Martin - Hammouda - 2011 - Role of Iron and Reducing Conditions On The Stability of Dolomite + Coesite

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Eur. J. Mineral.

2011, 23, 5–16


Published online November 2010

Role of iron and reducing conditions on the stability of dolomite þ coesite


between 4.25 and 6 GPa – a potential mechanism for diamond formation during
subduction
AUDREY M. MARTIN1,2,3,4,* and TAHAR HAMMOUDA1,2,3

1
Clermont Université, Université Blaise Pascal, Laboratoire Magmas et Volcans, BP 10448, 63000 Clermont-
Ferrand, France
2
CNRS, UMR 6524, LMV, 63038 Clermont-Ferrand, France
3
IRD, R 163, LMV, 63038 Clermont-Ferrand, France
4
NASA Johnson Space Center, Mailcode KT, 2101 NASA Parkway, Houston, TX 77058, USA
*Corresponding author, e-mail: audrey.m.martin@nasa.gov

Abstract: We have investigated the effect of iron and oxygen fugacity on dolomite þ coesite stability during subduction. For redox
conditions buffered by the assemblage itself, the presence of iron (Fe/(Fe þ Mg) ca. 0.4) lowers the decarbonation reaction by about
200  C at 4.25 GPa and by 300  C at 5.5 GPa, compared to the iron-free reaction. Clinopyroxene and CO2 form by a decarbonation
reaction similar to the iron-free system. At low temperature, however, graphite replaces CO2 through redox interactions with iron in
the carbonate. Melting occurs approximately 100  C above decarbonation and a carbonatitic melt is produced. In a second series of
experiments, we imposed lower oxygen fugacity by adding molybdenum, in order to study the potential redox mechanisms in contact
with the peridotitic mantle. In these samples, we observe systematic carbon reduction producing the assemblage clinopyroxene þ
graphite. Our results show that the stability of dolomite þ coesite on a subduction path is limited by redox interactions, in addition to
pressure and temperature. As our experiments were run in the stability field of diamond, we also demonstrate that diamond may form
from dolomite þ coesite during subduction. Chemical diffusion of iron and oxygen in the slab and at the slab/mantle interface appears
to be a key parameter to determine at what pressure–temperature conditions this may happen.
Key-words: dolomite, coesite, subduction, high pressure, redox interactions, decarbonation, experimental petrology, diamond.

1. Introduction composition, but dolomite CaMg(CO3)2, ferroan dolomite


(or ankerite) Ca(Mg,Fe)(CO3)2 and siderite FeCO3 are also
Analyses of mantle xenoliths show that carbon is present in found (Laverne, 1993; Nakamura & Kato, 2004). In order
the Earth’s mantle in the form of oxidized or reduced to constrain the destabilization conditions of carbonates
minerals (Ionov et al., 1993; Nixon, 1995; Viljoen, 1995; during subduction, low-pressure phase diagrams in the
Wang et al., 1996), fluids (Schrauder & Navon, 1993; system CaCO3 – MgCO3 – FeCO3 were determined by
Klein-Ben David et al., 2007), or dissolved in silicate and experiments (Goldsmith et al., 1962) and thermodynamic
carbonate melts (Schiano et al., 1994; Yaxley et al., 1998; modelling (Anovitz & Essene, 1987; Davidson, 1994).
Moine et al., 2004). Zhang & Zindler (1993) and Marty & However, we still lack high-pressure data on ternary solid
Tolstikhin (1998) infer from mass balance calculations that solutions. Interactions with the surrounding material were
the current concentration of carbon in the mantle (between also considered. Luth (1993, 1995, 1999) claims that dec-
80 and 400 ppm) can only be explained by the subduction arbonation takes place at higher pressure in sediment or
of oceanic material, and in particular, oceanic carbonates. altered basalt than in a peridotite. This would imply that a
Plank & Langmuir (1998) determined that 7 wt% of the depth interval exists, where the decarbonation of the super-
Global Subducting Sediment is of calcite CaCO3 composi- ficial layers of the slab can lead to the crystallization of
tion. Magnesium and iron carbonate are also observed, carbonate in the mantle wedge by CO2 transfer. Another
although more rare. Alteration of the oceanic crust by important parameter to consider is the oxidation state of the
hydrothermal activity also leads to the crystallization of system, because of the contrast between subducted mate-
carbonates which represent a large potential carbon reser- rial and peridotitic mantle. Redox conditions of silicate
voir for high pressure recycling (Staudigel et al., 1989; Alt (peridotitic) lithologies at high pressure are imposed in
& Teagle, 1999). These carbonates are mainly of calcite most cases by iron oxidation state (Ballhaus, 1993; Canil

0935-1221/10/0022-2067 $ 5.40
DOI: 10.1127/0935-1221/2010/0022-2067 # 2010 E. Schweizerbart’sche Verlagsbuchhandlung, D-70176 Stuttgart
6 A.M. Martin, T. Hammouda

