You are on page 1of 21

Lithos 220–223 (2015) 200–220

Contents lists available at ScienceDirect

Lithos
journal homepage: www.elsevier.com/locate/lithos

Invited review article

Diamond formation — Where, when and how?


T. Stachel ⁎, R.W. Luth
Department of Earth and Atmospheric Sciences, University of Alberta, Edmonton, AB, T6G 2E3, Canada

a r t i c l e i n f o a b s t r a c t

Article history: Geothermobarometric calculations for a worldwide database of inclusions in diamond indicate that formation of the
Received 25 August 2014 dominant harzburgitic diamond association occurred predominantly (90%) under subsolidus conditions. Diamonds
Accepted 29 January 2015 in eclogitic and lherzolitic lithologies grew in the presence of a melt, unless their formation is related to strongly
Available online 14 February 2015
reducing CHO fluids that would increase the solidus temperature or occurred at pressure–temperature conditions
below about 5 GPa and 1050 °C. Three quarters of peridotitic garnet inclusions in diamond classify as “depleted”
Keywords:
Inclusion in diamond
due to their low Y and Zr contents but, based on LREEN–HREEN ratios invariably near or greater than one, they nev-
Diamond formation ertheless reflect re-enrichment through either highly fractionated fluids or small amounts of melt. The trace element
Solidus signatures of harzburgitic and lherzolitic garnet inclusions are broadly consistent with formation under subsolidus
Redox reaction and supersolidus conditions, respectively. Diamond formation may be followed by cooling in the range
Peridodite of ~ 60–180 °C as a consequence of slow thermal relaxation or, in the case of the Kimberley area in South
Eclogite Africa, possibly uplift due to extension in the lithospheric mantle. In other cases, diamond formation and
final residence took place at comparable temperatures or even associated with small temperature increases
over time.
Diamond formation in peridotitic substrates can only occur at conditions at least as reducing as the EMOD buffer.
Evaluation of the redox state of 225 garnet peridotite xenoliths from cratons worldwide indicates that the vast
majority of samples deriving from within the diamond stability field represent fO2 conditions below EMOD.
Modeling reveals that less than 50 ppm fluid are required to completely reset the redox state of depleted cratonic
peridotite to that of the fluid. Consequently, the overall reduced state of diamond stable peridotites implies that
the last fluids to interact with the deep cratonic lithosphere were generally reducing in character. A further
consequence of the extremely limited redox buffering capacity of cratonic peridotites is that redox reactions
with infiltrating fluid/melt likely cannot produce large diamonds or high diamond grades. Evaluating the shift
in maximum carbon content in CHO fluids during either isobaric cooling or ascent along a cratonic geotherm,
however, reveals that isochemical precipitation of carbon from CHO fluids provides an efficient mode of diamond
crystallization. Since subsolidus fluids are permissible in harzburgites only, and supersolidus melts in lherzolite
we suggest that CHO fluid metasomatism may explain the long observed close association between diamonds
and harzburgitic garnets. In the absence of thermodynamic data we cannot evaluate if supersolidus carbonate-
bearing melts, stable at fO2 conditions below EMOD, would experience a similar decrease in maximum carbon
solubility during cooling or ascent along a geotherm. The absence of a clear association between diamond and
lherzolitic garnets, however, suggests that this is not the case. A very strong association between diamond and
eclogite likely relates to the fact that the transition from carbonate to diamond stable conditions occurs at
redox conditions that are at least about 1 log unit more oxidizing than EMOD. At this time we cannot quantita-
tively evaluate the redox buffering capacity of cratonic eclogites but given their much higher Fe contents it has
to be significantly higher than for peridotites. Alternatively, diamond in eclogite may precipitate directly from
cooling carbonate-bearing melts that may be too oxidizing to crystallize diamond in olivine-bearing lithologies.
© 2015 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
2. Diamond substrates in Earth's mantle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201

⁎ Corresponding author.
E-mail address: tstachel@ualberta.ca (T. Stachel).

http://dx.doi.org/10.1016/j.lithos.2015.01.028
0024-4937/© 2015 Elsevier B.V. All rights reserved.
T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220 201

3. Pressure–temperature conditions of diamond formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201


3.1. Peridotitic suite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
3.2. Eclogitic suite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
3.3. Comparison of temperatures derived from non-touching and touching inclusion pairs: evidence for diamond formation during transient
heating events? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
3.3.1. Peridotitic inclusion pairs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
3.3.2. Eclogitic inclusion pairs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
3.4. Pressure–temperature conditions of diamond formation and solidi of diamond host rocks . . . . . . . . . . . . . . . . . . . . . . . 204
3.5. Evidence from trace elements in support of melt-present and melt-absent diamond formation . . . . . . . . . . . . . . . . . . . . . 205
4. Timing of diamond growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
5. Diamond forming reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
5.1. Diamond formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
5.2. Oxidized and reduced mineral inclusions in diamond . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
5.3. Oxidation state of Earth's upper mantle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
5.4. Fluids in the upper mantle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
5.5. Ascent along lithospheric PT–fO2 path . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
5.6. “Isochemical” ascent or isobaric cooling of fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
5.7. Diamond precipitation from ascending or isobarically cooling melts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
5.8. The redox buffering capacity of cratonic peridotite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
6. Diamond forming processes based on co-variations in δ13C–N and on fluid inclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
7. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217

1. Introduction studies have shown that diamonds derive from subcontinental


lithospheric mantle extending into the diamond stability field
Following the discovery of the first kimberlite hosted diamond (e.g., Boyd and Gurney, 1986) or may have originated at an even greater
deposits in South Africa in 1870–1871, diamond formation was initially depth (extending to at least 700 km) in the sublithospheric mantle (Harte
linked to reaction of the kimberlite magma with abundant carbonaceous and Harris, 1994; Moore and Gurney, 1985; Scott Smith et al., 1984).
shale fragments (crustal xenoliths) present in these diatremes (Lewis, Sublithospheric diamonds are, however, rare and the subcratonic litho-
1887). The subsequent proposal of diamond representing a high- spheric mantle represents the primary source of over 99% (by mass) of
pressure phenocryst in kimberlite was widely accepted (e.g., Williams, the worldwide diamond production (Stachel and Harris, 2008). In addi-
1932) till the advent of geochemical studies of inclusions in diamond tion, the sublithospheric diamonds studied so far relate to the recycling
(Meyer, 1968; Meyer and Boyd, 1972; Sobolev et al., 1969) and radio- of oceanic lithosphere into the deep mantle (Harte et al., 1999b; Stachel
metric dating of diamond formation ages (Kramers, 1979; Richardson et al., 2000a, 2000b; Tappert et al., 2005; Walter et al., 2011) and hence
et al., 1984), both implying crystallization in Earth's mantle unrelated can provide only very limited insights into diamond formation and
to host kimberlite magmatism. The seminal suggestion of a xenocrystic storage in pyrolitic upper and lower mantle. For this contribution we
origin for diamond in kimberlite (based on the observation of diamond- will, therefore, focus exclusively on lithospheric diamonds.
iferous eclogite xenoliths), however, already dates back to Bonney Based on their mineral inclusion content, diamonds from the litho-
(1899). Since the 1970s, numerous studies covering all important de- spheric mantle (N = 2837) are divided into peridotitic (65%), eclogitic
posits around the globe have provided a detailed picture on the miner- (33%) and websteritic (pyroxenitic) suites (2%). Using garnet composi-
alogical and chemical environment for diamond formation, including tions (N = 685), the peridotitic inclusion suite can be subdivided into
the attendant pressure–temperature conditions (reviewed in Gurney, harzburgitic (56% of all diamonds), lherzolitic (8%) and wehrlitic
1989; Meyer, 1987; Stachel and Harris, 2008). Following the recognition (0.7%) parageneses. The 86:13:1 harzburgite:lherzolite:wehrlite split
of diamond as mantle-derived xenocrysts in kimberlite, the concept of the peridotitic suite is virtually unchanged from the original
emerged that diamond forms via redox reactions (Eggler and Baker, pioneering work of Gurney (1984). The 2:1 ratio of peridotitic:eclogitic
1982; Luth, 1993; Rosenhauer et al., 1977) that relate to migration of a suite diamonds is based on destructive studies on diamonds generally
fluid or melt through a mantle host rock (Haggerty, 1986) and in conse- b3 mm in size; it has, however, been speculated that among larger
quence, diamond is now generally viewed as a metasomatic mineral diamonds the relative proportion of the eclogitic suite may increase
(e.g., Stachel and Harris, 1997; Taylor et al., 1998). Little, however, is (e.g., Gurney, 1989; Stachel and Harris, 2008). In any case, 33% of all di-
known about the exact composition and the redox character amonds hosted in eclogite far exceed the b 1 to 5% estimated volumetric
(carbonate- versus methane-bearing) of the fluids or melts that precip- abundance of eclogite in subcratonic lithospheric mantle (Dawson and
itate smooth-surfaced monocrystalline diamonds. Stephens, 1975; McLean et al., 2007; Schulze, 1989). Equally, the ratio
In this contribution we use the large body of published data on of harzburgitic:lherzolitic paragenesis diamonds (~ 7:1) reverses the
diamonds and their mineral inclusions and less plentiful Fe3 +/Fe2 + relative proportions of harzburgite to lherzolite in diamond stable lith-
determinations on minerals in cratonic garnet peridotites to discuss ospheric mantle (ca. 1:4 for the Western Kaapvaal Craton; Griffin
the “where, when and how?” of diamond formation and to place et al., 2003). This suggests that compared to lherzolite, harzburgite
constraints on possible modes of diamond precipitation that invalidate and eclogite are strongly preferred substrates for diamond (Grütter
some popular models. et al., 2004; Gurney, 1984).

