Você está na página 1de 19

PHYSICS

OFTHE EARTH
AND P LAN ETA RY
INTERIORS

_________

ELSEVIER

Physics of the Earth and Planetary Interiors 85 (1994) 319337

Light elements in the Earths outer core: A critical review


Jean-Paul Poirier
Dpartement des Gomatriaux (URA CNRS 734), Institut de Physique du Globe de Paris, 4 Place Jussieu,
75252 Paris-cedex 05, France
Received 8 October 1993; revision accepted 8 February 1994

Abstract
There is little doubt that densities for the Earths outer core, inferred from seismology, require that it is
constituted of an alloy of liquid iron and light elements. However, the nature of the light alloying elements is still
uncertain as it depends in a large measure on the conditions of accretion of the Earth and formation of the core.
The arguments brought forward for or against silicon, oxygen, sulphur, hydrogen and carbon are critically reviewed.
There is no reason to consider that only one element is present in the outer core. Experimentally determined
and/or calculated ternary and quaternary phase diagrams are needed to provide constraints on the nature of the
light elements.
There is no reason to believe that the core is a particularly clean system D.J. Stevenson (1981)
All discussions of the nature of the light element suffer from too few data and too many extrapolations
(1984)

1. Introduction
More than 40 years ago, Birch (1952), from
seismic data, interpreted the outer core as liquid
iron, perhaps alloyed with a small fraction of
lighter elements. He suggested carbon and silicon. The fact that the outer core is mostly iron
was later established beyond reasonable doubt by
Birch (1961, 1964), who confirmed that the density of the core was about 10% lower than the
density of iron at the core pressures and temperatures, and that the seismic parameter (1 = K/p,
where K is the bulk modulus and p the density)
of the core was higher than that of iron. The
difference in density between the core and the
high-pressure phase of iron, on the basis of

R. Brett

shock-wave experiments, was later calculated to


be 10 2% (Jeanloz, 1979) and confinned by
static determination of the equation of state of
s-Fe up to 3 Mbar (Mao et al., 1990) (Fig. 1); the
uncertainty arises from the seismic density profile
and from the imperfect knowledge of the temperature in the core and the thermal expansion
coefficient of iron at core conditions. Birch (1964)
assigned the density deficit to lighter elements in
solution
silicon, sulphur or oxygen. Over the
years, a number of elements lighter than iron
silicon, sulphur, oxygen, hydrogen and carbon
were considered by various workers, singly or
more rarely in combination (Fig. 2).
The nature of the light elements in the outer
core is a standing problem of prime importance;

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


SSDI 0031-9201(94)02948-B

320

J.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

PREM (5000 K)

150

200

250

300

350

Pressure (CPa)

Fig. 1. Comparison of the density of solid pure a-Fe at high pressures and 5000 K with the core density (Preliminary Reference
Earth Model; PREM). The fusion volume of iron (about 3%) is not taken into account. After Badding et al. (1992).

it conditions in particular the existence and value


of a freezing point depression at the inner core
boundary (ICB), on which the answers to major
questions in geophysics and geochemistry hang:
e.g. to what extent are light elements released at
the ICB, thus inducing compositional convection
thought to power the geodynamo (Braginsky,
1964; Loper, 1978)? By how much is the temperature at the ICB lower than the melting temperature of pure iron at 3.3 Mbar (Stevenson, 1981)?

30

________________________________

20
Si

io

H
C

1950

1960

1970

1990

2000

Year

Fig. 2. Cumulative number of papers on light elements in the


core, as a function of publication date,

What is the composition of the inner core


(Jephcoat and Olson, 1987)? Are chalcophile elements depleted in the mantle by volatilization or
by being sequestered in the core? What is the
significance of the UPb age of the Earth (Oversby and Ringwood, 1971)?
It can be taken for granted that any light
solute element in suitable proportions in liquid
iron will decrease its density to make it compatible with seismological Earth models (Birch, 1952);
however, the number of potential alloying elements is restricted by the following cosmochemical and metallurgical constraints (Stevenson,
1981):
(1) The light elements must be sufficiently
abundant in the accreting Earth, thought to be of
chondritic composition. However, they must not
be volatilized by the heat of accretion and escape
entirely.
(2) It is currently accepted that most of the
core formed early during accretion (Oversby and
Ringwood, 1971; Allgre et al., 1982), hence before the Earth reached its present size. The light
elements should then be able to partition into
liquid iron at relatively low pressures (it may be
useful to remember that the central pressure of

J.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

the Moon is about 50 kbar and that of Mars


about 450 kbar) and preferably help form a
metallic liquid melt (e.g. eutectic) at temperatures lower than the melting point of pure iron.
(3) The outer core is presumably homogeneous, hence the light elements should remain
soluble in iron at the current pressures of the
core.
(4) At least some of the light elements must be
released in the melt during crystallization of the
inner core and the freezing point depressed, if
compositional convection is to be effective,
The nature of the light elements is tightly
linked to the mode of formation of the core and
it has been deplored that models of core formation are further complicated by our lack of knowledge of the chemical composition of the core
(Jones and Drake, 1986) or pointed out that the
nature of the most abundant light element in the
core is important for determining whether core
formation was dominated by high pressure or low
pressure processes (Newsom and Sims, 1991). I
take, of course, the opposite view and regret that
the present knowledge of core formation does not
provide more constraints on the nature of the
light elements.
The most valuable information on the possible
compositions of the outer core comes from phase
diagrams of binary, ternary and quaternary Fe
light elements systems at pressures up to the ICB
pressure. Some binary and ternary diagrams have
been experimentally determined to pressures of
about 100 kbar, and calculated to higher pressures when equations of state for the end-members are known. It may be noted here that about
4% nickel is thought to be present in the core
(Brett, 1971) and although it does not appreciably
change the density of liquid iron, its presence
should not be forgotten, as phase diagrams of
systems FeNilight elements may be significantly different from those of systems without
nickel (Urakawa et al., 1987).
More or less succinct reviews of the literature
on light elements in the outer core have been
written, as part of papers or books on the core
(Brett, 1976; Stevenson, 1981; Jacobs, 1987, 1992;
Jeanloz, 1990), or as an introduction to papers
supporting the presence of a given element (e.g.

321

Ringwood, 1977). In what follows, without any


preconceived idea nor theory of my own about
the nature of the light elements in the core, I will
systematically review the literature from Birch
(1952) to the present day, and try and sort out the
factual information and the speculation buttressed by ad-hoc models. In an admittedly
pedestrian way, I will successively examine the
case for or against silicon, sulphur, oxygen, hydrogen and carbon, or mixtures thereof, in an order
following more or less the evolution of fashion
during the last 40 years, as reflected by the cumulated number of papers on one or the other of
the light elements (Fig. 2). Although sometimes
listed among the possible light elements, nitrogen
was never the object of any attention. Alder (1966)
claimed that magnesium could be present in the
core, but this conclusion resulted from calculations of the solubility of MgO based on unavailable physical quantities. Ringwood and Hibberson (1991) later showed experimentally that MgO
is one of the least soluble oxides in molten iron,
even at high pressure.
Is the outer core in equilibrium or not with the
mantle? This problem is obviously related to the
composition and formation of the core. The controversy is centred on the measured abundances
of siderophile elements in the upper mantle, but
it is always possible to reproduce them in a somewhat ad-hoc fashion by accreting cocktails of different chondrites in various proportions at vanous times. I do not believe that the answer to this
problem may contribute much to selecting or
eliminating an element as the major light element
in the core. However, in view of the importance
of the topic in the literature, I will succinctly deal
with it.