& O’Neill, 1996), whereas oxygen fugacity (fO2) of carbo- GPa – in the cratonic mantle. Redox conditions in the
nate lithologies is likely to be controlled by carbon oxida- mantle wedge might be a little higher – around or above
tion state. FMQ – due to metasomatism (Parkinson & Arculus, 1999;
One of the decarbonation reactions investigated experi- Wang et al., 2007). However, percolation of reduced fluids
mentally by Luth (1995) for sediment or altered basalt was: coming from the slab might also lower the fO2 in the
overlying mantle (Frost & McCammon, 2008).
dolomite þ 2 coesite ¼ diopside þ 2 CO2 (1) Considering both pressure effect and chemical uncertain-
ties at subduction zones, these last experiments aim at
constraining potential redox mechanisms in subducted
CaMgðCO3 Þ2 þ 2 SiO2 ¼ CaMgðSiO3 Þ2 þ 2 CO2 dolomite þ coesite in contact with the mantle wedge and,
in particular, diamond formation from the assemblage
Luth found that the assemblage dolomite þ coesite is dolomite þ coesite.
stable to at least 6 GPa on a hot subduction path. In our
study, we first investigate the role of iron on this reaction
between 4.25 and 6 GPa over the temperature range 800 to 2. Experiments
1425  C. Iron was chosen because (1) the presence of iron-
bearing dolomite (or ankerite) is reported in various places
2.1. Starting material and fO2 buffering
(Laverne, 1993; Nakamura & Kato, 2004, see above) and
(2) iron is thought to control the fO2 in the mantle and The starting material is a mixture of natural high-purity
might diffuse into the subducted material. Iron (and mag- quartz from Saint-Paul-La-Roche (France) and a dolomite
nesium) might also diffuse into carbonates by exchange containing iron (Fe0.36Mg0.66Ca0.98(CO3)2) from Ben
reactions with subducted silicates occurring at high pres- Gasseur (Tunisia), which were crushed and dried at 140
sure. For example, studies on carbonated eclogites by 
C for 10 h. Compositions of the starting mixtures are
Hammouda (2003) and Yaxley & Green (2004) show that given in Tables 1 and 2. In the first series of experiments,
pure calcite is transformed to Fe-bearing dolomite by reac- oxygen fugacity is, in theory, imposed intrinsically by the
tions with garnet and clinopyroxene. Our experiments aim oxidation state of carbon whose concentration in the mix-
at constraining iron/carbon interaction mechanisms. ture (expressed in moles) is five times greater than that of
Therefore, fO2 is buffered intrinsically by the samples iron. In practice, oxidation conditions were established
themselves, leading to relatively oxidizing conditions using reactions occurring in our samples (see
(see Discussion). In a second series of experiments, we Discussion). AuPd capsules welded at the two extremities
study the role of reducing conditions on dolomite þ coesite were used in order to limit oxygen and hydrogen exchange
stability by adding molybdenum. Oxygen fugacity of the with the surrounding materials. As our experiments were
Mo/MoO2 (MMO) buffer, studied by O’Neill (1986) at 1 anhydrous, a double capsule technique using an external
bar, lies around IWþ0.5/þ1 and FMQ3.5/4.5 at 5 GPa buffer and H was excluded. For reducing experiments,
(Fig. 1). This value is in the range of oxidation conditions samples were loaded into a rolled and folded Mo foil that
determined by previous authors for peridotitic mantle was introduced into a platinum capsule welded at the two
(Ballhaus, 1993; McCammon et al., 2001). In particular, extremities. Computation of MMO equilibrium shows that
Frost & McCammon (2008) found that fO2 decreases when it lies approximately 3.5–4 log unit below the fayalite-
pressure increases – FMQ3 to FMQ4.5 between 4 and 6 magnetite-quartz buffer (FMQ), 0.5–1 log unit above the
iron-wüstite (IW) buffer, and about 2–3 log units below
diamond/CO2 (CCO from Frost 1998) or diamond/carbo-
nate buffers (DCDD, dolomite þ 2 coesite ¼ diopside þ 2
diamond þ 2 O2, relevant to eclogite or sediment after
Luth (1993), and EMOD, enstatite þ magnesite ¼ forster-
ite þ diamond þ O2, relevant to peridotite after Eggler &
Baker (1982)) at 5 GPa (Fig. 1). Thermodynamic calcula-
tions were made using the relations:
dG
f O2 ¼  (2)
RT

dG ¼ dH  TdS þ ðP  1Þ dV solid (3)

where G represents the Gibbs free energy, R the gas con-


stant (¼ 8.3144 J K1 mol1), T the temperature, H the
Fig. 1. Temperature – log (fO2) diagram comparing the position of
Mo–MoO2 buffer to Fe–FeO (IW), Fayalite–Magnetite–Quartz
enthalpy, S the entropy, P the pressure and V the volume.
(FMQ) and equilibria involving carbon species at 5 GPa. Data used for calculations were taken from Holland &
Calculations are described in the text. Powell (1998), Robie & Hemingway (1995) or the NIST
Role of iron and low fO2 on dolomite þ coesite stability 7

Table 1. Experimental conditions of the first series of experiments (intrinsic fO2).

Serie 1 Run# P (GPa) T ( C) Duration Wt% Dolomite Wt% Quartz Assemblage


ME625 4.25 800 6h 74.8 25.2 dol þ cs þ graph
ME710 4.25 950 94 h 75.0 25.0 dol þ cs
ME831 4.25 1100 18 h 74.5 25.5 dol þ cs þ cpx
ME380 4.25 1250 6h 61.7 38.3 dol þ cs þ cpx þ melt
ME614 5.5 800 6 h 30 min 74.7 25.3 dol þ arag þ mc þ cs þ graph
ME729 5.5 1000 2 h 15 min 75.0 25.0 dol þ cs
ME384 5.5 1300 6h 61.7 38.3 dol þ cs þ melt
ME382 5.5 1350 6h 64.1 35.9 dol þ cs þ melt
ME398 5.5 1425 6h 71.5 28.5 cs þ cpx þ melt
ME619 6 800 6 h 30 min 73.1 26.9 arag þ Fe-mc þ cs þ graph
ME592 6 900 6h 72.9 27.1 dol þ arag þ Fe-mc þ cs þ graph
ME645 6 1000 6h 75.2 24.8 dol þ cs þ graph
ME708 6 1100 34 h 75.0 25.0 dol þ cs
ME832 6 1200 18 h 74.5 25.5 dol þ cs þ cpx

Abbreviations: cpx ¼ clinopyroxene, cs ¼ coesite, dol ¼ dolomite, graph ¼ graphite, mc ¼ magnesite, arag ¼ aragonite. The presence of
CO2 is not indicated here.

Table 2. Experimental conditions of the second series of experiments (reducing conditions).

Serie 2 Run# P (GPa) T ( C) Duration Wt% Dolomite Wt% Quartz Assemblage


ME504 4.25 950 6h 74.6 25.4 dol þ cs þ cpx þ graph þ Fe-Mg oxide
ME470 4.25 1050 6h 72.9 27.1 dol þ cs þ cpx þ melt þ graph
ME464 4.25 1150 6h 63.4 36.6 arag þ cs þ cpx þ melt þ graph
ME483 5 1150 6h 74.8 25.2 dol þ cs þ cpx þ melt
ME474 5.5 1150 6h 75.1 24.9 dol þ cs þ cpx þ melt þ graph
ME517 5.5 1250 6h 75.1 24.9 dol þ cs þ cpx þ melt
ME469 5.5 1300 6h 74.8 25.2 cpx þ melt þ graph
ME826 6 900 48 h 74.5 25.5 dol þ cs þ cpxþ graph
ME505 6 1050 6h 74.4 25.6 dol þ cs þ cpx þ graph
ME484 6 1100 6h 73.9 26.1 dol þ cs þ cpx
ME471 6 1150 9 min 74.3 25.7 dol þ cs þ cpx þ graph
ME485 6 1200 6h 74.3 25.7 dol þ cs þ cpx þ melt
ME468 6 1250 6 h 30 min 73.5 26.5 cpx þ cs þ melt þ graph
ME512 6 1300 6h 74.5 25.5 cpx þ cs þ melt