2. Diamond substrates in Earth's mantle 3. Pressure–temperature conditions of diamond formation

The mineralogy and the mineral compositions of diamond host rocks 3.1. Peridotitic suite
in Earth's mantle are very well characterized through studies on mineral
inclusions in diamond (reviewed in Gurney, 1989; Meyer, 1987; Meyer Diamond represents a closed system, with even the mobility of hydro-
and Boyd, 1972; Shirey et al., 2013; Stachel and Harris, 2008). These gen being very low (Connell et al., 1998; Saguy, 2004). Non-touching
202 T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220

inclusions (i.e., isolated single-phase mineral inclusions) in diamond are, diamond transition (Fig. 1a). The regression line indicates a typical
therefore, assumed to have remained unchanged since encapsulation — diamond window – representing the thickness of diamond stable
rare cases with microscopically visible exsolution features aside (Leost lithospheric mantle bracketed by the intersections of the local
et al., 2003). Disequilibrium has been documented among non-touching geotherm with the graphite–diamond transition and the mantle
inclusions incorporated during diamond growth in chemically evolving adiabat – of 95 km thickness (110–205 km depth). Linear regression
environments (e.g., Bulanova, 1995; Griffin et al., 1988; Rickard et al., of the PNT 00–TNT 00 dataset of Grütter (2009) for garnet lherzolite
1989; Stachel et al., 1998). Multiple inclusions in single diamonds suitable xenoliths from the Central Slave Craton yields an almost identical
for the application of independent geothermometers allow testing for local geotherm, underscoring that the Central Slave Craton represents
possible disequilibrium. Worldwide, 13 diamonds have been studied a very appropriate choice as a reference geotherm for cratonic regions
containing inclusions of garnet + orthopyroxene + olivine (database of with high diamond potential (Grütter, 2009).
Stachel and Harris, 2008), allowing to derive temperature estimates
based on the Mg–Fe exchange between both garnet–olivine (O'Neill,
1980; O'Neill and Wood, 1979) and garnet–orthopyroxene (Harley,
1984); accounting for the systematic deviations between the two ther-
mometers documented by Brey and Köhler (1990), the two methods
yield temperature estimates agreeing within uncertainty for 12 of the
diamonds (i.e., N 90% show no evidence for internal disequilibrium; T [˚C]
Supplementary Fig. 1). Adding further evidence derived through detailed 600 800 1000 1200 1400
3.0
examination of tie-lines connecting coexisting mineral pairs (e.g., Otter Gr

Lherz+CH4+H2O
ap
and Gurney, 1989; Rickard et al., 1989) and observation of common hit
e

Solidus
Dia Harzburgitic
trace element equilibrium between multiple inclusions (e.g., garnet– mo
4.0 nd
clinopyroxene pairs; Stachel et al., 2000a), disequilibrium between non-
touching inclusions is considered the exception rather than the rule
(Gurney, 1989). On this basis, the pressure–temperature conditions of
Lherzolitic

P [GPa]
diamond formation can be derived through mineral exchange 5.0

Harzb+Carb+H2O
geothermobarometry employing silicate inclusions. In its primary sta- b

Solidus
bility field and at temperatures ≥1000 °C, diamond is not a rigid pres-
sure vessel but will transfer changes in ambient pressure to included 6.0
minerals by means of plastic deformation (De Vries, 1975). Touching Solidus
minerals will thus re-equilibrate under changing pressure–temperature Lherz+Carb+H2O

conditions and, therefore, reflect conditions of the final mantle storage. Solidus
7.0
The geothermobarometers applicable to a comparatively large num- Lherz+H2O
ber of inclusions are: (1.) for garnet–orthopyroxene pairs, the Al 3.0
Gr

TP=1300°C
exchange barometer of Brey and Köhler (1990; PBKN) combined with a

Adiabat
Dia phite
the Mg–Fe exchange thermometer of Harley (1984; THarley) and (2.) for mo
nd
clinopyroxene inclusions assumed to be in equilibrium with 4.0
orthopyroxene and garnet, the single crystal geothermobarometer of
Nimis and Taylor (2000; PNT 00 and TNT 00). The former is applicable to
the dominant harzburgitic and the lherzolitic inclusion paragenesis; 45
P [GPa]

5.0 a
the latter can only be applied to comparatively rare diamonds with
lherzolitic inclusions. The combination of PBKN–THarley owes its relatively
large uncertainty (1 sigma of ~100 °C and 0.5 GPa) mainly to systematic PNT 00-PNT 00
errors in the thermometer calibration (above 1000 °C the thermometer 6.0
Lherzolitic 40
increasingly underestimates temperature; Brey and Köhler, 1990). The PBKN-THarley
experimental calibration of the PBKN barometer extends to 6.0 GPa; Lherzolitic
extrapolation to higher pressures adds additional uncertainty. The better 7.0 Harzburgitic
precision of PNT 00–TNT 00 (~50 °C and 0.3 GPa) only holds to pressure of
35