2. Silicon
Birch (1952) first remarked that: any of the
most abundant elements will reduce the density
of iron. The effect of carbon and silicon are
perhaps the most familiar, a reduction of density
by 10% requiring only small percentages of these
elements. MacDonald and Knopoff (1958) then
pointed out that an Earth with an ironnickel

322

f-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

core and a penidotitic or eclogitic mantle would


have an (Fe + Mg + Ni)/Si ratio higher than that
of the chondritic meteorites, in contradiction with
the accepted tenet that the average composition
of the Earth is chondnitic; they concluded that
there must be silicon in the cone and that the
higher the silicon content assumed in the mantle,
the less silicon there must be in the core. MacDonald and Knopoff (1958) provided another independent argument for the presence of silicon

gases. The fact that some unreduced FeO is


present in the mantle is attributed to lack of
equilibrium. Support for this theory is found in
the mineralogy of meteorites, similar to blast
furnace assemblages (Ringwood, 1959), and in
the fact that there is 26 at.% Si in the metallic
phase of enstatite chondrites (Ringwood, 1961).
Ringwood (1959) also noticed, as had MacDonald
and Knopoff (1958), that there is a higher proportion of Si02 in the silicate phase of chondnites

(or indeed any light element) in the core: they


started from the observation that Bullens density
distribution and an interpolated Thomas
FermiDirac equation of state yield a weighted
mean atomic number Z = 22 for the outer core
(Knopoff and Uffen, 1954), obviously too low for
a core of iron (Z = 26) and nickel (Z = 28). The
value of Z for the outer core can be brought
down to 22 by the introduction in the core of
about 20 wt.% Si. They did not rule out the
presence of sulphur, although they saw difficulties in the fact that an FeS mixture with appropriate Z for the core would require an unrealistic
high initial abundance of sulphur or some process
by which silicon is lost to the Earth with respect
to the more volatile sulphur. Knopoff and MacDonald (1960), using equations of state determined from shock-wave data for iron and other
metals, later found that a material containing
between 20 and 30 wt.% Si is consistent with
seismic data.
Ringwood (1959) simultaneously proposed a
model of accretion of the Earth and formation of
the core resulting in the incorporation of silicon
in the core. He assumed that the Earth accreted
at low temperatures from oxidized dust and gas
of cosmic composition, trapping carbonaceous
compounds. As temperature rises during accretion, melting and convection occur and the
trapped carbon reduces the oxides, much as in a
blast furnace, according to the reactions
FeO+C~Fe+CO
FeO + CO
Fe + CO2

than in the Earths mantle and therefore, if the


compositions of Earth and chondrites are to be
similar, the silicon missing in the mantle must be
in the core; a satisfactory Earth model requires
the presence of about 20 wt.% Si in the core.
Urey (1960) took exception to the model proposed by Ringwood (1959), on the grounds that
reduction of both iron and silicon would entail
the evolution of very large quantities of CO and
that no satisfactory mechanism for the escape of
gases of molar weight 28 from a planet even the
size of Mars is known. He proposed that some
residual carbon and some sulphur dissolved in
iron would give a satisfactory explanation of the
density deficit of the core.
Balchan and Cowan (1966) performed shockwave experiments on FeSi alloys (4 and 19.8
wt.% Si) up to 2.7 Mbar and determined pressuredensity and sound speeddensity curves,
which they compared with the curves for the
Earth. They found that their results were consistent with a core containing 1420 wt.% Si, in
agreement with the suggestions of MacDonald
and Knopoff (1958) and Ringwood (1959).
From shock-wave theory, Stewart (1973) calculated all possible Hugoniots compatible with the
seismic properties of the outer core. He found
that his results were compatible with 820 wt.%
Si in the core, but did not rule out other light
elements.
Brett (1971), although agreeing that silicon is
likely to be a major element in the core, disagreed with
(1959,between
1961, 1966)
on and
the
question
of Ringwood
disequilibrium
core

SiO 2

mantle
andbymustered
arguments
for equilibrium,
answered
Ringwood
(1971). The
problem of

2C

Si + 2CO

A metallic phase containing iron, nickel and


silicon then segregates; CO and CO2 escape as

equilibrium vs. disequilibrium will be discussed in


Section 7. In a later review, Brett (1976) came to

J.-P. Poirier/Physics of the Earth ana Planetary Interiors 85 (1994) 319337

323

the conclusion that we totally lack solid evidence


on what the light element might be.
Ringwood (1977) acknowledged that his accretion model for incorporating silicon in the core
was very specific and met with difficulties,
pointed out by Brett (1971, 1976), essentially because it led to gross disequilibrium between mantie and core, and he thereupon suggested that
oxygen (as FeO dissolved in iron), and not silicon,

Wnkes model, raising doubts about the possibility of a nebula of solar composition ever achieving sufficiently reducing conditions for 1020% Si
in the core. Based on their study of the solubility
of mantle oxides in molten iron at high pressures,
and on Knittle and Jeanloz (1989) experiments in
a diamond-anvil cell, they suggested that the only
way to incorporate Si in the core would be by
dissolution of Si02 from mantle silicates at high

was the light element in the core. It must be


noted that Ringwood did not give very cogent
reasons (other than disequilibrium) for rather
abruptly abandoning the idea of silicon. Later on,
he eliminated altogether the need for silicon in
the core by suggesting that it is not the Earth
mantle that is depleted in silicon with respect to
the cosmic primitive composition, but the chondrites that are enriched (Ringwood, 1989).
In fact, the main objection to Ringwoods early
model for incorporating silicon in the core was
made by proponents of sulphur as the major light
element (see Section 3): reduction of silicon from
silicates requires a high temperature during accretion, which would have volatilized elements
more volatile than Si (e.g. 5) that, however, are
still present; it also produces vast quantities of
CO and CO2 that have to be blown off (Rama
Murthy and Hail, 1970, 1972). Wnke (1981) and
Wnke and Dreibus (1988) circumvented this difficulty by proposing an accretion in two stages:
first, a highly reduced, devolatilized material contaming Si in metallic form was accreted (thus
removing the problem of high-temperature reduction to the solar nebula or planetesimals
stage), then, when the Earth had reached about
two-thirds of its present mass, and after the metals and sulphides had segregated to form the
core, a more oxidized, volatile-rich material was
added (accounting for the chondritic proportion
of siderophiles in the upper mantle). Wnke could
therefore build an Earth with Cl chondrite abundances and a core containing 12.5% Si. He further suggested (Wnke, 1981) that the presence
of silicon in the core might result in the precipitation of the intermetallic compound Ni2Si to form
the inner core, as first proposed (without the
least justification) by Herndon (1979).
Ringwood and Hibberson (1991) objected to

pressure, but even then they saw difficulties owing to the fact that Ti02, which is more soluble
than SiO2, should be depleted in the mantle if
SiO2 was dissolved in the core; as it is not, they
concluded that there must be very little Si02 in
the core. It must be noted that the argument
against Si, based on Ti02, holds only if one
assumes that Si in the core comes from dissolution of the mantle at high pressure.
Allgre et al. (1994) calculated the ratio Si/Fe
in the core without assuming a priori the presence of silicon in the core (and without devising
an accretion model to acount for it). The ratio
(Si/Fe)core is calculated from the known ratio
(Si/Fe)mantie and from the ratio (Si/Fe)Earth,
which must be estimated:
(Si/Fe)
core

fm~(Si/Fe)mantie] /f~
where fm and ft are the mass fractions of iron in
the mantle and core, respectively.
The ratio (Si/Fe)Earth for the bulk Earth is
calculated from the Si/Al vs. Fe/Al trend established for meteorites (and assumed valid for the
Earth), and the value (Fe/Al)Earth is calculated
from the corresponding ratio for the mantle, taking into account the fact that aluminum does not
enter the core:
[(S1/Fe)Earth

(Al/Fe) Earth = fm~ (Al/Fe) mantle


Allgre et al. found about 11 wt.% Si in the
core. This would give a density deficit in the core
of only 67%. Other elements must therefore be
present (a few per cent S and 0 would give the
correct density).
From the metallurgical viewpoint, there is no
difficulty in having silicon dissolved in iron during
accretion at low pressures: the FeSi phase diagram at ambient pressure exhibits a eutectic point

324

J.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

wt % Si
10
20

35

30

____

1500

2S~j~1~

1400

C~

*20

1300
~

195

1212 1203

15

~;::~
~

______________________________

900

700

23 5

1200

800T

;)

/~f~~$W4.
f\

141

1500

825

t~,.

160d
F.