Abbreviations: cpx ¼ clinopyroxene, cs ¼ coesite, dol ¼ dolomite, graph ¼ graphite, arag ¼ aragonite.

website. The only molybdenum oxide observed in the 2.0 mm diameter capsules (see above). Before loading,
samples was MoO2 and Mo was still present at the end of platinum capsules were dried using the same method as
all experiments. for AuPd capsules and Mo foils were dried at 140  C for 1
h. Stepped LaCrO3 furnaces with Mo electrodes were used
for the two series of experiments. The furnace was sepa-
2.2. Experimental protocol rated from the capsule by an MgO spacer. A W95Re5-
W74Re26 thermocouple was inserted axially such that the
Experiments were conducted on a 1000 ton multi-anvil junction was in contact with one end of the capsule. Cr2O3-
apparatus (octahedral symmetry) at Laboratoire Magmas doped MgO octahedra were used in our experiments.
et Volcans (Hammouda, 2003) using an 18/11 type assem- Zirconia was placed between the furnace and the octahe-
bly (octahedron edge length in mm/truncation length of the dron for thermal insulation. Pressure calibration was made
anvils in mm). Experimental conditions are reported in using the coesite/stishovite (Akaogi, 1995) and fayalite/
Tables 1 and 2. In the first series of experiments (buffered spinel (Katsura & Ito, 1989) transitions at 1000  C.
intrinsically), the starting mixture was loaded in a 2.0 mm Pressure is believed to be accurate within 0.5 GPa
diameter Au80Pd20 (by weight) capsule. This capsule was and temperature uncertainty is about 20  C. The
dried at 1000  C before loading. It was then welded at the temperature gradient inside the capsules is not known;
two extremities in order to prevent CO2 or melt loss. In the however, considering the short length of every capsule
second series of experiments, the samples were contained (2.5–3.0 mm), we believe that this gradient is very small.
in Mo internal capsules introduced in outer platinum Pressure was first increased in about 2–3 h (depending on
8 A.M. Martin, T. Hammouda

pressure). Then the samples were heated in less than 20 min (Hammouda et al., 2010) with a eutectoid texture
by increasing the electrical power in the LaCrO3 furnace. (Fig. 2b) similar to that observed by Luth (2001).
Temperature was maintained at constant value by an auto- Graphite is present in all these samples. At 4.25 GPa –
matic regulation (Eurotherm controller). Running duration 1100  C and 6 GPa – 1200  C, clinopyroxene forms at
for most experiments was 6 h, based on Luth (1995). Some grain boundaries between dolomite and coesite (Fig. 2c).
longer-duration experiments were performed to test kinetic Melting occurs below 1250  C at 4.25 GPa, producing a
effects, in particular at low temperature (Tables 1 and 2). carbonate-silicate melt (Fig. 2d) in association with
Temperature was quenched by shutting off the electrical dolomite, coesite and clinopyroxene. At 5.5 GPa, carbo-
power (drop to ,100  C temperatures in 4–8 s). nate-silicate melt appears at 1300  C, in association with
Decompression was programmed for durations of 3–7 h. coesite and dolomite up to 1350  C, and with coesite
and clinopyroxene at 1425  C. The presence of a gas
phase was observed in some samples, either in the form
of bubbles going out from the capsule during polishing,
2.3. Sample preparation and analyses
or of rounded holes in some sample sections. Because
Recovered samples were mounted in resin and polished we could not perform systematic observations of fluid
using alumina powder and ethanol. Bright veins or pockets occurrence in our sample, we did not report them in the
of graphite were identified by optical reflected-light micro- Tables. The presence of clinopyroxene is sufficient evi-
scopy. Raman spectroscopy (Jobin-Yvon T64000, 514.5 dence for the decarbonation of the starting material.
nm) was used to study the carbon species (graphite or In the second series of experiments, graphite is found in
diamond). The volume of sample excited by the laser is almost all experimental samples (plucking during grinding
about 1 mm3. The laser power on the sample was main- and polishing is believed to be the cause of the absence of
tained at a low value (,15 mW) during analyses to mini- graphite in some samples). Clinopyroxene is also observed
mize sample alteration (which was controlled by optical below 1000  C at 4.25 GPa and below 1200  C at 6 GPa
microscopy). Phase relations were determined by scanning (Table 2; Fig. 3a). Above these temperatures, carbonate –
electron microscopy using energy-dispersive X-ray spec- silicate melt forms in association with coesite, dolomite,
trometry (Jeol 5910-LV). Chemical analyses were clinopyroxene and graphite in various proportions
obtained with an electron microprobe (Cameca SX100 (Fig. 3b–d).
using (rz) data reduction procedure) using wavelength- To determine if the graphite found in our samples formed
dispersive spectrometry (accelerating voltage of 15 kV; during the experiments or during the quench (at the end of
beam current of 15 nA for silicate minerals and 8 nA for the experiments), we investigated its structure by using
carbonate minerals or melts with a 5 mm defocused beam). Raman spectroscopy (Fig. 5). We focused on the first-
No damage was observed at the carbonate surface after order region of the Raman spectra of carbon, which
analyses. The standards used for the microprobe calibra- shows several bands corresponding to different vibration
tions are wollastonite for Ca and Si, Al2O3 for Al, forsterite modes of the carbon bonds (Beyssac et al., 2003). We looked
for Mg, fayalite for Fe, albite for Na and molybdenum for at three bands in particular: the G band (at 1580 cm1)
Mo. Counting time was 10 s on the peak for all elements which is due to the stretching vibration in aromatic layers
except for Mo (30 s) and Fe (40 s). Because of complex of graphite (E2g2 mode), the D1 band (at 1350 cm1)
quenching textures, melt composition was determined, called the defect band (A1g mode) and the D2 band (at
when possible, either with a large defocused beam or 1620 cm1) which is always present when D1 is present.
with a scanning beam over a large area. The last two bands characterize poorly ordered graphite
or carbonaceous material (CM). In our first series of
experiments, Raman spectra of graphite, which was only
found in the low-temperature experiments, show that it
3. Results is disordered. The intensity of D1 (at 1351–1360 cm1)
is slightly larger than that of G (at 1592–1600 cm1).
3.1. Phase relations D2 appears at approximately 1620 cm1, but cannot be
resolved from G and causes an offset of G to higher
Phase assemblages observed in the first and second series frequencies. In Series 2 (reducing experiments), where
of experiments are reported in Tables 1 and 2, respectively. graphite was found in almost all samples, ordering of
Typical sample textures are shown in Fig. 2 and 3 and graphite is a function of the presence or absence of melt.
phase diagrams of the two series of experiments are At subsolidus conditions, D1 is about the same height as
reported in Fig. 4. G, implying disorder. Above the solidus, G (at
In the first series of experiments, the dolomite þ coesite 1354–1356 cm1) is about three times higher than D1
decarbonation reaction is bracketed between 950 and 1100  C (at 1578–1581 cm1), i.e. the graphite is quite ordered
at 4.25 GPa, and between 1100 and 1200  C at 6 GPa and it appears that the presence of melt favours graphite
(Table 1; Fig. 2a). In the dolomite þ coesite stability field, ordering. From this strong link between ordering and
graphite is also found at 4.25 GPa – 800  C and at 6 melting, we conclude that graphite formed at high tem-
GPa – 1000  C. At 5.5 GPa – 800  C and 6 GPa below 900  C, perature (during the experiments) and is not a quench
dolomite breaks down to aragonite plus magnesite phase – otherwise, the D1/G value would be much more
Role of iron and low fO2 on dolomite þ coesite stability 9