up to ~ 4.5 GPa; at higher pressure the barometer increasingly underesti- 600 800 1000 1200 1400
mates pressure (Nimis, 2002). Application of PNT 00–TNT 00 requires rigor- T [˚C]
ous filtering of clinopyroxene analyses (using cation totals as a quality
criterion and then excluding all compositions outside the experimental
ranges; see Grütter, 2009 for details), which for the current study elimi- Fig. 1. (a): Pressure–temperature estimates for peridotitic suite diamonds based on garnet–
nated 40% of the available data (N = 134). orthopyroxene inclusion pairs (PBKN–THarley) and clinopyroxene inclusions (PNT 00 and
TNT 00). The latter combination is only applicable to the lherzolitic inclusion paragenesis.
Fig. 1 illustrates P–T estimates (including 26% touching inclusion
The filled red circle and surrounding pale red error ellipse represent the average and 1 stan-
pairs) using the two geothermobarometer combinations discussed dard deviation for the entire dataset (5.3 ± 0.8 GPa and 1130 ± 120 °C; N = 157). A linear
above. Based on 157 independent estimates (82 diamonds with regression through the dataset (r2 = 0.53) is shown as a red line and represents an average
lherzolitic and 75 with harzburgitic inclusions), the average peridotitic geotherm for global diamondiferous lithospheric mantle. Continental geotherms (blue
suite diamond derives from 5.3 ± 0.8 GPa and 1130 ± 120 °C. dashed lines with surface heat flow values in mW/m2; Hasterok and Chapman, 2011), a
mantle adiabat for a potential temperature of 1300 °C (Hasterok and Chapman, 2011), and
Lherzolitic and harzburgitic inclusions yield similar means (5.1 ± the graphite–diamond transition (Day, 2012) are shown for reference. (b): Fields of P–T
0.8 GPa and 1120 ± 140 °C, and 5.6 ± 0.8 GPa and 1140 ± 90 °C, respec- conditions for lherzolitic and harzburgitic inclusions in diamonds compared to a range of
tively). The lherzolitic data, however, extend to low temperature condi- solidus temperatures in the presence of CHO-fluids. The solidus for hydrous melting of
tions (640–970 °C) that are not observed for the harzburgitic inclusion carbonated lherzolite (red short dashed line) is from Wyllie and Ryabchikov (2000). The
solidi of lherzolite in the presence of a hydrous fluid (red long dashed line) and a reduced
set. A regression line through the combined dataset is slightly oblique
CHO fluid (red dotted line) are from Wyllie and Ryabchikov (2000) and Litasov et al.
to the continental heat flow geotherms of Hasterok and Chapman (2014), respectively. The solidus of harzburgite in the presence of carbonate and H2O (blue
(2011), crossing from a 39 mW/m2 model geotherm at the intersection dashed-dotted line) is taken from Wyllie (1987). The error ellipse around the average
with the mantle adiabat to a 36 mW/m2 geotherm at the graphite value of P–T estimates for peridotitic suite diamonds is shown as a black dashed line.
T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220 203

The significance of the relatively small difference in average P–T con- entire 900–1400 °C experimental range. Purwin et al. (2013) con-
ditions for peridotitic suite diamonds of lherzolitic and harzburgitic para- ducted eclogite system experiments specifically designed to evalu-
genesis is difficult to evaluate as it likely results from iterations in two ate the effect of highly variable Fe3 +/Fe2 + in their starting materials
different geothermobarometer combinations (e.g., Nimis and Grütter, over an 800–1300 °C temperature range. They found that, if total iron
2010). Assuming an ~5.0 GPa average pressure, the temperature condi- is used as input for TKrogh 88, their experiments are reproduced over
tions for diamonds hosted by lherzolite and harzburgite can be the entire temperature range within 60 °C, whilst significant systematic
compared based only on the garnet–olivine Mg–Fe exchange temperature estimation errors occur when measured Fe2+ is employed.
geothermometer (O'Neill, 1980; O'Neill and Wood, 1979; T O'Neill). Consequently, results obtained for eclogitic inclusions in diamond using
Based on TO'Neill (Fig. 2a), harzburgitic (N = 144, average = 1170 ± total Fe measurements via EPMA and the Krogh (1988) thermometer
120 °C, median = 1170 °C) and lherzolitic (N = 23, average = can be considered both accurate and precise.
1150 ± 170 °C, median = 1170 °C) inclusions yield the same median A reliable barometer suitable for mantle eclogites and eclogitic inclu-
values and an equality of means cannot be rejected (Student's t-test at sions in diamond is currently not available (Nimis and Grütter, 2010).
α = 5%). This suggests that harzburgitic and lherzolitic paragenesis Assuming that diamonds hosted by eclogite form at similar pressure as
diamonds grew in different substrates but under comparable P–T peridotitic suite diamonds, a fixed value of 5.0 GPa is applied for thermo-
conditions. metric calculations.
On this basis, 144 eclogitic garnet–clinopyroxene pairs (including
11% touching inclusion pairs) yield an average equilibration tempera-
3.2. Eclogitic suite ture of 1170 ± 110 °C (median = 1160 °C; see Fig. 2b). This compares
very well to temperatures derived from equally abundant (N = 164)
The most widely used and tested thermometer for mantle eclogites is garnet–olivine pairs (TO'Neill) in peridotitic suite diamonds, which
the garnet–clinopyroxene Mg–Fe exchange thermometer of Krogh yield a mean of 1160 ± 110 °C (median = 1140 °C; see Fig. 2a) for
(1988; TKrogh 88). Brey and Köhler (1990) found that TKrogh 88 reproduces the same assumed pressure. This excellent agreement suggests that
their lherzolite system experiments within about 40 °C (1 sigma) over the diamonds hosted by peridotite and eclogite indeed share a common
origin within diamond stable lithospheric mantle. In the absence of a
reliable barometer, pressure estimates for mantle eclogites are usually
based on projections of calculated temperatures on independently
TKrogh 88 [°C]
derived local geotherms. Following a similar approach (Fig. 3), the
800 900 1000 1100 1200 1300 1400 average temperature and one sigma range (1170 ± 110 °C) for eclogitic
40
inclusions in diamond are projected on both the average Precambrian
Eclogitic: shield geotherm of Hasterok and Chapman (2011; 40 mW/m2) and
N=144 the peridotitic inclusion-based diamond geotherm (Fig. 1a). A depth of
30 Mean=1170 origin for 70% (samples within 1 sigma) of eclogitic suite diamonds
±110 °C between 135–190 km is obtained for the average shield geotherm, and
Median=1160 °C 145–200 km for the peridotitic inclusion-based diamond “geotherm”
(Fig. 3). The latter projection results in P–T conditions that are about
N 20 b 0.5 GPa and 70 °C higher than those for peridotitic suite diamonds.

10 3.0
Solidus

TP=1300°C
Adiabat
Basalt+H2O Graphite-
Diamond
4.0 Transition
0
Peridotitic
Peridotitic: Diamond
Harzburgitic 45
N=164
P [GPa]

5.0 P-T
30
Mean=1160 ±110 °C Regression
Median=1140 °C
Eclogite+CH4+H2O

6.0
40
Solidus

N 20 a 1 sigma

7.0 Solidus
Eclogite+Carbonate
Lherzolitic
35

10 600 800 1000 1200 1400


T [˚C]

0 Fig. 3. Average temperature and 1 sigma range (1170 ± 110 °C; TKrogh 88 at 5 GPa) for 144
800 900 1000 1100 1200 1300 1400 eclogitic garnet–clinopyroxene inclusion pairs shown as solid red circles and bold red lines,
in a pressure corrected projection onto two geotherms: the 40 mW/m2 average Precambrian
TO’Neill [°C] shield geotherm of Hasterok and Chapman (2011; as labeled) and the peridotitic suite
diamond geotherm at slightly higher pressures (from Fig. 1a). The solidi for hydrous eclogite
Fig. 2. Temperature estimates for diamonds with peridotitic (a; garnet–olivine pairs, TO'Neill) (Kessel et al., 2005; solid green line) and carbonated eclogite (Dasgupta et al., 2004; dashed
and eclogitic inclusions (b; garnet–clinopyroxene pairs, TKrogh 88). Calculations assume a fixed green line) occur at lower temperatures, implying melt present conditions for diamond
pressure of 5.0 GPa. Temperatures above 1400 °C were excluded as they exceed the mantle precipitation in eclogite, except under reducing conditions (IW buffer, with CH4–H2O
adiabat (Fig. 1a) by more than 50 °C. For the eclogitic inclusion set (b), all data from the present; Litasov et al., 2014; long dashed blue line). The hydrous solidus terminates at a
Argyle Diamond Mine were excluded as they represent atypically “hot” conditions (Stachel second critical endpoint (star) between 5–6 GPa (Kessel et al., 2005). Mantle adiabat,
et al., in press). conductive geotherms and graphite-diamond transition as in Fig. 1.
204 T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220