-.5

,~

~
.~

.~

.~

.~

wt%Si

Fig. 4. FeSSi miscibility gap at 1 atm (after Raghavan,


1988).

600

500
Fe

io

I
20

iT
30 40

50

At % Si
Fig. 3. FeSi phase diagram at 1 atm (after Raghavan, 1988).

at 1200Cand 20.5 wt.% Si (Fig. 3), and there is


no reason to believe that Si would be less soluble
at high pressure. Silicon also lowers the melting
point of iron at atmospheric pressure. However,
the ternary diagram FeSiS at 1 atm exhibits a
vast miscibility gap in the liquid state, which
widens as temperature increases (Raghavan,
1988); this, of course, would limit the solubility of
sulphur in FeSi (Fig. 4). The gap possibly shrinks
at high pressure, but we do not know that it does

calculations were based on the ambient pressure


densities of ironnickel and troilite!); sulphur
would then obviate the necessity of having Si in
the core, with the apparently unpleasant (why?)
consequence of coremantle disequiiibrium.
It was, however, Rama Murthy and Hall (1970,
1972) who really started an interest in sulphur.
They noticed that sulphur was depleted in the
crust and mantle relative to the other volatile
elements by several orders of magnitude, whereas
halogens, water and rare gases were present in
about their chondritic abundances (Fig. 5). The
depletion could not be due to preterrestrial frac-

Mason (1966) was the first to suggest definitely


that sulphur might be the major element in the
core: assuming that the Earth had a composition
of enstatite chondrite, he calculated the proportion of troilite FeS in the core and found that it
would give the correct density deficit (but his

~
~

/~

10

io-3~-

,,,

H~ ~F

ci

B,

____________

Fig. 5. Abundance of light elements in the Earth, referred to

chondritic abundance (after Rama Murthy and Hall, 1970).

J.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337


soc

of the FeNiS system from 30 to 100 kbar and,


using equations of state for Fe and for nonstoichiometric Fe09S (King and Ahrens, 1973), he

o
1400

I.

1188.
i

915

800

412
Fe~-

,~

Boo

~
400

36.5

200

Fe

10

20

38.8
I

30

40

325

50

wt % s
Fig. 6. FeS phase diagram at 1 atm (after Raghavan, 1988).

tionation, as meteorites exhibit no anomaly, nor


could it be due to volatilization during accretion,
as the other29Xe
volatiles
would by
have
lost
(in
produced
thebeen
decay
of too
shortparticular,
i
lived 1291) They thought it unlikely that sulphur
could be hidden in the lower mantle, as any
process that would have removed metallic Fe into
the core would also have removed sulphur. The
only remaining possibility was that sulphur was
sequestered in the core during core formation,
thanks to the existence of a low melting point
FeS eutectic (Brett and Bell, 1969) (Fig. 6).
They found that a mixture of 40% Cl chondrites,
50% ordinary chondrites and 10% iron meteorites provides a satisfactory composition for the
Earth, with a core containing 15 wt.% S (but a
mantle richer in FeO than pyrolite).
The idea that sulphur might segregate with
iron in the proto-core was quickly accepted (e.g.
Lewis, 1971) and a great number of experimental
data on the melting of the FeFeS system under
pressure, as well as on the equation of state of
sulphides, was obtained in the following years.
Usselman (1975) investigated the pressure dependence of the liquidus in the iron-rich portion

extrapolated the densities of the compositions on


the liquidus to the pressure of the ICB; comparing these values with the seismically determined
density, he found an average composition of the
outer core between 8 and 11 wt.% S.
Urakawa et al. (1987) studied the solubility of
sulphur and oxygen in the system FeSO up to
150 kbar. They found that the immiscibility gap in
the liquid region narrows with increasing pressure, and suggested that it might disappear above
250 kbar. Addition of nickel reduces the immiscibility of the liquids. Urakawa et al. (1987) also
found that alloying with S and 0 reduces the
interfacial tension between the metallic melt and
silicates and oxides: at high pressure the metallic
melt wets the grain boundaries and forms a network of liquid. Goarant et al. (1992) also found
complete wetting of grain boundaries of magnesiowstite and perovskite by FeOS melt at
pressures between 700 and 1300 kbar.
The melting temperatures of sulphides and
FelO wt.% S mixtures were measured in diamond
anvil cells
by Williams
and Jeanloz
up to about
1 Mbar
and by Boehler
(1992)(1990)
up to
about 500 kbar. Although the temperatures found
by Williams and Jeanloz are at least 500Chigher
than those of Boehler, there is agreement on the
fact that the eutectic behaviour in the FeS systern persists up to very high pressures and that
the Fe10% S alloy melts several hundreds of
degrees below the melting point of pure iron or
of iron sulphide.
Using solid-state and liquid-state physics models and results of shock-wave and melting experiments, Svendsen et al. (1989) performed somewhat involved calculations of the liquidus at high
pressure in the FeFeS system, assuming ideality
of the liquid and complete immiscibility of Fe and
FeS in the solid state. They found that the cornposition of a S-rich core is more likely to lie on
the Fe-rich side of the eutectic (which implies a
welcome depression of the freezing point of pure
iron). However, in a follow-up paper the same
workers (Anderson et al., 1989) concluded that
solid solution between S and Fe is possible at

326

f-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

high pressures and that the FeS system exhibits


continuous solid and liquid solutions (still with a
freezing point depression, as the melting point of
FeS is lower than that of Fe). The possibility of
high-pressure solid solution between S and Fe
was confirmed by electronic structure calculations
by Boness and Brown (1990) and Sherman (1991).
Shock-wave equations of state were determined for pyrrhotite FE7S8 by Ahrens (1979) and
Brown et al. (1984) and for pyrite FeS2 by Ahrens
and Jeanloz (1987); they were used, together with
equations of state for iron, to estimate the mixing
ratio of sulphur and iron matching densities in
the liquid outer core. The inferred sulphur contents
the only light
element
in
the(assuming
core) are sulphur
in goodis agreement:
912
wt.%
(Ahrens, 1979), 10 4 wt.% (Brown et al., 1984)
and 11 2 wt.% (Ahrens and Jeanloz, 1987).
Brown and McQueen (1982), using the pyrrhotite
data from Ahrens (1979) and their own shockwave equation of state for c-Fe, found a smaller
percentage of sulphur
510 wt.%
essentially because they took into account the difference in volume between solid and liquid iron.

Although sulphur is a most convenient light


element to have in the core as it forms a eutectic
with iron at low pressures, is still soluble in iron
at high pressures and lowers its freezing point,
there has never been a very satisfactory answer to
an embarrassing problem (e.g. Ringwood, 1977):
if sulphur, a very volatile element, were hidden in
the core in sufficient quantity to account for the
density deficit, it would be less depleted in the
bulk Earth than several less volatile, nonsiderophile elements, such as Na, K or Cl.
4. Oxygen
Some 20 years ago, Dubrovskiy and Pankov
(1972), as an alternative to the theory according
to which the core was not iron but a high-pressure metallic phase of the mantle silicate
(Ramsey, 1949), suggested that the iron outer
core of the Earth might contain more than 50%
in mass of iron oxide FeO, metallized under
pressure, that would have gravitationally segregated from partially melted lower mantle; the

~
20000;

05

1800L /
LI

%
22

o
23

24

L
11

1600

1523

1371

~II\
50.92

51.26

1200

wustite

1000

FeO

800

60

560

_____________

400
200Fe 1

51.42
2 3

50

51

52

53

Al % 0
Fig. 7. FeO phase diagram at 1 atm (after Raghavan, 1988).