Fig. 2. SEM images (BSE) of typical experimental textures observed in samples of the first series of experiments. (a) Dolomite (light grey) and
coesite (dark grey) in a non-decarbonated sample. (b) Fine-grained eutectoid texture formed by aragonite (lighter) and magnesite (darker). Black
phase is graphite. Dark grey phase is coesite. White phase is a small piece of the capsule which moved during polishing. (c) Clinopyroxene (white
phase) formed at grain boundaries between dolomite (light grey) and coesite (dark grey) in a decarbonated sample. (d) Carbonate-silicate melt
with dendritic clinopyroxene exsolved during quench. Dark grey phase is coesite and light grey phase is clinopyroxene.

scattered, as it would depend on quench speed and phase composition of which varies depending on redox condi-
relations. Moreover, a graphite precipitation from CO2 tions (Fig. 6 and supplementary material S2 and S3). In the
fluid during quenching (Wopenka & Pasteris, 1993; first series of experiments, Mg# (defined by Mg/(Mg þ
Cesare, 1995) would not explain why only low-tempera- Fe)) of cpx is 0.60 while Mg# of carbonate is 0.67. In
ture, non-decarbonated samples contain graphite in Series 2 (reducing conditions), Mg# of cpx and carbonate
Series 1. In addition, Luque et al. (1998) also showed are equivalent but with a larger variability, between 0.60
that natural (low temperature) fluid-deposited graphite and 0.72. A more remarkable feature is the difference
(except that contained in inclusions) exhibits a high between XCa values of cpx (defined by Ca/(Ca þ Mg þ
degree of order, whereas the graphite formed in our Fe) in Series 1 and 2. While XCa in carbonate remains
low-temperature experiments is disordered. Considering constant (0.50), XCa in cpx is 0.35 in Series 1, but 0.43
all this, we can definitely rule out graphite precipitation to 0.49 in Series 2. Therefore, cpx from the first series of
from a fluid during quench. Rather, graphite formed experiments is less calcic than those formed at more redu-
during the experiments by a redox reaction between Fe cing conditions. Moreover, the Ca content of cpx from
and C inside carbonates in Series 1 (see Discussion). reduced experiments is close to that of carbonate, as
expected from mass-balance calculations. As no Ca-rich
carbonate or calcite was found in the samples, the low Ca
3.2. Phase compositions content of cpx in Series 1 requires the presence of an
additional phase that takes up calcium relative to magne-
The composition of the low-temperature carbonates is sium. The only possibility is a CO2 fluid. Although cal-
given in Table S1, freely available online as cium solubility in CO2 is not well documented from
Supplementary Material on the GeoScienceWorld website experiments at high pressure, natural examples of Ca-
of the journal, linked to the relevant article at http://eur- bearing high density fluids (HDFs) in diamond inclusions
jmin.geoscienceworld.org/. Subsolidus decarbonation of have been reported (Klein-Ben David et al., 2007). This
the samples produces clinopyroxene (cpx), the kind of fluid is thought to be stable in the diamond
10 A.M. Martin, T. Hammouda

Fig. 3. SEM images (BSE) of typical experimental textures observed in samples of the second series of experiments (reducing conditions).
(a) Decarbonated sample with graphite veins (black) and clinopyroxene (white) which formed at grain boundaries between dolomite (light grey)
and coesite (dark grey). (b) Clinopyroxene rims (white) around coesite (dark grey) in a melted sample (melt is not shown on this picture).
Dolomite is light grey. (c) Carbonate-silicate melt (grey) with dendritic clinopyroxene exsolved during quench (white stripes). Light grey
euhedral crystals are clinopyroxene. Coesite is dark grey and dolomite is grey (at the top of the image). (d) SEI image of molybdenum oxide
(medium grey) embedded in the carbonate-silicate melt (dark grey with white stripes). Molybdenum capsule is at the bottom (light grey).