3.3. Comparison of temperatures derived from non-touching and in lithosphere thickness for the Western Kaapvaal Craton of ~20 km oc-
touching inclusion pairs: evidence for diamond formation during curring between kimberlite emplacement at Finsch (124 Ma) and Kim-
transient heating events? berley (94–86 Ma; Griffin et al., 2003) and may suggest that the uplift
reflected in the inclusion in diamond data was related to extension of
3.3.1. Peridotitic inclusion pairs the lithospheric mantle.
An extensive geothermobarometric study to evaluate differences For diamonds from the Panda kimberlite in the Central Slave Craton,
between the conditions of diamond formation (8 non-touching pairs) Canada, thermometric data obtained from non-touching (4 pairs) and
and final mantle residence (27 touching pairs) was conducted on touching (4 pairs) garnet–olivine (TO'Neill) and garnet–orthopyroxene
garnet–orthopyroxene inclusions in diamonds from the De Beers Pool (THarley) pairs indicate that diamond formation was followed by cooling
kimberlite pipes in South Africa (Phillips and Harris, 1995; Phillips (Stachel et al., 2003) by about 150 °C (Fig. 4b). Nitrogen aggregation
et al., 2004). Based on a combination of PBKN and THarley it was found based thermometry (Leahy and Taylor, 1997; assuming 3 Ga mantle
that diamond formation (average non-touching inclusion pairs: residence) indicates that time averaged mantle residence temperatures
6.2 GPa and 1200 °C) was followed by cooling (average touching for Panda diamonds with non-touching pairs (giving diamond forma-
pairs: 5.4 GPa and 1080 °C) by about 120 °C (Phillips et al., 2004). tion temperatures) are on average about 30 °C lower than the mineral
Recalculating the entire dataset of Phillips et al. (2004) at a fixed pressure exchange temperatures but about 120 °C higher for diamonds contain-
of 5.0 GPa, however, strongly reduces the average temperature difference ing touching inclusion pairs (giving final residence temperatures). This
to about 60 °C (Fig. 4a). The decrease of 0.8 GPa in pressure may indicate suggests that slow lithosphere-scale cooling (100 s of million to billion
that the residence history of diamonds in the lithospheric mantle be- year time scale) rather than a short lived local heating event associated
neath Kimberley (Western Kaapvaal) was more complicated than with diamond formation (million year time scale) may have occurred in
simple cooling and involved differential uplift of almost 30 km as the Central Slave Craton.
well (Phillips et al., 2004). This number compares well to the loss Evidence that diamond formation in peridotite is not invariably
followed by cooling can be drawn from two diamonds each containing
a clinopyroxene and a clinopyroxene–orthopyroxene pair (diamonds
MW-80 from Mwadui in Tanzania and Nam-211 from Namibia). Calcu-
T O’Neill or Harley [˚C] lating TNT 00 at 5.0 GPa, in both cases the single clinopyroxene inclusion
900 1000 1100 1200 1300 records temperatures about 40 °C lower than the touching pair. These
4
temperature differences are slightly larger than the error on the temper-
Non- ature estimates (1 sigma of ±30 °C) and suggest that thermal regimes
Panda
touching may also slightly increase following diamond formation.
3
Touching 3.3.2. Eclogitic inclusion pairs
Frequency

For diamonds hosted by eclogite the presence of touching and non-


touching pairs of garnet and clinopyroxene inclusions can only be confi-
2 b dently identified in literature data for the kimberlites at De Beers Pool
(Phillips et al., 2004), George Creek, USA (Chinn, 1995), and Jagersfontein,
South Africa (Tappert et al., 2005). For De Beers Pool and George Creek,
1 touching inclusions record temperatures that are distinctly lower (differ-
ences in average TKrogh 88 of 100 and 180 °C, respectively; Fig. 5a and b)
than obtained from non-touching pairs, suggesting that diamond forma-
tion was followed by cooling. These results are consistent with a
0 ~ 170 °C cooling estimate (recalculated using TKrogh 88) for eclogitic
clinopyroxene and touching garnet–clinopyroxene in an unsourced
De Beers
Touching diamond initially described by Meyer and Tsai (1976), based on analyses
Pool
of Prinz et al. (1975). In the case of Jagersfontein, the temperatures calcu-
15 lated for touching and non-touching pairs agree within error of the ther-
mometer (touching inclusions indicate temperatures that are 20 °C
Frequency

higher; Fig. 5c). Similarly, at Argyle, for a diamond containing one touch-
10 a ing garnet–clinopyroxene pair and two separate garnets, calculated
temperatures (TKrogh 88) agree to within 10 °C (i.e., within error). So
Non- similar to peridotitic suite diamonds, in some cases formation of eclogitic
touching diamonds is followed by 100–180 °C cooling whilst in other cases
5 diamond formation and ultimate storage take place in a thermally stable
mantle environment.

3.4. Pressure–temperature conditions of diamond formation and solidi of


0
900 1000 1100 1200 1300 diamond host rocks
T Harley [˚C]
Based on a layer-by-layer growth mode, smooth-surfaced octahe-
Fig. 4. Temperature estimates for non-touching (blue) and touching (red) peridotitic inclu- dral diamond is presumed to grow into open space (Sunagawa, 1984),
sion pairs in diamonds from the De Beers Pool Mines (a) and the Panda kimberlite of the implying the presence of a melt or a high density fluid. This premise
Ekati Mine (b). Temperatures were calculated assuming a fixed pressure of 5.0 GPa and are can be tested by comparing inclusion-based temperature estimates to
based on garnet–orthopyroxene (THarley; De Beers Pool and Panda) and garnet–olivine experimentally determined solidus temperatures for possible diamond
(TO'Neill; only Panda) inclusion pairs. For De Beers Pool diamonds (a), the average tempera-
ture for non-touching inclusions is 1120 °C versus 1060 °C for touching inclusions (implying
host lithologies.
cooling by 60 °C from diamond formation to final mantle residence). For Panda diamonds (b), Diamond forms as a consequence of the influx of carbon-bearing
the average temperature for non-touching inclusions is 1190 °C versus 1040 °C for touching fluids/melts into pre-existing substrates (e.g., Haggerty, 1986; Stachel
inclusions (150 °C drop in temperature from diamond formation to final mantle residence). and Harris, 1997; Taylor and Green, 1988). Since carbonation reactions
T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220 205

800 900 1000 1100 1200 1300 1400 peridotite plus a H2O–carbonate fluid and carbonated peridotite plus
3
H2O are equivalent. The solidus for hydrous melting of carbonated
Jagersfontein
lherzolite shown in Fig. 1b is taken from Wyllie and Ryabchikov (2000)
Non- and based on extrapolation of experiments b3.5 GPa. However, there is
touching
excellent agreement between this extrapolation and the hydrous solidus
2 of carbonated lherzolite between 4.0–6.0 GPa experimentally determined
Frequency

by Foley et al. (2009). In the pressure region shown in Fig. 1b, the hydrous
solidus of lherzolite (from Wyllie and Ryabchikov, 2000; principally based
c on Kushiro et al., 1968) is only about 50 °C higher. P–T estimates
for lherzolitic inclusions dominantly fall above the hydrous solidi for
1 lherzolite and carbonated lherzolite. At lower pressures (≤5.0 GPa equiv-
Touching
alent to ≤160 km depth) there is, however, a group of “low temperature”
(b1050 to b1000 °C) lherzolitic inclusions that were encapsulated by
diamond under sub-solidus conditions. Because of the low solubility of
CH4 and H2 in silicate melts, reduced fluids inhibit melting (Green,
0 1990). In mixed CH4–H2O fluids, due to strong non-ideality, with increas-
George Creek ing CH4 the activity of H2O decreases more rapidly than the actual H2O
6 content in the fluid, leading to a rapid increase in CH4 activity and solidus
temperature (Foley, 2011). The solidus for lherzolite in the presence of a
reduced CHO fluid shown in Fig. 1b (Litasov et al., 2014) is for very reduc-
Non- ing conditions (IW buffered), resulting in a methane dominated fluid with
Frequency

touching a water content increasing from 13% at 3 GPa to 25% at 6 GPa. Tempera-
4
tures in conductively equilibrated peridotites are below the solidus of
b lherzolite in the presence of a reduced fluid (Fig. 1b). The solidus temper-
Touching
ature for harzburgite in the presence of carbonate and H2O (Fig. 1b;
estimate of Wyllie, 1987) is also high and exceeded by only a few
2
harzburgitic inclusions pairs derived from deeper portions of lithospheric
mantle (N 6.0 GPa equivalent to N 190 km). We conclude that subsolidus
fluids (H2O–CO2− 3 to H2O–CH4) are the principal diamond forming
agent in harzburgitic substrates but only play a similar role in lherzolites
0 if either they are reducing (CH4-rich) or diamond formation occurs
De Beers Pool below the H2O–carbonate solidus (i.e., at b 5 GPa along model geotherms
6 b40 mW/m2; Fig. 1).
Under diamond stable conditions the solidus of anhydrous and non-
Non- carbonated metabasalt distinctly exceeds the mantle adiabat (Yasuda
touching et al., 1994). However, similar to their peridotitic counterparts, the
Frequency

