
Of course, it would be highly desirable to use
the variation of compressional and shear velocity
in addition to density and seismic parameter
throughout the lower mantle, to place tighter
constraints on the range of acceptable chemical
models. However, the absence of the pressure,
temperature and mixed derivatives of compressional and, especially, shear velocities for magnesium-, iron-, aluminum- and calcium-bearing silicate perovskites significantly hinders the robustness of such an approach. It is also worth re-emphasizing that the value of the shear modulus of
MgSiO3 perovskite measured at ambient conditions is comparable with the shear modulus of the
Earths lower mantle at a depth of 971 km (see
table II of the PREM model). Normally, acoustic
velocities tend to increase with increasing pressure and decrease with increasing temperature;
such a high ambient condition shear velocity can-

120

A. Yeganeh-Haeri/Physics of the Earth and Planetary Interiors 87(1994)111121

not be easily reconciled with the velocity structure of the lower mantle short of invoking some
anomalous behavior in a state variable, either by
requiring a large negative temperature derivative
or by requiring a large negative second pressure
derivative. The distinction between these various
scenarios must await further study. Moreover,
although the pressure derivative of the bulk modulus of silicate perovskite has been measured to
over 100 GPa, the uncertainty in this parameter is
still rather large and in need of further refinement. For example, perturbations of K 3.9
0.4 within its stated uncertainty can have a significant effect on the computed seismic parameter
and hence on the inferred chemical composition,
as shown by the shaded region in Fig. 4(b). In
addition, no measurements have yet been conducted on the acoustic properties of magnesiowstite at sufficiently high states of compression
(e.g. P> 50130 GPa) and at high temperatures
characteristic of the Earths deep mantle.
With all these uncertainties in mind, we will
stop short of a complete perturbation of composition (i.e. perovskite to magnesiowstite ratio),
temperature and thermoelastic properties of perovskite in the exhaustive search for a unique
lower mantle composition. In conclusion, the current seismic and mineral physics data cannot unequivocally constrain the chemical composition of
the Earths lower mantle. Not until these vanables have been accurately characterized can
questions concerning the exact constitution of the
Earths lower mantle and whether it is chemically
dissimilar from the upper mantle, be answered.
=

Acknowledgments
I thank two anonymous referees for their constructive and helpful reviews. I am indebted to
my advisor Don Weidner for many valuable discussions and comments, and for the use of the
Brillouin scattering facility at Stony Brook, without which this study would not have been possible. I also thank Y. Wang for experimental assistance in synthesis runs. I am grateful to Thomas
J. Ahrens for financial support that made possible the completion of the manuscript. The emis-

sion spectra of MgSiO3 perovskite were collected


in the laboratory of D. Schiferl at Los Alamos
National Laboratory, to whom I am thankful. The
crystal synthesis experiments were conducted at
Stony Brooks High-Pressure Laboratory which is
supported by the National Science Foundation
Center for High Pressure Research (CHiPR) and
the State University of New York at Stony Brook
under grant EAR89-20239. The Brilbouin scattering laboratory is supported by NSF grant EAR8804087. Mineral Physics Institute Contribution No.
100.

References
Anderson, DL., i989. Theory of the Earth. Blackwell Scientific, Boston.
Bukowinski, MS. and Wolf, G.H., 1990. Thermodynamically
consistent decompression: Implications for the lower mantie composition. J. Geophys. Res., 95: 12 58312 593.
Dziewonski, A. and Anderson, DL., 1981. Preliminary Reference Earth Model. Phys. Earth Planet. Inter., 15: 297356.
Fei, Y., Mao, H.-K., Shu, J. and Hu, J., 1992. PVT equation of state of magnesiowstite (Mg06,Fe04)O. Phys.
Chem. Mineral., 18: 416421.
Funamori,
N. and
T., 1993.
High pressure
and high
temperature
in Yagi,
situ X-ray
observation
of MgSiO3
perovskite under lower mantle conditions. Geophys. Res.
Lett., 20: 387390.
Gasparik, T., 1989. Transformation of enstatitediopside
jadeite pyroxenes to garnet. Contrib. Mineral. Petrol., 102:
389405.
Hemley, R. and Cohen, R.E., 1992. Silicate perovskite. Annu.
Rev. Earth Planet. Sci., 20: 553600.
Horiuchi, H., Ito, E. and Weidner, D.J., 1987. Perovskite-type
MgSiO3: single crystal x-ray diffraction study. Am. Mmeral., 72: 357360.
Ito, E. and Matsui, Y., 1978. Synthesis and crystal chemical
characterization of MgSiO3 perovskite. Earth Planet. Sci.
Lett., 38: 443450.
Ito, E. and Weidner, D.J., 1986. Crystal growth of MgSiO3
perovskite. Geophys. Res. Lett., ii: 464466.
Jackson, I., 1983. Some geophysical constraints on the chemical composition of the earths lower mantle. Earth Planet.
Sci. Lett., 62: 91103.
Knittle, E. and Jeanloz, R., 1987. Synthesis and equation of
state of (Mg, Fe)Si03 perovskite to over 100 gigapascals.
Science, 235: 668670.
Kudoh, Y., Ito, E. and Takeda, H., 1989. Effect of pressure on
the crystal structure of perovskite-type MgSiO3. Phys.
Chem. Mineral., 14: 350354.
Mao, H-K., Hemley, R., Shu, J., Chen, L.-C., Jephcoat, A.,
Wu, Y. and Bassett, W., 1989. The effect of pressure,
temperature and composition on the lattice parameters