stability field, i.e. 4 GPa and higher. Additionally, Luth 4. Discussion


(2006) reported significant Mg and Ca solubility in CO2
fluid in equilibrium with diopside at pressures between 3 4.1. Effect of iron on the stability of dolomite þ coesite
and 5 GPa, with preferential uptake of Ca, as in the
present experiments. Unfortunately, polishing difficul- 4.1.1. Phase relations
ties did not allow us to determine precisely carbonate, In the first series of experiments (self-buffered) the
cpx and CO2 fractions in the samples, thus we cannot Fe-bearing dolomite starting material breaks down to mag-
estimate how much Ca dissolved in the vapour in our nesite þ aragonite at about the same conditions as those
experiments.
found by Luth (2001) and Buob et al. (2006) for the Mg
Melting produces carbonate liquids that have a small
end-member (Fig. 4a). Considering the experimental
fraction of dissolved silicate (,15 wt% SiO2;
Supplementary material S4). At the P–T conditions of brackets, the dolomite breakdown boundary in this study is
our study, no immiscibility was observed between the also consistent with the results of Morlidge et al. (2006) on
carbonate and silicate fractions of the melts. In the first Fe end-member (ankerite) stability. According to Luth
series of experiments, melts have a dolomitic composition (1995), the decarbonation reaction dolomite þ 2 coesite
with an XCa value of 0.5. The Mg# value of the solid ¼ diopside þ 2 CO2 occurs at 1500  C at 6 GPa and
phases formed in equilibrium with the melts is 0.60–0.80 1200  C at 4 GPa. In the iron-bearing system, the
for cpx and 0.60–0.72 for carbonates. In clinopyroxene, high-temperature assemblage is favoured and the decarbona-
XCa is between 0.35 and 0.50, whereas in carbonates XCa tion reaction is shifted to lower temperature by at least 200  C
is between 0.50 and 0.70. In the second series of experi- at 4.25 GPa and 300  C at 5.5 GPa (Fig. 4a). Melting
ments, some molybdenum was found in the melt (up to 5 occurs between 1100 and 1250  C at 4.25 GPa and between
wt% MoO3). Thus, the use of the Mo/MoO2 buffer is not 1200 and 1400  C at 6 GPa. Compared to the iron-free system
appropriate for high-temperature – supersolidus – experi- studied by Luth (2006), we note a solidus depression of at
ments in the presently studied chemical system. least 325  C at 4 GPa. In our experiments, the solidus P–T
Role of iron and low fO2 on dolomite þ coesite stability 11

Fig. 5. Raman spectra of graphite in the experimental samples (S1 ¼


Series 1, S2 ¼ Series 2).
Fig. 4. Pressure–temperature diagrams of (a) the first series of
experiments (intrinsic fO2) and (b) the second series of experi-
ments (reducing conditions). The precision on the pressure value is with dolomite2 containing slightly less iron than dolo-
less than 0.5 GPa; the precision on temperature is 20  C. mite1. A few iron oxide grains were found in the gra-
Black points indicate the experimental conditions; diamond sym- phite-bearing samples, but they were too small to obtain
bols represent samples with graphite. Black curves are reactions good microprobe analyses for iron valence state deter-
deduced from our samples. Dashed grey curve represents the Fe- mination. The loss of iron from the dolomite may be
free decarbonation reaction dolomite þ 2 coesite ¼ diopside þ 2
CO2 and the solidus calibrated by Luth (2006). Dol ¼ dolomite,
either compensated by slight Ca and Mg content varia-
cs ¼ coesite, cpx ¼ clinopyroxene, arag ¼ aragonite, mc ¼ tions in the dolomite structure, or by the formation of
magnesite, C ¼ graphite. Solidus in Series 1 is indicated for new phases (carbonates or oxides). We could not, how-
information; however, our experiments do not allow one to deter- ever, observe any variations or new phases, probably
mine if its slope is positive or if it is vertical. because of the very small amounts of iron oxide and
graphite produced.
In order to quantify the oxygen fugacity in the samples,
slope appears to be either positive or vertical whereas Luth we modelled the stability of siderite with regard to
(2006) found a negative slope between 3 and 8 GPa in the magnetite and graphite over a P–T range corresponding to
iron-free system. This is probably an effect of the presence of our experiments. Siderite was chosen as iron-carbonate end-
iron in our system. From all the above discussion, we can member because thermodynamic values for ankerite are
deduce that the sole addition of iron to dolomite lowers the poorly known. The results are illustrated on a polybaric
dolomite þ coesite stability regarding both decarbonation temperature–fO2 section in Fig. 7. Using the redox reaction:
and melting.
6 siderite ¼ 2 magnetite þ 6 C þ 5 O2 (5)

4.1.2. Oxidation conditions 6 FeCO3 ¼ 2 Fe3 O4 þ 6 C þ 5 O2


In these experiments, fO2 is assumed to be buffered by the
samples themselves. Carbonate reduction to graphite was we estimate a value of log(fO2) 7.5 at 6 GPa – 1000  C
observed at low temperature. This can only be explained in our samples. Below this temperature (carbonate and
by a redox reaction with the iron contained in the starting graphite also present), log(fO2) decreases with temperature
dolomite, which could be written in a simplified way: along reaction (5). Above this temperature (graphite
absent), log(fO2) is above reaction (5). At 5 GPa – 900  C,
dolomite1 ¼ dolomite2 þ iron oxide þ graphite (4) log(fO2) 9.4 (FMQ–0.3). Below this temperature,
12 A.M. Martin, T. Hammouda

Fig. 6. Ternary projection (Ca–Mg–Fe) of coexisting carbonate and clinopyroxene compositions at subsolidus conditions: dol ¼ dolomite, cpx
¼ clinopyroxene. Series 1 ¼ experiments with intrinsic fO2; Series 2 ¼ experiments at reducing conditions. Note clustering within each
mineral type, although pressure and temperature conditions vary.

log(fO2) decreases with temperature along reaction (5). dolomite þ 2 coesite þ 2 Mo


Above this temperature, log(fO2) is above reaction (5). ¼ cpx þ 2 C þ 2 MoO2 (7)
These values are about 0.2 log unit fO2 below the CCO
buffer at the same P–T conditions (Fig. 7).
At higher temperature, decarbonation of the sample pro- CaMgðCO3 Þ2 þ 2 SiO2 þ 2 Mo
duces cpx and CO2. Considering the intersection between ¼ CaMgðSiO3 Þ2 þ 2 C þ 2 MoO2
CCO and the decarbonation reaction dolomite þ 2 coesite ¼
cpx þ 2 CO2 (DCDV) that we calibrated experimentally
CO2 absence is deduced from textural evidence and clin-
around 1100  C in the iron-bearing system, we deduce that
opyroxene composition (see Results), and can be explained
the fO2 increases to above 7.5 log units fO2 (FMQ–0.7) at
by a simultaneous occurrence of reactions (1) and (6). By
5 GPa when the decarbonation occurs (all phases of DCDV
lowering the CO2 activity, Mo lowers the temperature of
present and graphite absent). No constraints can be given on
reaction (1). Oxidation conditions in our samples are buf-
the fO2 value between the graphite stability boundary and
fered by reaction (7), which means that, at 5 GPa, the fO2
the decarbonation temperature, nor above the decarbonation
lies somewhere between FMQ–0.5 and FMQ–4.5 at 1000  C
temperature. However, a temperature dependence of the
and between FMQ–1 and FMQ–3.5 at 1300  C (Fig. 1).
oxidation conditions in the samples (log(fO2)9.4 at
Although the oxidation conditions could not be deter-
900  C to above 7.5 at 1100  C in the 5 GPa experiments)
mined more precisely, we demonstrate that dolomite þ
is demonstrated.
coesite stability during subduction may be lower than
previously thought due to redox interactions, and that
diamond or graphite may form from the subduction of
4.2. Effect of reducing conditions on dolomite þ coesite this assemblage. In the natural case, however, its stabi-
stability lity should be dependent on the FeO/Fe2O3 equilibria in
the surrounding silicates.
All the samples synthesized at reducing conditions (Series
2) contain clinopyroxene. Therefore, reducing conditions
imposed by Mo dramatically contract the dolomite þ coe- 4.3. Reactions kinetics and equilibrium achievement
site stability field – by more than 300  C compared to
oxidizing conditions. Reaction mechanisms can be divided 4.3.1. Self-buffered experiments
into two parts: a decarbonation reaction similar to that In our first series of experiments, the starting mixture is
occurring in Series 1 (reaction (1)), and a redox reaction: never completely decarbonated. According to Lüttge &
Mo þ CO2 ¼ MoO2 þ C (6) Metz (1991, 1993), kinetics of the reaction dolomite þ 2
quartz ¼ diopside þ 2 CO2 is very sluggish as soon as
in agreement with our thermodynamic calculations clinopyroxene grains (which grow at the boundary
(Fig. 1). The global reaction can be written: between dolomite and quartz) separate the two starting
Role of iron and low fO2 on dolomite þ coesite stability 13