density of the mixture, estimated from approximate equations of state, was consistent with that
of the core. Bullen (1973), with a similar rationale, proposed that the outer core consisted of
Fe
2O (12.5 wt.% 0).
However, oxygen as a major light element in
the core did not really become fashionable until
Ringwood (1977) revived the idea of Dubrovskiy
and Pankov and argued that, instead of silicon
that he had favoured earlier, or sulphur (the
current contender at the time), FeO ought to be
considered. The phase diagram of the FeFeO
system at atmospheric pressure exhibits a large
liquid miscibility gap and the solubility of FeO in
molten iron is very small near the liquidus (Fig.
7). However, the solubility increases rapidly with
temperature, and Ringwood, following Dubrovskiy and Pankov (1972), suggested that FeO becomes metallic at high pressure, and that consequently the miscibility gap between ionic and
metallic liquids should disappear at a pressure
estimated at about 300 kbar. Using a rough estimate of the density of metallic FeO at core
pressures, Ringwood concluded that, to fit the
seismic data, the core should contain about 44
wt.% FeO, equivalent to 10 wt.% 0.
Shock-wave experiments (Jeanloz and Ahrens,
1980) showed that Fe0 940 does indeed undergo a
phase transition, with density increase of 4% at

J.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

about 700 kbar, and that the Hugoniot data were


consistent with the outer core containing about
10 wt.% 0. However, Yagi et al. (1985) found no
large discontinuous density increase at 700 kbar
in FeO statically compressed at room temperature up to 1.2 Mbar. Knittle and Jeanloz (1986,
1991a) later showed that above 700 kbar the
electrical conductivity of FeO becomes almost
equal to that of iron (Fig. 8), and concluded that
the transition observed by Jeanloz and Ahrens
(1980) is due to the metallization of FeO and
that there is complete miscibility of the FeFeO
liquids above 700 kbar, in reasonable agreement
with the prediction of McCammon et al. (1983)
based on extrapolation of the FeFeO phase
diagram at high pressures.
Sherman (1989) remarked that the metallization of FeO is not necessarily related to a structural change owing to a change of the character
of the Fe0 bonding from ionic to metallic; it is
more likely that the sudden increase in electrical
conductivity is due to a Mott transition (insulatormetal transition) consisting in a delocalization of the Fe(3d) electrons over the metallic
sublattice, while the FeO bonding remains ionic;
this could explain why Yagi et al. (1985) did not

WUstite (DAC)

~ 10
WUstite (Shock-wave)

Fe

IC

~3~
~
Pressure (GPo)

Fig. 8. Electrical resistivity measurements on Fe0 ~O under


pressure. Above 700 kbar, the resistivity becomes comparable

with that of iron (after Knittle and Jeanloz, 1986).

327

2000

~ 1900
L.
~.

i~oo

\
\

Lm

Lm

1670

L,,+Li

/1760

Lm+ F.Oc

1875
~

1700
Fe

5. Fe05

______________________________
0

Fe

20

40

60

wt %

80

100

FeO

Fig. 9. Phase diagram for the system FeFeO at 160 kbar

(after Ringwood and Hibberson, 1990). Lm: metallic FeO


melt, Li: ionic FeOFe melt, Fec: crystalline iron, FeOc:
crystalline wiistite.

see anything at low temperature, where delocalization probably does not occur.
The solubility of FeO in liquid iron at high
pressures was investigated in a series of experiments at the Australian National University, first
in a piston cylinder apparatus up to 40 kbar
(Ohtani et al., 1984), then in a multi-anvil apparatus at 160 kbar (Kato and Ringwood, 1989; Ringwood and Hibberson, 1990, 1991); these experiments confirmed the increase of solubility of FeO
at high pressures (Fig. 9) and allowed an estimate
of the eutectic composition and temperature on
the Fe-rich side of the phase diagram (10 wt.%
0, 1670C), as well as of the depression of the
freezing point (about 28Cper 1 wt.% 0 at 160
kbar).
Knittle and Jeanloz (1991b) measured the vanation of the melting point of FeO up to 800 kbar
in a laser-heated diamond-anvil cell, and found
that the slope of the melting curve increases
above 700 kbar and that FeO melts at a higher
temperature than Fe. At 830 kbar, the alloy of
composition Fe069O031 is found to melt at a
temperature intermediate between the melting
point of Fe and that of FeO. Knittle and Jeanloz
concluded that at these high pressures, the eutectic disappears to be replaced by a two-phase

328

f-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

spindle between solid and liquid solution, and


that, consequently, oxygen, far from depressing
the freezing point of the alloy, would increase it.

4000

FeO
present in the core to counteract this effect, but
Another element (e.g. sulphur) should then be
the(1989)
overallremarked
depression
notthat
be Anderson
more
than eta
few
of degrees.
However,
al.
thatshould
the
fact
the melting
pointhundreds
of an FeFeO
alloy
is intermediate
between those of Fe and FeO does not necessarily
rule out the existence of a eutectic (Fig. 10), such
as the one they calculated. In any case, the composition of an FeFeO alloy consistent with outer
core density should fall on the FeO-nich side of
the phase diagram, leading to an enrichment of
the solid phase (inner core) in oxygen during
freezing.
On the basis of electronic structure calculations, the existence of a solid solution between Fe
and 0 at high pressures is also thought to be
unlikely (Boness and Brown, 1990; Sherman,
1991).
Boehler (1992, 1993) investigated melting of
FeO and FeFeO alloys at pressure up to 580
kbar, in a laser-heated diamond-anvil cell. Although he found values of the melting point of Fe
and FeO much lower than those of Knittle and
Jeanloz, he did find, like them, that FeO melts at
a higher temperature than Fe, and that there was
no significant difference between the melting
points at high pressure of iron and Fe8 wt.%
FeO and Fe30 wt.% FeO alloys (Fig. 11). There

~id_
Solid

Fe

________________FeO

________________
Fe
reQ

Fig. 10. Schema showing that a value intermediate between


the melting points of Fe and FeO for the melting point of an
alloy FeFeO (dot) can be compatible with a eutectic (right)
as well as with a solid solution (left) (after Anderson et al.,
1989).

(BoenIer,1992)\~~

~3000
2000

~Rin~00da~Hibirson,
//

7
1000

Fe:10~.%Fe0

Fe:8wt%FeO

+
0

1990

Fe: 3Owt. % FeC

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

P ( Mbar)
Fig. ii. Melting curves of Fe, FeO and FeFeO alloys (after
Boehler, 1993).

seems therefore to be no disagreement as to the


fact that, at core pressures, oxygen does not lower
the freezing point of iron.
Knittle and Jeanloz (1991b) and Goarant et al.
(1992) investigated the reaction between lowermantle material (silicate perovskite and magnesiowstite) and molten iron above 700 kbar. They
found that liquid iron infiltrated between the
grains of the solid phase and dissolved FeO from
it. Analytical transmission electron microscopy of
the samples after reaction in the laser-heated
diamond-anvil cell (Goarant et al., 1992) showed
depletion of the magnesiowstite in FeO and
complete wetting of the grain boundaries, in
agreement with the results at lower pressure of
Urakawa et al. (1987).
There is therefore little doubt that molten iron
of the core and oxides of the lower mantle react
at high pressure, thus enriching the core metal in
oxygen. Knittle and Jeanloz (1991b) even claimed
that it may be that the entire budget of light
alloying component in the outer core has come
from chemical reactions with the mantle. This,
however, seems unlikely, for the coremantle reaction is a very high pressure process and there
must surely have been some light elements introduced at low pressure during the early stages of
core formation.

f-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

329

5. Hydrogen
6000

Hydrogen, although a possible candidate for


lowering the density of the core, was long neglected for two reasons: it was thought that it
would escape during accretion (e.g. Jeanloz,
1990), and Ringwood (1966) had argued that, as
hydrogen enters into interstitial sites in iron, it
would not significantly decrease the density of

P
5000

L
4000
6

the core. As to the last point, Stevenson (1977)

mustered evidence to the contrary: interstitial


hydrogen expands the host metal; he also calculated that hydrogen might form an FeH hydride
at the pressure of only a few kilobars. About 1
wt.% of hydrogen would be enough to account
for the density deficit of the core and this quan
tity would be provided if only 10% of the accreting material was a low-temperature condensate
containing water, and if the water reacted with
Fe at a pressure such that H could enter in
solution and be retained. Stevenson concluded
that hydrogen was most probably one of the elements contributing to the density deficit of the
core (Stevenson, 1977), but later eliminated it on
the basis that it was too insoluble (Stevenson,
1981).
All the work on hydrogen in Fe was done
recently in Japan and at the Geophysical Laboratory of the Carnegie Institution of Washington.
Suzuki et al. (1984) studied the reaction between
enstatite, iron and hydrous minerals (simulating a
mixture of enstatite and Cl chondrites) at 50 kbar
and temperatures between 1000 and 1200C.
From the presence of olivine (containing FeO)
and iron in the reacted product, they indirectly
inferred the presence of hydrogen in the iron, on
the basis of the reaction
Fe

water

FeH + FeO
X

The metal had melted 500Cbelow the melting


point of pure iron.
Fukai and Suzuki (1986), using calculated or
estimated values of the atomic volumes of various
light elements (H,C,O,S) at high pressures, calculated their solubility in iron and the resulting
decrease in density. Fukai (1992) proposed a
high-pressure phase diagram for the FeH systern (Fig. 12) and re-examined the results of

100GPa

.~

L+LH2

3000

2000

C +

Li2

1000 I

~ 4SH2

/
-______________

Fe

_____________

Fe H

Fig. 12. Proposed phase diagram of the FeH system at 1


Mbar (after Fukai, 1992).