4 formation of diamonds in eclogite is related to the influx of fluids and/


a or melts (Ickert et al., 2013; Keller et al., 1998; Taylor et al., 1998). Con-
trary to peridotite, in the presence of a hydrous fluid and along typical
Touching
cratonic geotherms CO2 is not buffered in bimineralic eclogite (Knoche
2 et al., 1999; Luth, 1993). The experimentally determined solidi of
metabasalt in the presence of water (Kessel et al., 2005) or carbonate
(Dasgupta et al., 2004) provide the two most relevant end-member
conditions (Fig. 3). Kessel et al. (2005) found that the hydrous eclogite
0 solidus terminates at a second critical end-point at 5–6 GPa, after
800 900 1000 1100 1200 1300 1400 which the onset of melting cannot be defined anymore. Experiments
T Krogh 88 [˚C] on melting of eclogite under strongly reducing conditions (Litasov
et al., 2014; IW buffered), i.e. in the presence of a methane dominated
Fig. 5. Temperature estimates for non-touching (blue) and touching (red) eclogitic inclusion
fluid (b 20% H2O in the pressure range of Fig. 3), document an ~200 °C
pairs in diamonds from the De Beers Pool Mines (a), George Creek, USA (b) and Jagersfontein, increase in solidus temperature relative to hydrous and carbonated
South Africa (c). Temperatures were calculated assuming a fixed pressure of 5.0 GPa and are eclogite at 5.0 GPa. Our thermometric results for diamonds in eclogitic
based on garnet–clinopyroxene (TKrogh 88) inclusion pairs. For De Beers Pool diamonds (a), substrates indicate that in the presence of water or carbonate, their
the average temperature for non-touching inclusions is 1170 °C versus 1070 °C for touching
formation occurred in a source that was partially molten (Fig. 3). In the
inclusions (i.e., a 100 °C drop in temperature from diamond growth to final mantle
residence). For diamonds from George Creek (b), the average temperature for non-touching presence of a reducing fluid, eclogite melting would only occur in the
inclusions is 1110 °C and for touching inclusions pairs 930 °C, implying cooling by 180 °C. For deeper portions of subcratonic lithospheric mantle (N 5.5 GPa equivalent
Jagersfontein diamonds (c), the average temperature for non-touching inclusions is 1140 °C to N175 km).
versus 1160 °C for touching inclusions (implying an increase in temperature by 20 °C from
diamond genesis to final mantle residence). Excluding the one touching pair yielding a
very high temperature estimate (1380 °C), cooling by 30 °C is derived. Both temperature
3.5. Evidence from trace elements in support of melt-present and melt-
differences for Jagersfontein are within error of the thermometer. absent diamond formation

Fluid and melt metasomatism have distinctive trace element


prevent the migration of CO2 through olivine-bearing rocks at high signatures, with the ratios of highly to mildly incompatible elements
pressure (Wyllie and Huang, 1976), the spectrum of oxidized to reduced (e.g., LREE/HREE or Zr/Y) decreasing strongly from fluids to melts
fluid species during peridotite melting ranges from H2O–carbonate (e.g., Griffin and Ryan, 1995; Stachel et al., 2004; Stosch and Lugmair,
through pure H2O and H2O–CH4 to H2O–CH4–H2 (Foley, 2011; French, 1986). Griffin and Ryan (1995) empirically devised fields in a diagram
1966; Taylor and Green, 1989). With respect to solidus temperature, of Y versus Zr content in garnet that permit discrimination between
206 T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220

low temperature (fluid) metasomatism and melt metasomatism (Fig. 6a). which may not necessarily be the diamond forming event; e.g., during
The majority (about 75%) of garnet inclusions analyzed for these elements a thermal pulse a harzburgitic diamond substrate may have been affected
fall into the “depleted” field (Fig. 6a), where no clear distinction of meta- by melt metasomatism, but diamond only formed later, after thermal
somatic style can be made below Y b5 ppm and Zr b20 ppm. Plotting the relaxation, during a mild fluid metasomatic event that essentially remains
HREE content (YbN, with N = C1-chondrite normalized) versus the invisible in the trace element signature of inclusions. In fact, Richardson
degree of sinuosity of REEN patterns (NdN/ErN; Fig. 6b), nevertheless et al. (1984) concluded that a melt metasomatic event must have preced-
allows to derive evidence for ubiquitous metasomatic overprint for the ed diamond formation in the lithospheric mantle beneath Kimberley and
strongly depleted group of garnets (as predicted by Frey and Green, Finsch (Western Kaapvaal) by about 300 Ma, based on unsupported high-
1974). This implies that formation of diamond is typically associated ly radiogenic Sr in harzburgitic garnet inclusions.
with mild metasomatic events reflecting low fluid–rock or melt–rock
ratios. For HREE depleted garnets, LREEN–HREEN ratios near or greater 4. Timing of diamond growth
than one imply overprint by a strongly LREE enriched fluid (e.g., Stachel
et al., 2004). For depleted garnets showing some re-enrichment in Cathodoluminescence imaging reveals that diamonds often have
HREE, sinuosity of REE patterns is comparatively low (NdN/ErN b 10; complex internal growth structures, including oscillatory zoning or
Fig. 6b), suggestive of overall mild metasomatism involving a melt. If all the presence of sharp boundaries that may show evidence for intermit-
depleted garnets with NdN/ErN N 20 are interpreted to relate to fluid tent periods of diamond resorption (e.g., Harte et al., 1999a; Wiggers de
metasomatism and all garnets with YbN N 2 to indicate mild melt metaso- Vries et al., 2013a). In situ studies of micron scale variations in nitrogen
matism, then (with one exception) depleted lherzolitic garnets reflect content and carbon isotopic composition demonstrate that such sharp
melt metasomatism whilst depleted harzburgitic garnets reflect both internal boundaries are often associated with abrupt compositional
mild fluid and mild melt metasomatism (Fig. 6b). About half the depleted changes, which may include switches in the redox character (carbonate
garnets (with YbN b 2 and NdN/ErN b 20), although clearly derived from a versus methane bearing) of the precipitating fluid/melt (Palot et al.,
metasomatized source (otherwise NdN/ErN should be ≪1), cannot be 2013; Peats et al., 2012; Wiggers de Vries et al., 2013a). Grouping sul-
assigned to a particular style of overprint. For the more strongly fides from Mir and 23rd Party Congress (Yakutia) based on their compo-
metasomatized inclusions (falling outside the “depleted” field in Fig. 6a) sition, the nitrogen characteristics and carbon isotopic compositions of
it is evident that lherzolitic garnets have high Y/Zr, typical of melt infiltra- the associated growth zones in their host diamonds, and coexisting
tion into their host rocks, whereas harzburgitic garnets predominantly silicate inclusions, Wiggers de Vries et al. (2013b) inferred that these
show low Y/Zr (typical for low-temperature metasomatism) but with a sulfide groupings likely document protracted diamond growth span-
number of exceptions (as predicted from the position of the hydrous ning up to ~1 Ga. In this context, diamond formation ages that are not
solidus; Fig. 1b). Overall, the trace element evidence hence supports the based on the dating of single inclusions have to be viewed with some
conclusions made in Section 3.4. The low-temperature (fluid) metasoma- caution. However, the consistency of certain diamond formation ages
tism evident in Fig. 6a is, however, not a simple extension of the fluid within and across cratons (see below) and, in particular, the observed
metasomatic trend visible in Fig. 6b, since sinuosity of garnet REEN close agreement of Re–Os sulfide and Sm–Nd silicate isochron ages for
patterns actually decreases towards high Zr contents (e.g., all garnets diamonds from individual occurrences (reviewed in Stachel and Harris,
with Zr N 50 ppm have NdN/ErN b 5). This suggests that increasing Zr 2008) cannot be coincidental and suggest that the temporal extent of
contents are accompanied by decreasing LREE–HREE fractionation in individual diamond growth events is usually well contained within the
metasomatic fluids. uncertainty of the age dates.
It should, however, be noted that such an approach will only identify Isotopic dating using garnet, clinopyroxene and sulfide inclusions
the dominant re-enrichment event affecting a diamond source region, (reviewed in Gurney et al., 2010; Pearson and Shirey, 1999; Stachel