A. Yeganeh-Haeri /Physics of the Earth and Planetary Interiors 87 (1994) 111121


and density of (Mg,Fe)Si0

3 perovskites to 30 GPa.
Carnegie Inst. Washington Year Book, 19881989, pp.
8289.
Mao, H.-K., Hemley, R., Shu, J., Chen, L.-C., Jephcoat, A.,
Wu, Y. and Bassett, W., 1991. The effect of pressure,
temperature and composition on the lattice parameters
and density of (Mg,Fe)5i03 perovskites to 30 GPa. J.
Geophys. Res., 96: 80968079.
Parise, J., Wang, Y., Yeganeh-Haeri, A., Fei, Y. and Cox, D.,
1990. Crystal structure and thermal expansion of (Mg,
Fe)Si03 perovskite. Geophys. Res. Lett., 17: 20892092.
Ross, N. and Hazen, R.M., 1989. Single crystal x-ray diffraction study of MgSiO3 perovskite from 77 to 400 K. Phys.
Chern. Mineral., 16: 415420.
Ross, N. and Hazen, R.M., 1990. High pressure crystal chernistry of MgSiO3 perovskite. Phys. Chem. Mineral., 17:
228237.
Stixrude, L., Hemley., R.J., Fei, Y. and Mao, H.-K., 1992.
Thermoelasticity of silicate perovskite and magnesiowOstite
and the stratification of the earths mantle. Science, 257:
10991101.
Vaughan, M.T. and Bass, J.D., 1983. Single-crystal properties
of protoenstatite: A comparison with orthoenstatite. Phys.
Chem. Mineral., 10: 6268.

121

Wang, Y., Guyot, F., Yeganeh-Haeri, A. and Liebermann,


R.C., 1990. Twinning in MgSiO3 perovskite. Science, 248:
468471.
Wang, Y., Weidner, D.J. and Liebermann, R.C., 1993. PVT
equation of state of (Mg,Fe)Si03 perovskite: Constraints
on composition of the lower mantle. Phys. Earth Planet.
Inter., submitted.
Weidner, D.J. and Carleton, H., 1977. Elasticity of Coesite. J.
Geophys. Res., 82: 13341346.
White, T.J., Segal, R.L., Barry, J.C. and Hutchinson, L., 1985.
Twin boundaries in perovskite. Ada. Cryst., B41: 9398.
Yagi, T., Mao, H-K. and Bell, P.M., i982. Hydrostatic cornpression of perovskite-type MgSiO3, In: S.K. Saxena (Editor), Advances in Physical Geochemistry, Vol. 2. Springer,
New York, pp. 317325.
Yeganeh-Haeri, A., Weidner, D.J. and Ito, E., 1989a. Elasticity of MgSiO3 in the perovskite structure. Science, 243:
787789.
Yeganeh-Haeri, A., Weidner, D.J. and Ito, E., i989b. Singlecrystal elastic moduli of magnesium metasilicate perovskite In: A. Navrotsky and D.J. Weidner (Editors), Perovskite: A Structure of Great Interest to Geophysics and
Material Science. AGU Monogr. No. 45, Am. Geophys.
Union, pp. 1326.

You might also like