Fukai and Suzuki (1986), using two different


models of compressibility of the light elements to
fit the observed density deficit. They proposed
two core compositions, corresponding respectively to 10% and 15% of the low-temperature
condensate in the accreting material: FeH041
C0050013S003 and FeH063C007O023S005. It
should be noted that Si was omitted without
justification.
Badding et al. (1991, 1992) investigated the
reaction between iron and hydrogen in situ in a
diamond-anvil cell up to 620 kbar, at room ternperature, using X-ray diffraction with synchrotron
radiation. They noticed a sudden increase in volume of iron at 35 kbar, corresponding to the
formation of the FeH hydride, which they found
stable
up determined
to the maximum
pressureand
investigated,
and they
its structure
its equation of state. They also performed thermodynamic calculations relative to the reaction
(2+x)Fe+H20-*2FeH+FeO

and found that it was favoured at high temperatune and pressure. From the experimentally determined equation of state, they found that the
core density deficit can be accounted for by more
than 40 mol % hydrogen.

330

f.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

6. Carbon
Although Birch (1952) mentioned it, carbon
was only occasionally given lip service. Carbon is
available in the accretional material and it is
soluble in iron at low pressure (4.3 wt.% at the 1
atm eutectic point). However, Ringwood (1966)
eliminated it for the same reason as hydrogen,
because it formed interstitial solid solutions with
iron. Brett (1976) even thought that it could increase the density of iron. This reasoning does
not take into account the expansion that can
occur around an interstitial, and indeed carbon,
albeit modestly, lowers the density of liquid iron
near its melting point (Ogino et al., 1984). Ringwood (197) eventually recognized that carbon
could decrease the density of iron at high pres-

retained in appreciable quantities during accretion. He allowed 1 wt.% C at most in the core.
The only serious investigation of carbon in the
core is very recent. Wood (1993) first addressed
the question of carbon volatility, and showed that
the volatility of carbon could be greatly reduced
by even the modest pressures (less than 50 kbar)
at which the differentiation of the Earths core is
thought to occur. He calculated the solubility of
carbon in liquid Fe in equilibrium with the gas
produced from Cl carbonaceous chondrites and
found that, between 1 and 10 kbar, carbon would
enter liquid iron at concentrations between 2 and
4 wt.%. He then proceeded to calculate the phase
diagram of the FeC system at high pressures,
using thermodynamic properties of y-Fe, C, Fe,
FeC liquids and Fe3C (the equation of state of

sure, but pointed out that it was too volatile to be

Fe3C, for which there are no experimental data,

300c

2400

/.

2600

2200

-.

2000

1~

F.3C.

Fe

~oo

Fe3C

1200
S

w.i:ht % C:,bon

2600

tO

(b)

Weight % CirbOti

tO

/
5200

Ca,bon

2400

Liquid (L)

00~

____

(c)

Fe, Fe3C

ash Ic

looc

(a)

Grapt~iIe.
Liquid
Fe3C,

_______

1400

Liquid

Weight % Carbon

(d)

Weight % Carbon

Fig. 13. Phase diagram of the system FeC at 1 bar (a), 50 kbar (b), 150 kbar (c) and 1.36 Mbar (d) (after Wood, 1993).

f-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

was estimated), as well as constraints from experiments at 30 and 50 kbar. It was found that,
between atmospheric pressure and 150 kbar (Fig.
13), the stability field of the carbide Fe3C increases dramatically, and the eutectic composilion shifts to even lower carbon concentrations
(2.1 wt.% C at 150 kbar against 4.3% at 1 bar);
extrapolation to the pressure of the coremantle
boundary shows an enhancement of these featunes. It therefore appears that carbon cannot
contribute more than half of the budget of light
elements in the core, if melt segregation occurs at
low to moderate pressures. Wood then investigated the FeCS system and found that the
miscibility gap gradually closes with increasing
temperature and pressure: at core temperatures,
most of the FeSC liquids with a composition
consistent
density
of the core
would
formwith
one the
stable
liquid.deficit
Furthermore,
the
extrapolations yield the result that, for even very
low values of the C/S ratio, the first phase to
crystallize would be the carbide Fe
3C, thus leading Wood to the conclusion that the inner core
probably consists of Fe3C.
7. Coremantle equilibrium or disequilibrium?
The controversy about coremantle equilibnium or absence thereof essentially hinges on
three points: the abundances of siderophile elements in the upper mantle, the state of oxidation
of the mantle and the segregation process of the
core liquid,
(1) The abundance of siderophile elements
(Ni, Co, Au, Pt, etc.) in the mantle, as determined from analyses of upper-mantle penidotites,
is much higher than the concentrations expected
from equilibrium partitioning between mantle and
core, assuming solar composition for the bulk
Earth (Ringwood, 1966) and assuming the upper
mantle is representative of the whole mantle. For
instance, nickel concentration in the mantle material is about 2000 ppm, whereas 10100 ppm
would be expected for equilibrium partitioning
(Urakawa, 1991). The conclusion that core and
mantle were in gross disequilibrium appeared to
Brett (1971) to be highly unlikely on intuitive

331

grounds. Bretts calculations of reaction kinetics


and partition coefficients with new data (and
assumed values of certain thermodynamic quantities) tended to show that a disequilibrium state
would be unlikely during core formation, but they
were contested by Ringwood (1971). Urakawa
(1991) experimentally investigated the partition
coefficient of nickel between magnesiowstite and
Fe-rich metal up to 170 kbar and 2200C, and
used the results to estimate the partition of nickel
between metal and bulk mantle silicate. He found
that at high pressure, Ni transfers into the silicate
from the iron alloy; but even though he advocated nickel equilibrium partitioning, he still
needed a specific model of mantle differentiation
to account for the nickel concentration in the
present upper mantle.
3 ~/Fe2 + ratio of
(2) basaltic
To account
forand
the penidotites,
Fe
fresh
glasses
a pyrolite
mantle should have a ratio Fe3~/Fe2~~
0.050.1,
whereas it is expected to be at least an order of
magnitude lower for metal silicate equilibrium
(Ringwood, 1966). Brett (1971) suggested that the
rocks of the upper mantle had had time to reequilibrate at higher crustal oxygen fugacities, a
conclusion again disputed by Ringwood (1971).
McCammon (1993), using Mssbauen spec3+/
troscopy,
determined
the Fe with
EFe ratio experimentally
in magnesiowstite
in equilibrium
iron up to 180 kbar, and found that it decreases
with increasing pressure. Extrapolation of her
results to conditions at the top of the lower
mantle suggest values of the ratio Fe3 7EFe
lower than 0.05 for equilibrium, too low to be
consistent with electrical conductivity measurements on lower-mantle material, which point to
conduction by charge transfer between Fe3 + and
Fe2~(see e.g. Shankland et al., 1993); McCammon concluded that it is unlikely that the present
lower mantle is in equilibrium with the core.
The problems of the discrepancy of the
siderophile abundances and the oxidation degree
of the lower mantle have usually been avoided
rather than resolved by devising specific models
of core formation and differentiation (Jones and
Drake, 1986; Newsom and Sims, 1991) or of inhomogeneous accretion of various proportions of
reduced and oxidized condensates (Ringwood,

332

1.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

1977; Wnke, 1981; Wnke and Dreibus, 1988).