40 6
Undepleted Lherzolitic
Garnets a Only garnets with b
Harzburgitic Y<5 and Zr<20 ppm
30
Melt

m 4
is
at
om
Y 20 as Yb N
et
M
e lt
M
2
e)
is pit

10
go
m
lo

h
(P
at

Low-T om Fluid
s
Meta
0 0
0 Depleted 50 100 150 0 20 40 60 80
Zr NdN/ErN

Fig. 6. (a): Y versus Zr (wt. ppm) content in peridotitic garnet inclusions in diamond from worldwide sources. Empirical compositional fields and trends are from Griffin and Ryan (1995). The
formation of subcratonic lithospheric mantle (SCLM) involved intense melt extraction and consequently, if no major re-enrichment event occurred, garnets from SCLM plot in the depleted field
(Y b7 ppm and Zr b30 ppm). Low temperature (fluid) metasomatism involves preferential re-enrichment in Zr and is restricted to harzburgitic garnet inclusions. Simultaneous addition of Zr
and Y, attributed to melt metasomatism, is observed for lherzolitic and some harzburgitic garnet inclusions. This is consistent with the observation (see Fig. 1b) that diamond formation mostly
occurs above the solidus of lherzolite (in the presence of a CHO fluid) but below the solidus of harzburgite. The majority of garnet inclusions, however, plot into the depleted field, and for garnets
with Y b5 ppm and Zr b20 ppm assignment to extrapolated metasomatic trends is no longer practical. For such strongly “depleted” garnets (b, right), ubiquitous metasomatic overprint is
nevertheless documented by LREE–HREE ratios (NdN/ErN) invariably near or greater than one. Modest HREE enrichment at low NdN/ErN is interpreted to represent mild melt metasomatism
(trend along the Y-axis) whilst strongly sinusoidal REEN at low HREE (trend along X-axis) document overprint by a strongly fractionated fluid Normalization to C1-chondrite (N) after
McDonough and Sun (1995).
T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220 207

and Harris, 2008) documents that formation of diamonds occurred or carbonate melts, so precipitation of diamond from such melts would
through most of Earth's history (from the Paleoarchean to at least the require an oxidation or reduction reaction. In general, such reactions
Mesozoic). Diamond forming episodes in subcratonic lithospheric mantle could involve reduction of oxidized species such as CO23 − or CO2, or
occur on regional to global scales in response to tectonothermal events oxidation of reduced species such as CH 4 . In addition to being
such as suturing, subduction and plume impact (e.g., Aulbach et al., dissolved in silicate or carbonate melts, oxidized species could also
2009; Gurney et al., 2010; Shirey et al., 2002). Individual diamond be present as crystalline carbonate or in a fluid. Reduced carbon
forming episodes may be associated with particular substrates, with could be present in a melt, a fluid, or under very reduced conditions
harzburgitic paragenesis diamonds generally yielding Paleoarchean in phases such as moissanite (SiC; see Di Pierro et al., 2003; Shiryaev
(3.6–3.2 Ga) ages and lherzolitic paragenesis diamonds forming mostly et al., 2011; Ulmer et al., 1998).
in the Paleoproterozoic at about 2 Ga (see references in Gurney et al., Fluids at mantle conditions may contain oxidized carbon-bearing
2010; Stachel and Harris, 2008). Harzburgitic garnet inclusions from species such as CO 2 or reduced species such as CH4, depending on
Udachnaya, however, yield a 2 Ga isochron age (Richardson and Harris, the oxidation state. There is a constraint on CO2 -bearing fluids in
1997), coinciding with the worldwide peak in lherzolitic diamond forma- peridotitic lithologies, however. Since the landmark work of a num-
tion and younger episodes of peridotitic diamond growth are established ber of research groups in the 1970s, it has been clear that CO2-rich
through Mesoproterozoic isochron ages for lherzolitic sulfide inclusions fluids would not be stable in oxidized peridotitic mantle, but
in diamonds from Ellendale (Western Australia, 1.4 Ga; Smit et al., would react to form carbonates such as dolomite or magnesite by
2010), Mir and 23rd Party Congress (the second diamond growth event reactions such as forsterite + diopside + CO2 → enstatite + dolomite
in Yakutia, 1.0–0.9 Ga; Wiggers de Vries et al., 2013b), a Mesozoic and forsterite + CO2 → enstatite + magnesite (Fig. 7). Another reaction
isochron age obtained from two peridotitic sulfide inclusions in a single relevant to peridotitic mantle is the exchange reaction by which
diamond from Koffiefontein (within error of kimberlite emplacement at enstatite and dolomite react to form the higher-pressure assemblage
90 Ma; Pearson et al., 1998), and an inferred similar age for a peridotitic magnesite and diopside. Comparing these reactions with possible
sulfide inclusion from Jagersfontein (Aulbach et al., 2009). Formation of geotherms for lithospheric mantle (Fig. 7), it is apparent that magnesite
diamonds hosted by eclogite is documented from the Mesoarchean to would be the stable carbonate under most circumstances in oxidized
the Neoproterozoic (2.9 and 0.6 Ga) and may well continue up to the lithospheric peridotitic mantle, and thus carbon would be present in
present. magnesite rather than in a CO2-rich fluid. At more reduced conditions,
For fibrous coats (overgrowths over pre-existing monocrystalline magnesite becomes unstable relative to graphite or diamond via the
diamonds) from Aikhal, Siberia, an Ar–Ar study suggested that the reaction enstatite + magnesite → forsterite + graphite/diamond + O2
coat-forming event occurred close to the time of host kimberlite erup- (the EMOG/EMOD reactions of Eggler and Baker, 1982). Whether the
tion (Burgess et al., 2002). This is consistent with infrared spectroscopic mantle is sufficiently oxidized to stabilize carbonate requires evaluation
studies on fibrous diamonds (coats and cuboids) from southern and of the evidence that constrains the oxidation state of diamond substrates
western Africa, Siberia and Australia, documenting low nitrogen aggre- in the upper mantle.
gation states (pure Type IaA; Boyd et al., 1987, 1992) and thereby indi-
cating that their formation likely occurred penecontemporaneously 5.2. Oxidized and reduced mineral inclusions in diamond
with and probably initiated by host kimberlite activity (Boyd et al.,
1987, 1992; Gurney et al., 2010). Fibrous coats on diamonds from tertia- Mineral inclusions in diamond place constraints on the range of redox
ry kimberlites at Lac de Gras, Canada (Gurney et al., 2004), consequently, conditions associated with natural diamond formation. The upper limit of
may represent the youngest diamond samples available.