Fig. 8. Pressure – temperature diagram showing the destabilization


conditions of the dolomite þ coesite assemblage at intrinsic fO2
(Series 1) and reducing conditions (Series 2). (1) Graphite or dia-
mond is produced by the reduction of the subducted dolomite in
Fig. 7. Temperature – log(fO2) diagram illustrating the location of contact with the mantle. (2) ‘Hot’ subduction path crosses the dolo-
siderite (sid) breakdown to magnetite (mt) þ graphite (C) (black mite decarbonation reaction at intrinsic conditions, i.e. far from the
curves). Siderite breakdown equilibrium was modelled using mantle. In the shaded range, graphite or diamond forms by a redox
Thermocalc software (Holland & Powell, 1998) version 3.26. At reaction between Fe and C in the subducted dolomite.
each pressure investigated, the observed stability limit of graphite Graphite–diamond transition is from Sung (2000).
was reported in order to delimit (black dashed curve) the upper
stability of graphite. Graphite – CO2 equilibrium (grey curves) was
positioned using the experimental data of Frost (1998). Note that the
siderite breakdown reaction should form CO2 instead of C þ O2 after therefore, the interaction between Mo and C equilibria was
it intersects CCO (thus, the higher temperature part of the carbonate probably not complete. The remaining dolomite and coe-
breakdown curve is metastable).
site were chemically ‘‘isolated’’ by a mechanism similar to
that observed in self-buffered experiments – i.e. growth of
minerals, in spite of the presence of a CO2 fluid. Our cpx at grain boundaries – as shown by the large number of
samples exhibit identical clinopyroxene compositions dolomite and coesite inclusions in cpx (Fig. 3b). Therefore,
within each sample; thus, we consider that a temporary we expect that reaction (7) occurs at much lower tempera-
chemical equilibrium is attained. The determination of the ture than the lower experimental temperatures we tested.
precise reaction temperatures would require either very This strengthens our conclusion that the assemblage dolo-
long duration experiments (months, years?), or reversal mite þ coesite may be destabilized at very shallow depth
experiments (which, in our case, were limited by the melt- during subduction due to interactions with the mantle.
ing temperature of the capsule metal). Therefore, our dec-
arbonation temperatures are maximum temperatures, and
the decrease of the decarbonation temperature caused by 5. Implications for dolomite þ coesite high-
iron represents a minimum. This implies that the stability of pressure stability and a potential way of
dolomite þ coesite during subduction is much lower than diamond formation during subduction
previously determined (Luth, 1995) if the presence of iron
is considered. Kinetics of the low-temperature graphite Our results show that the destabilization of dolomite þ
formation reaction are unknown but might also be slow coesite is favoured by the presence of iron and by reducing
enough that a temporary chemical equilibrium was conditions. If this assemblage is present in the lower part of
reached. Therefore, the fO2 calculated for these experi- the subducted crust, i.e. far enough from the mantle wedge
ments represent maximum values. to prevent redox interactions, and if there is no interaction
with other minerals or fluids (which means, at an intrinsic
fO2), CO2 is produced from less than 6.5 GPa (195 km) to
4.3.2. Reducing experiments 7.5 GPa (225 km) and melt above 7.5 GPa (Fig. 8). At
In Series 2 (reducing conditions), molybdenum oxides low pressure and temperature, graphite forms by the break-
formed between the Mo0 capsule and samples in all experi- down of the carbonate ferrous end-member up to 4.5
ments that used the reducing buffer. Oxide stoichiometry GPa. In contact with the peridotitic mantle wedge, slab
deduced from microanalyses indicates a MoO2 (IV) com- delamination and chemical diffusion probably produce
position. No other molybdenum oxide formed. No iron was more reducing conditions for the subducted material. In
detected either in the Mo oxide, or in the remaining Mo this case, dolomite is reduced to graphite above 3.5 GPa
capsule. Graphite is observed in most samples, but carbo- (90 km). Diamond is absent in our samples, although
nate is also still present. Thermodynamic calculations most of the experiments were performed in the diamond
(Fig. 1) show that the reaction Mo þ O2 ¼ MoO2 stability field. This feature is common in experiments
(MMO) lies 2–4 log units fO2 below the reaction CO2 ¼ performed at moderate pressure (Gunn & Luth, 2006)
C þ O2 (CCO) at the P–T conditions of our experiments; and is related to the high energy required for diamond
14 A.M. Martin, T. Hammouda

Fig. 9. Sketch view of the subducting slab illustrating the possibilities for dolomite þ coesite reaction. Case (1) represents material remaining
in the slab, whereas case (2) corresponds to material assimilated to the mantle (delamination). Diamond symbols represent graphite or
diamond formation (a) by reduction at the mantle interface and (b) by a Fe/C redox reaction inside the dolomite. Grey areas represent portions
of the slab that would not be affected by the reducing conditions of the mantle, i.e. far enough to prevent oxygen diffusion. Details of the
various reactions are given in the text.