As pointed out by Newsom and Sims (1991),
there is a correlation between the degree of success of a model and the number of adjustable
parameters.
(3) A somewhat different question, of a less
geochemical nature, is also relevant to the equilibrium vs. disequilibrium controversy: did the
segregating liquid metal gather into large blobs
that sank rapidly through the slag without having
the time to equilibrate (e.g. Stevenson, 1981), or
did the cone fluid wet the grain boundaries of the
silicate and equilibrate while percolating down
(e.g. Arculus et al., 1990)? The answer depends
on the nature of the solute light element and its
influence on the interfacial energy between oxides and liquid iron.
There are very few measurements of interfacial energy between solid oxides and liquid metals, but they indicate that alloying of iron with
light elements may lower the interfacial energy:
only 1 at.% (less than 0.3 wt.%) of oxygen in
liquid iron lowers the interfacial energy with alumina from 2.4 to 0.6 J ~ 2 (Chaklader et al.,
1981). Surface energy of liquid iron, a good guide

1.8
0

Fe-C

1.6

1 4

Fe-O
~

0.810.6

Fe-S

0.4
0001
0.0005 -

0 01

Weight

prevent local equilibrium.


Local equilibrium, however, even if it is
achieved,are,
in no
the cone andThe
the
mantle
on way
everimplies
were, that
in equilibrium.

\\

12

to interfacial energy, is also reduced in about the


same proportions by oxygen (Fig. 14) and by
sulphur: 1 wt.% S reduces the surface energy by a
factor of three (lida and Guthrie, 1988). Carbon,
however, does not seem to have any significant
effect. There are no data on silicon in iron, but
the value of the surface energy of liquid silicon (a
metal) is 0.865 J m2, lower than that of pure
liquid iron: 1.872 J m2 (lida and Guthrie, 1988).
As it is a rule of thumb that solutes with a low
surface energy tend to lower the surface energy
of the solution, it is reasonable to expect that
silicon will lower the surface energy of liquid
iron.
At high pressures, we have direct observational evidence that liquid iron, containing oxygen in greater concentrations than at 1 atm, as
well as sulphur, completely wets grain boundaries
of oxides (Urakawa et al., 1987; Goarant et al.,
1992).
I therefore think that it is relatively safe to
assume that, if the cone liquid contains several
light elements (a reasonable supposition as we
have seen), some of these will lower the intenfacial energy with solid oxides enough for the liquid
to wet the grain boundaries and percolate down
as thin intergnanular films, at low as well as at
high pressures. It follows that it is possible that
local equilibrium is achieved between the segregating metal and the solid slag. This may not be
the case if the silicate is molten: drops of metal
might then be able to sink rapidly enough to

addition

Fig. 14. Effect of C, S and 0 on the surface tension of liquid


Fe (after lida and Guthrie, 1988).

liquid percolating down samples a whole range of


temperatures, pressures and oxygen fugacities (as
already pointed out by Ringwood (1959)), and the
resulting segregated cone is certainly not in equilibrium with any part of the mantle, let alone the
upper mantle.
Although abundances of elements and oxidation state measured on upper-mantle rocks must
be taken into account in models of Earth accrelion and differentiation, I do not believe that
these data can provide useful constraints on the
nature of the light elements in the outer core.

J.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

8. Conclusion
(1) The density deficit of the core can be
accounted for by most of the light elements considered
Si,S, 0, H, C. However, each element
is subject to metallurgical constraints and is cornpatible with only a limited class of Earth accnetion or core formation models:
(a) Silicon is available and is soluble in iron at
low (and very probably high) pressures; it lowers
the melting point of iron. It is compatible with
early core formation, but it demands very neducing conditions during accretion, or accretion of
an already reduced material.
(b) Sulphur is available and is soluble in iron
at low and high pressures; it lowers the melting
point of iron. It is compatible with early cone
formation. However, geochemical questions concerning the abundance of sulphur are still pending.
(c) Oxygen is available, but it is soluble in iron
in reasonable quantities only at high pressures. It
is incompatible with early core formation, and
can be introduced into the core by reaction with
the mantle only when the Earth has reached a
large size. Also, oxygen does not significantly
lower the melting point of iron and would not be
released during crystallization of the inner core.
(d) Hydrogen is available and might be retamed at low or moderate pressures by forming
an iron hydride. It probably would lower the
melting point of iron,
(e) Carbon is available, but it is not soluble
enough in iron (even at high pressures) to account for the whole density deficit of the core. It
lowers the melting point of iron.
(2) Most of the light elements considered lower
the interfacial energy between liquid iron and
solid slag. The segregating core fluid probably
percolates down, achieving local equilibrium with
the solid mantle at temperature and pressures
varying with depth. If the silicate is molten, even
local equilibrium might not be reached. Hence, it
is very likely that the core is not in equilibrium
with the upper (or lower) mantle. However, establishing whether the core is or is not in equilibrium with the mantle affords no constraint on the
nature of the light elements.

333

(3) The conclusion that there are several light


elements in the outer core seems inescapable. It
is not even obvious that one element should be
particularly dominant.
The best constraints on the proportions of
different elements compatible with the geophysical data (and on the composition of the inner
cone) are, in my opinion, of a metallurgical nature: phase diagrams of iron-rich (and iron
nickel-rich) ternary and quaternary systems
should be experimentally determined up to high
pressures and/on calculated, using thermodynamic data and equations of state.
Acknowledgements
I gratefully acknowledge fruitful discussions
with Francois Guyot. I thank Claude Allgre,
Francois Guyot, Jean-Louis Le Moul and David
Price for reading the manuscript and providing
useful comments. This work was partly supported
by CNRS (URA 734). This is IPG Contribution
1312.
Appendix: Calculation of the density deficit (with
respect to pure Fe) of a solution of light elements
in iron
Most workers announce that the density deficit
of the core is accounted for by a given mass
fraction of light element, but do not generally
explain how this result was obtained. Although
the calculation is simple in principle, I thought it
may be useful to give it here, if only to show that
it is not independent of the equation of state of
the light element. Instead of the light element, it
may be more convenient to consider a light-element-rich end-member (e.g. FeO or FeS) whose
equation of state is known.
Let us consider a solution of a light element X
in Fe. Let PFe be the density of pure iron at core
pressures and temperatures and p~the density of
light element in the same conditions (from equations of state). Let mFe, VFe and m~,v~be the
molar masses and partial molar volumes of Fe
and light element in the core, respectively (con-

334

f-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

sidering partial molar volumes takes into account


the excess volume, which may not be negligible;
however, in view of the fact that the available
data usually are relative to an end-member and
not to the element in solution, the excess volume
is neglected and v,.~ is the molar volume of the
light end-member). The mass fraction of the light
element X is

Ic

(Al)

mFe + m~

The density of the core fluid is


(A2)

mFe + m~
VFe

and Eq. (A6) becomes

1)

0.11

(A8)

As an example, let us consider successively the


cases where the light elements are S, 0 and Si.
Using the equations of state for s-Fe (Brown and
McQueen, 1982), FeO (Jeanloz and Ahrens,
1980), FeS (Brown et al., 1984) and Fe20 wt.%
Si (Balchan and Cowan, 1966), we obtain from
(A6) approximate values of the mass fraction of
light element compatible with a core density
deficit
of 10%:
PFe
1.3

hence fFeO

40 wt.%,

P FeO

We have
and
=(lf~)+f~=+f~(
1
1
1
1
/l
P
PFe
Px
PFe
\P~

1)
P~e

(A3)

f0~9wt.%
137

PFe
PFeS

and

and

=f~~(
/ 1
\Px

PFe

(A4)

Carrying p from (A3) into (A4), we obtain


1 PFe

f~(~

PFe

(AS)

1) + 1

and

3Owt%,

f~~llwt.%
1.12

PFe

PFe

hencefFes

hence fFe-20S

40 wt.%,

f~1~l8wt.%

PFe2OSj

The density deficit for various mass fractions


of each element alone is shown in Fig. Al. The
fact that the curve for silicon is below the curves
for oxygen and sulphur (instead of between them
as might be surmised, as the atomic weight of Si
is intermediate between those of 0 and 5) is

Px

Taking
formula

~P/P~e
=

0.01, we obtain the simple

0.11

f~=________

(A6)

0.25

C)
>.