5. Diamond forming reactions 1


2
3
100
5.1. Diamond formation
Gr
Dia
adiabat

Diamond may form in Earth's mantle by a variety of processes: 4 3


recrystallization of the low-pressure graphite polymorph, precipitation

Depth [km]
from a fluid or melt saturated with carbon, or by oxidation–reduction 140
P [GPa]

reactions involving carbonate or methane. 45


5
The direct conversion of graphite to diamond is a reconstructive
phase transition, and in the absence of a flux or solvent requires consid-
erable overstepping of the reaction boundary; Irifune et al. (2004) found 180
that 15 GPa and 1800 °C, or 12 GPa and 2000 °C were required for this 6
40

reaction to proceed in the laboratory in a static compression mode. On


the other hand, graphite is commonly used as a carbon source in exper-
iments that grow diamond in the presence of metal, carbonate, and 7 220
silicate melts, as well as CHO fluids at pressure–temperature conditions
35

much closer to the graphite–diamond reaction curve. The operative


700 900 1100 1300 1500
mechanism in these cases would be one of dissolution of graphite into
T [°C]
the melt or fluid, and re-precipitation as diamond resulting from the
higher solubility of the metastable polymorph relative to the stable
Fig. 7. Pressure–temperature diagram showing the location of the carbonation/
one at the experimental P–T conditions. decarbonation reactions (1) enstatite + dolomite = forsterite + diopside + CO2 (Wyllie
Industrial syntheses of diamond at high P and T use metallic melts as et al., 1983) and (2) enstatite + magnesite = forsterite + CO2 (Newton and Sharp, 1975).
growth media. Similar metallic melts may be responsible for diamond In both cases, CO2 is present on the high-temperature, low-pressure side of the reaction
growth in the mantle at depths greater than ~250 km where conditions boundary. (3) is the exchange reaction diopside + magnesite = enstatite + dolomite
(Brey et al., 1983) that stabilizes magnesite on the high-pressure side of the reaction. The
become sufficiently reduced to stabilize metal (Frost and McCammon,
35, 40, and 45 mW/m2 geotherms for lithospheric mantle and the 1300 °C mantle adiabat
2008; Rohrbach et al., 2007, 2011). In the shallower lithospheric mantle, are from Hasterok and Chapman (2011). Gr–Dia is the graphite–diamond transition from
conditions are typically too oxidized for metal stability. There is no Day (2012). Magnesite would be the stable carbonate in peridotitic lithospheric mantle
evidence at present for dissolution of elemental carbon in either silicate along most geotherms.
208 T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220

diamond stability with respect to oxygen fugacity is defined by the pres- fO2 (FMQ) = log fO2 (sample) − log fO2 (FMQ), changing the T-log fO2
ence of carbonates. Primary inclusions of carbonate (versus secondary diagram to one expressing T versus Δlog fO2 (FMQ) one (Supplementary
precipitates from original fluid or melt inclusions) are rare but have Fig. 2b). These buffer reactions also depend on pressure to varying extents
been observed in diamonds hosted by both peridotite (magnesite: (Supplementary Fig. 3), so using values of Δlog fO2 (FMQ) to compare
Harris et al., 2004; Phillips et al., 2004; Wang et al., 1996) and eclogite samples from different pressures should be done with the awareness
(calcite and dolomite: Meyer and McCallum, 1986; Sobolev et al., that such values do not normalize out the pressure dependencies as
2009). On the reducing side, diamond stability in mantle peridotite is effectively as the temperature dependencies.
limited through metal saturation (Frost and McCammon, 2008; Wood, Turning to the oxidation state of the mantle in the diamond stability
1993). Ni-free to Ni-poor native iron inclusions have been observed in field, the most direct evidence comes from garnet-bearing peridotites.
diamonds from Yakutia (coexisting with silicates of harzburgitic para- All four minerals in a garnet peridotite (olivine, orthopyroxene,
genesis; Sobolev et al., 1981), Sloan (Meyer and McCallum, 1986) and clinopyroxene, and garnet) contain Fe2 +; the pyroxenes and garnet
Mwadui (rimmed by wüstite; Stachel et al., 1998). Oxygen fugacities can accommodate Fe3+ to variable extents as well. Various reactions
distinctly below the iron–wüstite buffer (Ulmer et al., 1998) are docu- have been proposed as “oxybarometers” for garnet peridotites; Luth
mented by the occurrence of moissanite as inclusion in diamonds from et al. (1990) outlined five:
Monastery (Moore and Gurney, 1989), Argyle (Jaques et al., 1989), and
2þ 3þ 2þ 2þ
George Creek (Chinn, 1995). In the latter two cases, moissanite co- Fe3 Fe2 Si3 O12 ¼ 2 Fe2 SiO4 þ Fe SiO3 þ 0:5O2 ð1Þ
existed with silicate inclusions of the eclogitic suite.
Attempts to constrain the fO2 conditions associated with common 3þ 2þ
Mg3 Fe2 Si3 O12 ¼ MgSiO3 þ Mg2 SiO4 þ Fe2 SiO4 þ 0:5O2 ð2Þ
inclusions in diamond have so far been limited to the application
of the olivine–orthopyroxene–spinel oxybarometer to diamonds of 3þ
2Ca3 Fe2 Si3 O12 þ 6MgSiO3 þ 4Fe

SiO3
peridotitic paragenesis from several kimberlites on the Kaapvaal Craton 2þ
¼ 6CaMgSi2 O6 þ 4 Fe2 SiO4 þ O2 ð3Þ
(Daniels and Gurney, 1991) and the Limpopo Belt (Kopylova et al.,
1997). Considering all uncertainties, including significant differences 3þ 2þ
between the various existing calibrations (Ballhaus et al., 1991; O'Neill 2 Ca3 Fe2 Si3 O12 þ 2 Fe3 Al2 Si3 O12
and Wall, 1987), the results provide no additional constraints ¼ 2 Ca3 Al2 Si3 O12 þ 4 Fe2 SiO4 þ 2 FeSiO3 þ O2 ð4Þ
beyond the limitation of diamond stability to fO2 conditions bounded

by carbonate and Fe–metal forming reactions. 2Ca3 Fe2 Si3 O12 þ 2Mg3 Al2 Si3 O12 þ 4FeSiO3
¼ 2Ca3 Al2 Si3 O12 þ 4Fe2 SiO4 þ 6MgSiO3 þ O2 : ð5Þ
5.3. Oxidation state of Earth's upper mantle
The thermochemical data available to Luth et al. (1990) restrict-
Discussion of the oxidation state of natural samples is usually couched ed their analysis to the reactions involving the Ca 3 Fe 2 Si 3 O12 end-
in terms of their oxygen fugacity, which stems from the use of oxygen member. Following the determination of the free energy of forma-
buffers in experimental petrology pioneered by Eugster (1957). These tion of the Fe3Fe2Si3O12 end-member by Woodland and O'Neill
buffers are assemblages that control the chemical potential of oxygen by (1993), Gudmundsson and Wood (1995) experimentally constrained
reactions such as: reactions (1) and (5) at 1300 °C and 2.5–3.5 GPa. They found (1) to be
:: more precise, and until recently all subsequent studies have used their
2x Fe þ O2 ¼ 2 Fex O ðiron–wustite; IWÞ calibration for Δlog fO2 (FMQ) based on reaction (1), with the correc-
2Ni þ O2 ¼ 2NiO ðnickel–nickel oxide; NNOÞ tion to a typographical error as noted by Woodland and Peltonen
:
3 Fe2 SiO4 þ O2 ¼ 2Fe3 O4 þ 3SiO2 ðfayalite–magnetite–quartz; FMQ Þ (1999). Oxybarometry of garnet peridotites was revisited recently by
4 Fe3 O4 þ O2 ¼ 6 Fe2 O3 ðmagnetite−hematite; MHÞ
Stagno et al. (2013), who did experiments at 3, 6, and 7 GPa at temper-
atures between 1300 and 1600 °C. They proposed a new calibration
For each of these reactions, an equilibrium expression of the form
based on reaction (5), which reproduced their data to better precision
than the previous calibration of Gudmundsson and Wood (1995).
a6Fe2 O3
KMH ¼ There have been a number of studies on the oxidation state of
4
a Fe3 O4 f O2 cratonic garnet peridotites from various localities (Canil and O'Neill,
1996; Creighton et al., 2009, 2010; Goncharov and Ionov, 2012;
may be written. Combining this equation with that relating the equilib- Goncharov et al., 2012; Lazarov et al., 2009; Luth et al., 1990;
rium constant to the standard free energy for this reaction McCammon and Kopylova, 2004; Woodland, 2009; Woodland and
o
Koch, 2003; Woodland and Peltonen, 1999; Yaxley et al., 2012). These
RT lnK MH ¼ −Δr G studies have produced a dataset of 225 samples from cratons world-
wide, with almost half of the data coming from the Kaapvaal Craton in
produces an expression for fO2 (by convention expressed as log10 fO2) of southern Africa. Of the samples studied, very few have been document-
the form: ed to contain diamond or graphite. The one graphite-bearing sample
studied by Canil and O'Neill (1996) is a low-pressure sample (2.1 GPa,
Δr Go
log f O2 ðMH Þ ¼ þ 6 log a Fe2 O3 −4 log a Fe3 O4 ; 710 °C). Two diamond-bearing samples from Finsch have been studied;
2:303RT
one (865) by both Canil and O'Neill (1996) and Lazarov et al. (2009). The
which in the case of pure magnetite and hematite reduces to: other (F556) is from the Canil and O'Neill (1996) study.
There have also been recent studies of ultrahigh-pressure orogenic
o
Δr G peridotites interpreted to be samples of mantle wedges (Malaspina
log f O2 ðMH Þ ¼ : et al., 2009, 2010) that yield fascinating insights into the oxidation state
2:303RT
in these environments. However, given the focus of this review on
The calculated locations of these reactions in T-log fO2 space at 4 GPa diamond formation in cratonic lithospheric mantle, space prohibits a full
are shown in Supplementary Fig. 2a. Because these reactions are nearly exploration of variations in oxidation state with tectonic environment.
parallel in this diagram, much of the relative temperature dependence is To place the available data on cratonic samples in context in terms of
removed by expressing the values of fO2 relative to a reference reaction. their depths of origin and the geothermal conditions recorded by their
The most common reference in use is the FMQ reaction, such that Δlog mineral compositions, they are shown on a P–T diagram (Fig. 8) along
T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220 209