nucleation. In nature, one can anticipate the role of seeds, Alt, J.C. & Teagle, D.A.H. (1999): The uptake of carbon during altera-
nuclei or high strain rates at slab/mantle interface (King & tion of ocean crust. Geochim. Cosmochim. Acta, 63, 1527–1535.
Bebout, 2006) to promote diamond nucleation. Therefore, Anovitz, L.M. & Essene, E.J. (1987): Phase equilibria in the system
our results illustrate a potential mechanism for diamond CaCO3 – MgCO3 – FeCO3. J. Petrol., 28, 389–414.
formation by the reduction of dolomite þ coesite inside the Ballhaus, C. (1993): Redox states of lithospheric and asthenospheric
subducted plate, far from or in contact with the mantle upper mantle. Contrib. Mineral. Petrol., 114, 331–348.
(Fig. 9). Chemical diffusion – in particular, of iron and Beyssac, O., Goffé, B., Petitet, J.-P., Froigneux, E., Moreau, M.,
Rouzaud, J.-N. (2003): On the characterization of disordered and
oxygen – at slab/mantle interface and in the slab appears to
heterogeneous carbonaceous materials by Raman spectroscopy.
be a key parameter.
Spectrochim. Acta Pt. A, 59, 2267–2276.
Buob, A., Luth, R.W., Schmidt, M.W., Ulmer, P. (2006):
Experiments on CaCO3-MgCO3 solid solutions at high pressure
Acknowledgements: D. Laporte and D. Andrault are sin- and temperature. Am. Mineral., 91, 435–440.
cerely acknowledged for their comments. The authors are Canil, D. & O’Neill, H.St.C. (1996): Distribution of ferric iron in
also grateful to K. Righter for English corrections and some upper-mantle assemblages. J. Petrol., 37, 609–635.
interesting comments. Dolomite starting material was Cesare, B. (1995): Graphite precipitation in C-O-H fluid inclu-
kindly provided by B. Devouard. Assistance by J.-L. sions: closed system compositional and density changes, and
Devidal on electron microprobe, by J.-M. Hénot and F. thermobarometric implications. Contrib. Mineral. Petrol.,
Faure on scanning electron microscope, by J.-M. Nédélec 122, 25–33.
on Raman microprobe and by N. Bolfan-Casanova, F. Davidson, P.M. (1994): Ternary iron, magnesium, calcium carbo-
Pointud and J.-L. Fruquière on multi-anvil apparatus is nates: A thermodynamic model for dolomite as an ordered deri-
gratefully acknowledged. We also thank R.W. Luth and vative of calcite-structure solutions. Am. Mineral., 79, 332–339.
G.M. Yaxley for their constructive reviews, and B. Fritz, Eggler, D.H. & Baker, D.R. (1982): Reduced volatiles in the system
C. Shaw and C. Chopin for their editorial work. Financial C-O-H: implications to mantle melting, fluid formation, and
support from CNRS-INSU (DyETI program) is also diamond genesis. in ‘‘High-Pressure Research in Geophysics’’,
acknowledged. The multi-anvil apparatus of Laboratoire S. Akimoto & M.H. Manghnani, eds. Advances in Earth and
Magmas et Volcans is financially supported by the Centre Planetary Sciences, 12, Tokyo, 237–250.
National de la Recherche Scientifique (Instrument Frost, D.J. (1998): Erratum to Frost & Wood (1997) ‘‘Experimental
National de l’INSU). measurements of the fugacity of CO2 and graphite/diamond
stability from 35 to 77 kbar at 925 to 1650 C’’. Geochim.
Cosmochim. Acta, 62, 725.
Frost, D.J. & McCammon, C.A. (2008): The redox state of Earth’s
mantle. Annu. Rev. Earth Planet. Sci., 36, 389–420.
References Frost, D.J. & Wood, B.J. (1997): Experimental measurements of the
fugacity of CO2 and graphite/diamond stability from 35 to 77 kbar
Akaogi, M. (1995): Thermodynamic properties of a-quartz, coesite, at 925 to 1650 C. Geochim. Cosmochim. Acta, 61, 1565–1574.
and stishovite and equilibrium phase relations at high pressure Goldsmith, J.R., Witters, J., Northrup, D.A. (1962): Studies in the
and high temperatures. J. Geophys. Res., 100, 22337–22347. system CaCO3-MgCO3-FeCO3: A method for major-element
Role of iron and low fO2 on dolomite þ coesite stability 15