Px

0.2

.~

0.15

(0

sityIfp~(at
conditions)
and mass X~,
fraction
f~
there core
are several
light elements
of denEq. (AS) becomes
~P

PX

(A7)

V
~
0

C.)

0 1
0.05

0~i~
0

0.05

0.1

0.15

0.2

0.25

Mass fraction of light element


1)

PFe
px

Fig. Al. Core density deficit as a function of the mass fraction


of oxygen, sulphur and silicon.

1.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

________________________________________
I

0.2

5
.._

0.15

~43J

0.1
en

Ct)
(0

0.05

_______________________________________

sure chemistry of hydrogen in metals: in-situ study of iron


hydride. Science, 253: 421424.
Badding, J.V., Mao, H.K. and Hemley, R.J., 1992. High-pressure crystal structure and equation of state of iron hydride: implications for the Earths core. In: Y. Syono and
M.H. Manghnani (Editors), High-pressure Research: Application to Earth and Planetary Science. Terrapub, Tokyo,

C
0

335

0.05
Mass
fraction 0.1
of s

0.15

Fig. A2. Mass fractions of Si, S and 0 giving a core density


deficit of 10%.

probably due to the large negative excess volume


of silicon in solution in liquid iron: 36% for the
composition FeSi (Wilson, 1965), which causes
the density of the solution to be greater than that
of an ideal solution. If silicon, sulphur and oxygen
were simultaneously present, the composition of
the alloys giving a core density deficit of 10%
could be found from (A8) (Fig. A2).

References
Ahrens, TJ., 1979. Equations of state of iron sulfide and
constraints on the sulfur content of the Earth. J. Geophys.
Res., 84: 985998.
Ahrens, T.J. and Jeanloz, R., 1987. Pyrite shock compression,
isentropic release and composition of the Earths core. J.
Geophys. Res., 92: 1036310375.
Alder, B.J., 1966. Is the mantle soluble in the core? J. Geophys. Res., 71: 49734979.
Allgre, C.J., Dupr, B. and Brvart, 0., 1982. Chemical
aspects of the formation of the core. Philos. Trans. R. Soc.
London, Ser. A, 306: 4959.
Allgre, CJ., Poirier, J.P., Humler, E. and Hofmann, A.,
1994. The chemical composition of the Earth in preparation.
Anderson, W.W., Svendsen, B. and Ahrens, T.J., 1989. Phase
relations in iron-rich systems and implications for the
Earths core. Phys. Earth Planet. Inter., 55: 208220.
Arculus, R.J., Holmes, R.D., Powell, R. and Righter, K., i990.
Metalsilicate equilibria and core formation. In: H.E.
Newsom and J.H. Jones (Editors), Origin of the Earth,
Oxford University Press, Oxford, pp. 251271.
Badding, J.V., Hemley, R.J. and Mao, H.K., 1991. High-pres-

pp. 363371.
Balchan, A.S. and Cowan, G.R., 1966. Shock compression of
two ironsilicon alloys to 2.7 Megabars. J. Geophys. Res.,
71: 35773588.
Birch, F., 1952. Elasticity and constitution of the Earths
interior. J. Geophys. Res., 57: 227286.
Birch, F., 1961. Composition of the Earths mantle. Geophys.
J.R. Astron. Soc., 4: 295311.
Birch, F., 1964. Density and composition of mantle and core.
J. Geophys. Res., 69: 43774388.
Boehler, R., 1992. Melting of the FeFeO and the FeFeS
systems at high pressure: constraints on core temperature.
Earth Planet. Sci. Lett., 111: 217227.
Boehler, R., 1993. Temperatures in the Earths core from
melting point measurements of iron at high static pressures. Nature, 363: 534536.
Boness, D.A. and Brown, J.M., 1990. The electronic band
structure of iron, sulfur and oxygen at high pressures and
the Earths core. J. Geophys. Res., 95: 2172121730.
Braginsky, SI., 1964. Magnetohydrodynamics of the Earths
core. Geomagn. Aeron., 4: 698712.
Brett, R., 1971. The Earths core: speculations on its chemical
equilibrium with the mantle. Geochim. Cosmochim. Acta,
35: 203221.
Brett, R., 1976. The current status of speculations on the
composition of the core of the Earth. Rev. Geophys. Space
Phys., 14: 375383.
Brett, R., 1984. Chemical equilibrium of the Earths core and
upper mantle. Geochim. Cosmochim. Ada, 48: 11831188.
Brett, R. and Bell, P.M., 1969. Melting relations in the Fe-rich
portion of the system FeFeS at 30 kbar pressure. Earth
Planet. Sci. Lett., 6: 479482.
Brown, J.M. and McQueen, R.G., 1982. The equation of state
for iron and the Earths core. In: S. Akimoto and M.H.
Manghnani (Editors), High-Pressure Research in Geophysics. D. Reidel, Dordrecht, pp. 611623.
Brown, J.M., Ahrens, T.J. and Shampine, D.L., 1984. Hugoniot data for pyrrhotite and the Earths core. J. Geophys.
Res., 89: 60416048.
Bullen, K.E., 1973. Cores of terrestrial planets. Nature, 243:
6870.
Chaklader, A.C.D., Gill, W.W. and Mehrotra, S.P., 1981.
Predictive model for interfacial phenomena between
molten metals and sapphire in varying oxygen partial pressures. In: J. Pask and A. Evans (Editors), Surfaces and
Interfaces in Ceramics and CeramicMetal Systems,
Plenum, New York, pp. 421431.
Dubrovskiy, V.A. and Pankov, V.L., 1972. On the composition of the Earths core. Acad. Sci. USSR, Phys. Solid
Earth (transl.), 7: 452455.

336

1.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337

Fukai, Y., 1992. Some properties of the FeH system at high


pressures and temperatures and their implications for the
Earths core. In: Y. Syono and M.H. Manghnani (Editors),
High-Pressure Research: Application to Earth and Planetary Science. Terrapub, Tokyo, pp. 373385.
Fukai, Y. and Suzuki, T., 1986. Ironwater reaction under
high pressures and its implication in the evolution of the
Earth. J. Geophys. Res., 91: 92229230.
Goarant, F., Guyot, F., Peyronneau, J. and Poirier, J.P., 1992.
High-pressure and high-temperature reactions between silicates and liquid iron alloys, in the diamond anvil cell,
studied by analytical electron microscopy. J. Geophys.
Res., 97: 44774487.
Herndon, J.M., 1979. The nickel silicide inner core of the
Earth. Proc. R. Soc. London, Ser. A, 368: 495500.
lida, T. and Guthrie, R.I.L., 1988. The Physical Properties of
the Liquid Metals. Clarendon Press, Oxford, 287 pp.
Jacobs, J.A., 1987. The Earths core 2nd edn. Academic Press,
London.
Jacobs, J.A., 1992. Deep interior of the Earth. Chapman and
Hall, London.
Jeanloz, R., 1979. Properties of iron at high pressures and the
state of the core. J. Geophys. Res., 84: 60596069.
Jeanloz, R., 1990. The nature of the Earths core. Annu. Rev,
Earth Planet. Sci., 18: 357386.
Jeanloz, R. and Aiirens, T.J., 1980. Equations of state of FeO
and CaO Geophys. J.R. Astron. Soc., 62: 505528.
Jepheoat, A. and Olson, P., 1987. Is the inner core of the
Earth pure iron? Nature, 325: 332335.
Jones, J.H. and Drake, M.J., 1986. Geochemical constraints
on core formation in the Earth. Nature, 322: 221228.
Kato, T. and Ringwood, A.E., 1989. Melting relationships in
the system FeFeO at high pressures: implications for the
composition and formation of the Earths core. Phys.
Chem. Minerals, 16: 524538.
King, D. and Abrens, T.J., 1973. Shock compression of iron
sulfide and the possible sulfur content of the Earths core.
Nature Phys. Sci., 243: 8284.
Knittle, E. and Jeanloz, R., 1986. High pressure metallization
of FeO and implications for the Earths core. Geophys.
Res. Lett., 13: 15411544.
Knittle, E. and Jeanloz, R., 1989. Simulating the coremantle
boundary: an experimental study of high-pressure reactions between silicates and liquid iron. Geophys. Res.
Lett., 16: 609612.
Knittle, E. and Jeanloz, R., 1991a. The high-pressure phase
diagram of Fe0 940: a possible constituent of Earths core.
J. Geophys. Res., 96: 1616916180.
Knittle, E. and Jeanloz, R., 1991b. Earths coremantle
boundary: results of experiments at high pressures and
temperatures. Science, 251: 14381443.
Knopoff, L. and MacDonald, G.J.F., 1960. An equation of
state for the core of the Earth. Geophys. J.R. Astron. Soc.,
3: 6877.
Knopoff, L. and Uffen, R.J., 1954. The densities of cornpounds at high pressures and the state of the Earths
interior. J. Geophys. Res., 59: 471484.

Lewis, J.S., 1971. Consequences of the presence of sulfur in


the core of the Earth. Earth Planet. Sci. Lett., 11: 130134.
Loper, D.E., 1978. The gravitationally powered dynamo. Geophys. J. R. Astron. Soc., 54: 389404.
MacDonald, G.J.F. and Knopoff, L., 1958. On the chemical
composition of the outer core. Geophys. J.R. Astron. Soc.,
1: 284297.
Mao, H.K., Wu, Y., Chen, L.C., Shu, J.F. and Jephcoat, A.P.,
1990. Static compression of iron to 300 GPa and Fe
0 8Ni02
alloy to 260 GPa: implications for composition of the core.
J. Geophys. Res., 95: 2173721742.
Mason, B., 1966. Composition of the Earth. Nature, 211:
616618.
McCammon, C., 1993. Effect of pressure on the composition
of the lower mantle end member Fe,O. Science, 259:
6668.
McCamrnon, C.A., Ringwood, A.E. and Jackson, I., 1983.
Thermodynamics of the system FeFeOMgO at high
pressure and temperature and a model for formation of
the Earths core. Geophys. J.R. Astron. Soc., 72: 577595.
Newsom, H.E. and Sims, K.W.W., 1991. Core formation during early accretion of the Earth. Science, 252: 926933.
Ogino, K., Nishiwaki, A. and Hosotani, Y., 1984. Density of
molten FeC alloys. Nippon Kinzoku Gakkaishi, 48:
10041010.
Ohtani, E., Ringwood, A.E. and Hibberson, W., 1984. Composition of the core: II. Effect of high pressures on the
solubility of FeO in molten iron. Earth Planet. Sci. Lett.,
71: 94103.
Oversby, V.M. and Ringwood, A.E., 1971. Time of formation
of the Earths core. Nature, 234: 463465.
Raghavan, V., 1988. Phase diagrams of ternary iron alloys.
Part 2: Ternary systems containing iron and sulphur. Indian Institute of Metals, Calcutta.
Rama Murthy, V. and Hall, H.T., 1970. The chemical composition of the Earths core: possibility of sulfur in the core.
Phys. Earth Planet. Inter., 2: 276282.
Rama Murthy, V. and Hall, H.T., 1972. The origin and
chemical composition of the Earths core. Phys. Earth
Planet. Inter., 6: 123130.
Ramsey, W.H., 1949. On the nature of the Earths core. Mm.
Nat. R. Astron. Soc. Geophys. Suppl., 5: 409426.
Ringwood, A.E., 1959. On the chemical evolution and density
of planets. Geochim. Cosmochim. Acta, 15: 257283.
Ringwood, A.E., 1961. Silicon in the metal phase of enstatite
chondrites and some geochemical implications. Geochim.
Cosmochim. Acta, 25: 113.
Ringwood, A.E., 1966. Chemical evolution of terrestrial planets. Geochim. Cosmochim. Acta, 30: 41104.
Ringwood, A.E., 1977. Composition of the core and implications for the origin of the Earth. Geochem. J., 11: 111135.
Ringwood, A.E., 1989. Significance of the terrestrial Mg/Si
ratio. Earth Planet. Sci. Lett., 95: 17.
Ringwood, A.E. and Hibberson, W., 1990. The system FeFeO
revisited. Phys. Chem. Minerals, 17: 313319.
Ringwood, A.E. and Hibberson, W., 1991. Solubilities of
mantle oxides in molten iron at high pressures and tem-

1.-P. Poirier/Physics of the Earth and Planetary Interiors 85 (1994) 319337


peratures: implication for the composition of the Earths
core. Earth Planet. Sci. Lett., 102: 235251.
Shankland, T.J., Peyronneau, J. and Poirier, J.P., 1993. Electrical conductivity of the Earths lower mantle. Nature,
366: 453455.
Sherman, D.M., 1989. The nature of the pressure-induced
metallization of FeO and its implications to the coremantie boundary. Geophys. Res. Lett., 16: 515518.
Sherman, D.M., 1991. Chemical bonding in the outer core:
high-pressure electronic structures of oxygen and sulfur in
metallic iron. J. Geophys. Res., 96: 1802918036.
Stevenson, D.J., 1977. Hydrogen in the Earths core. Nature,
268: 130131.
Stevenson, D.J., 1981. Models of the Earths core. Science,
214: 611619.
Stewart, R.M., 1973. Composition and temperature of the
outer core. J. Geophys. Res., 78: 25862597.
Suzuki, T., Akimoto, S. and Fukai, Y., 1984. The system
ironenstatitewater at high pressures and temperatures

formation of iron hydride and some geophysical implications. Phys. Earth Planet. Inter., 36: 135144.
Svendsen, B., Anderson, W.W., Ahrens, T.J. and Bass, J.D.,
1989. Ideal FeFeS, FeFeO phase relations and Earths
core. Phys. Earth Planet. Inter., 55: 154186.
Urakawa, 5., 1991. Partitioning of nickel between magnesiowstite and metal at high pressure: implications for core
mantle equilibrium. Earth Planet. Sci. Lett., 105: 293313.
Urakawa, S., Kato, M., and Kumazawa, M., 1987. Experimen-

337

tal study on the phase relations in the system FeNiOS


up to 15 GPa. In: M.H. Manghnani and Y. Syono (Editors),
High-Pressure Research in Mineral Physics. Terrapub,
Tokyo, pp. 95111.
Urey, H.C., 1960. On the chemical evolution and density of
planets. Geochim. Cosmochim. Acta, 18: 151153.
Usselman, T.M., 1975. Experimental approach to the state of
the core: Part 1. The liquidus relations of the Fe-rich
portion of the FeNiS system from 30 to 100 kbar. Am.
J. Sci., 275: 278290. Part 2. Composition and thermal
regime. Am. J. Sci., 275: 291303.
Wnke, H., 1981. Constitution of terrestrial planets. Philos.
Trans. R. Soc. London, Ser. A, 303: 287302.
Wnke, H. and Dreibus, G., 1988. Chemical composition and
accretion history of terrestrial planets. Philos. Trans. R.
Soc. London, Ser. A, 325: 545557.
Williams, 0. and Jeanloz, R., 1990. Melting relations in the
ironsulfur system at ultra-high pressures: implications for
the thermal state of the Earth. J. Geophys. Res., 95:
1929919310.
Wilson, JR., 1965. The structure of liquid metals and alloys.
Metall. Rev., 10: 381590.
Wood, B., 1993. Carbon in the core. Earth Planet. Sci. Lett.,
117: 593607.
Yagi, T., Suzuki, T. and Akimoto, 5., 1985. Static compression
of wstite (Fe
0 990) to 120 GPa. J. Geophys. Res., 90:
87848788.

Você também pode gostar