spectrochemical analysis; 3. Composition of some ferroan dolo- Marty, B. & Tolstikhin, I.N. (1998): CO2 fluxes from mid-ocean
mites. J. Geol., 70, 659–687. ridges, arcs and plumes. Chem. Geol., 145, 233–248.
Gunn, S.G. & Luth, R.W. (2006): Carbonate reduction by Fe-S-O McCammon, C.A., Griffin, W.L., Shee, S.R., O’Neill, H.S.C. (2001):
melts at high pressure and high temperature. Am. Mineral., 91, Oxidation during metasomatism in ultramafic xenoliths from the
1110–1116. Wesselton kimberlite, South Africa: implications for the survival
Hammouda, T. (2003): High pressure melting of carbonated eclogite of diamond. Contrib. Mineral. Petrol., 141, 287–296.
and experimental constraints on carbon recycling and storage in Moine, B.N., Grégoire, M., O’Reilly, S.Y., Delpech, G., Sheppard,
the mantle. Earth Planet. Sci. Lett., 214, 357–368. S.M.F., Lorand, J.P., Renac, C., Giret, A., Cottin, J.-Y. (2004):
Hammouda, T., Andrault D., Koga, K., Katsura, T., Martin, A.M. Carbonatite melt in oceanic upper mantle beneath the Kerguelen
(2010): Ordering in double carbonates and implications for Archipelago. Lithos, 75, 239–252.
processes at subduction zones. Contrib. Mineral. Petrol., DOI: Morlidge, M., Pawley, A., Droop, G. (2006): Double carbonate
10.1007/s00410-010-0541-z. breakdown reactions at high pressures: an experimental study
Holland, T.J.B. & Powell, R. (1998): An internally consistent ther- in the system CaO–MgO–FeO–MnO–CO2. Contrib. Mineral.
modynamic data set for phases of petrological interest. J. Petrol., 152, 365–373.
Metamorphic Geol., 16, 309343. Nakamura, K. & Kato, Y. (2004): Carbonatization of oceanic crust
Ionov, D.A., Dupuy, C., O’Reilly, S.Y., Kopylova, M.G., Genshaft, by the seafloor hydrothermal activity and its significance as a
Y.S. (1993): Carbonated peridotite xenoliths from Spitsbergen: CO2 sink in the Early Archean. Geochim. Cosmochim. Acta, 68,
implications for trace element signature of mantle carbonate 4595–4618.
metasomatism. Earth Planet. Sci. Lett., 119, 283–297. Nixon, P.H. (1995): A review of mantle xenoliths and their role in
Katsura, T. & Ito, E. (1989): The system Mg2SiO4-Fe2SiO4 at high diamond exploration. J. Geodyn., 20, 305–329.
pressures and temperatures: precise determination of stabilities O’Neill, H.St.C (1986): Mo-MoO2 (MOM) oxygen buffer and the
of olivine, modified spinel, and spinel. J. Geophys. Res., 94, free energy of formation of MoO2. Am. Mineral., 71,
15663–15670. 1007–1010.
Klein-Ben David, O., Izraeli, E.S., Hauri, E., Navon, O. (2007): Parkinson, I.J. & Arculus, R.J. (1999): The redox state of subduction
Fluid inclusions in diamonds from the Diavik mine, Canada zones: insights from arc-peridotites. Chem. Geol., 160, 409–423.
and the evolution of diamond-forming fluids. Geochim. Plank, T. & Langmuir, C.H. (1998): The chemical composition of
Cosmochim. Acta, 71, 723–744. subducting sediment and its consequences for the crust and
King, R.L. & Bebout, G.E. (2006): Metamorphic evolution along the mantle. Chem. Geol., 145, 325–394.
slab/mantle interface within subduction zones. Geochim. Robie, R.A. & Hemingway, B.S. (1995): Thermodynamic properties
Cosmochim. Acta, 70, 18 (Supplement 1), A319. of Minerals and Related Substances at 298,15K and 1Bar (105
Laverne, C. (1993): Occurrence of siderite and ankerite in young Pascals) Pressure and at Higher Temperatures. U.S. Geol. Surv.
basalts from the Galapagos Spreading Center (DSDP Holes Bull., 2131.
506G and 507B). Chem. Geol., 106, 27–46. Schiano, P., Clocchiatti, R., Shimizu, N., Weis, D., Mattielli, N.
Luque, F.J., Pasteris, J.D., Wopenka, B., Rodas, M., Barrenechea, (1994): Cogenetic silica-rich melts trapped in mantle minerals in
J.F. (1998): Natural fluid-deposited graphite: mineralogical Kerguelen ultramafic xenoliths: implications for metasomatism in
characteristics and mechanisms of formation. Am. J. Sci., 298, the oceanic upper mantle. Earth Planet. Sci. Lett., 123, 167–178.
471–498. Schrauder, M. & Navon, O. (1993): Solid carbon dioxide in a natural
Luth, R.W. (1993): Diamonds, eclogites, and oxidation state of the diamond. Nature, 365, 42–44.
Earth’s mantle. Science, 261, 66–68. Staudigel, H.R., Hart, S.R., Schmincke, H.-U., Smith, B.M. (1989):
— (1995): Experimental determination of the reaction dolomite þ Cretaceous ocean crust at DSDP Sites 417 and 418: Carbon
2coesite ¼ diopside þ 2CO2 to 6 GPa. Contrib. Mineral. Petrol., uptake from weathering versus loss by magmatic outgassing.
122, 152–158. Geochim. Cosmochim. Acta, 53, 3091–3094.
— (1999): Carbon and carbonates in the mantle. in: ‘‘Mantle Petrology: Sung, J. (2000): Graphite-diamond transition under high pressure: a
Field Observations and High Pressure Experimentation. A Tribute kinetics approach. J. Mater. Sci., 35, 6041–6054.
to Francis (Joe) Boyd’’, Y. Fei, C.M. Bertka, B.O. Mysen, eds. Viljoen, K.S. (1995): Graphite- and diamond-bearing eclogite xeno-
Geochem. Soc. Spec. Publ., 6, 297–316. liths from the Bellsbank kimberlite, Northern Cape, South
— (2001): Experimental determination of the reaction aragonite þ Africa. Contrib. Mineral. Petrol., 121, 414–423.
magnesite ¼ dolomite at 5 to 9 GPa. Contrib. Mineral. Petrol., Wang, A., Pasteris, J.D., Meyer, H.O.A., Dele-Duboi, M.L. (1996):
141, 222–232. Magnesite-bearing inclusion assemblage in natural diamond.
— (2006): Experimental study of the CaMgSi2O6–CO2 system at Earth Planet. Sci. Lett., 141, 293–306.
3–8 GPa. Contrib. Mineral. Petrol., 151, 141–157. Wang, J., Hattori, K.H., Kilian, R., Stern, C.R. (2007):
Lüttge, A. & Metz, P. (1991): Mechanism and kinetics of the reac- Metasomatism of sub-arc mantle peridotites below southern-
tion 1 Dolomite þ 2 Quartz ¼ 1 Diopside þ 2 CO2. Can. most South America: reduction of fO2 by slab-melt. Contrib.
Mineral., 29, 803–821. Mineral. Petrol., 153, 607–624.
—, — (1993): Mechanism and kinetics of the reaction 1 Dolomite þ Wopenka, B. & Pasteris, J.D. (1993): Structural characterization
2 Quartz ¼ 1 Diopside þ 2 CO2: a comparison of rock-sample of kerogens to granulite-facies graphite: Applicability of
and powder experiments. Contrib. Mineral. Petrol., 115, Raman microprobe spectroscopy. Am. Mineral., 78,
155–164. 533–557.
16 A.M. Martin, T. Hammouda

Yaxley, G.M. & Green, D.H. (2004): Phase relations of carbonate- Zhang, Y. & Zindler, A. (1993): Distribution and evolution of
bearing eclogite assemblages from 2.5 to 5.5 GPa: implications carbon and nitrogen in Earth. Earth Planet. Sci. Lett., 117,
for petrogenesis of carbonatites. Contrib. Mineral. Petrol., 146, 331–345.
606–619.
Yaxley, G.M., Green, D.H., Kamenetsky, V. (1998): Carbonatite Received 10 March 2010
metasomatism in the Southeastern Australian lithosphere. J. Modified version received 28 June 2010
Petrol., 39, 1917–1930. Accepted 22 September 2010

You might also like