with conductive reference geotherms and the mantle adiabat from increasing depth. Some localities show trends with less scatter than
Hasterok and Chapman (2011). The samples from the Slave Craton others (e.g., the Early Cretaceous Group II kimberlite at Finsch versus
mostly plot between the 35 and 40 mW/m2 geotherms, with a the Late Cretaceous Group I kimberlites of Kimberley). For the Kaapvaal
few recording hotter conditions. The Siberia samples scatter more Craton, there is also a correlation with texture; the “sheared” peridotites
widely, whereas those from the Kaapvaal record conditions near the (open symbols) tend to be more oxidized than the “coarse” ones (filled
40 mW/m2 geotherm, with the deepest samples actually plotting symbols).
along the extrapolation of that geotherm beyond the mantle adiabat A first order observation is that there is considerable variability in
of Hasterok and Chapman (2011). the relative oxidation states recorded even by samples from the same
To show the oxidation state of these samples, we can plot them on a locality. The Kimberley samples from the Kaapvaal Craton and both
depth/pressure versus Δlog fO2 (FMQ) diagram, on which the samples the Diavik and Jericho suites from the Slave Craton are excellent exam-
are placed at the pressure derived from their mineral thermobarometry ples of this variability, in that these samples present ranges in Δlog fO2
(Fig. 9). Temperature would also be increasing with depth (cf. Fig. 8). For (FMQ) on the same order as that of the global dataset. This variability
comparative purposes, we also plot the iron–wüstite (IW), EMOG/D, begs the question as to whether these values reflect the differences in
and graphite–diamond reactions on these diagrams. These reactions oxidation state(s) at the time of diamond formation, or processes that
are all calculated for the P–T conditions along a 40 mW/m2 model post-date diamond formation. If the latter is the case, then the signifi-
geotherm (Hasterok and Chapman, 2011). cance of the current values is as a record of whether the lithospheric
The IW curve approximates the fO2 at which metal becomes stable; mantle was “diamond-friendly” or “diamond-hostile”, at least at the
as shown by Frost and McCammon (2008), for example, the fO2 at time of sampling by the kimberlite host magma.
which Ni-rich metal becomes stable is very close to the location of the What causes the variability observed in either the global dataset or
end-member IW reaction (after Ballhaus et al., 1991). The EMOG/D in individual suites? To address this question, we first need to under-
reactions represent the upper fO2 bound for the stability of graphite or stand how a peridotite of a given fixed composition, including its FeO
diamond in olivine-bearing mantle relative to crystalline carbonate. and Fe2O3 contents, would change in relative oxidation state with pres-
The “SF10” curves shown in Fig. 9 represent the upper bound when sure and temperature along the geotherm. Then, to explain the variation
the carbonate is present as a carbonatitic melt (Stagno and Frost, 2010). at a given depth, we need to determine how robust the relative oxida-
Comparing the oxidation states of these samples calculated using the tion state values are to modification by interactions with fluids or melts.
calibration of Stagno et al. (2013; “S13”) with those from the calibration Wood et al. (1990) showed that in an isochemical mantle, the rela-
of Gudmundsson and Wood (1995; “GW95”) (Fig. 9a and b), there is tive oxidation state of garnet peridotite should decrease with depth
reasonable agreement at low pressure (b 5 GPa), but increasing differ- along the geotherm because of the change in volume for reactions
ences at higher pressure (Fig. 10), with S13 yielding more oxidized that stabilize Fe3+-bearing garnet. This result has been confirmed by
values. S13 also yields a wider range of values at a given pressure than subsequent modeling by a number of workers (Ballhaus and Frost,
GW95 (Fig. 9a and b). For both oxybarometers, calculated Δlog fO2 1994; Creighton et al., 2009; Frost and McCammon, 2008; Luth and
(FMQ) for a given Fe3 +/ΣFe in garnet becomes more reduced with Stachel, 2014) (Supplementary Fig. 5).
increasing depth (Supplementary Fig. 4). Superimposing the results of the “Luth/Stachel” calculations for
Separating the global dataset allows us to understand better what is pyrolite with different bulk Fe3+/ΣFe on the global xenolith dataset
happening on a regional basis. We will consider the two cratons with (Fig. 11), it is noted that a bulk Fe3+/ΣFe greater than 0.05 would stabilize
the most data: the Kaapvaal and Slave Cratons. The results are illustrated carbonate rather than diamond in a pyrolite mantle composition, and that
in Fig. 9c and d, which show the trend to more reduced values with most samples lie in a field bounded on the low-fO2 side by the 0.02 curve.
It is also clear in Fig. 11 that the variability in Δlog fO2 (FMQ) at a given
depth can be explained by changing the bulk Fe3+/ΣFe of a peridotite —
45
by adding or removing oxygen, in other words, without requiring any
3
100 other compositional change. Confounding this straightforward interpre-
40

tation of the dataset, however, is the fact that many of the natural samples
are more depleted than pyrolite (and variably so), and thus the effects of
35

adiabat

4 depletion – which would be reflected in both modal mineralogy and


mineral composition – need to be evaluated. How will the more deplet-
140
Depth (km)

ed nature of cratonic peridotite affect the sensitivity of the calculated


P [GPa]

5 Δlog fO2 (FMQ) to addition or removal of oxygen by interactions with


Gr fluids or melts? It is reasonable to suppose that decreasing the modal
Dia abundance of Fe3 +-bearing minerals – clinopyroxene and garnet in
180 particular – should decrease the amount of oxygen required to shift
6
Kaapvaal the fO2 of the rock.
Siberia To quantify this, Luth and Stachel (2014) calculated the Δlog fO2
Slave
(FMQ) of a peridotite of an arbitrary bulk composition as a function of
Tanzania
7
Fennoscan 220 bulk Fe3 +/ΣFe, P, and T. Their model, conceptually based on those of
Frost and McCammon (2008) and Stagno et al. (2013), allowed explora-
700 900 1100 1300 1500 tion of the effects of changing bulk composition on the relationship
between bulk Fe3+/ΣFe and the calculated Δlog fO2 (FMQ). Using the
T [°C]
pyrolite mantle composition of McDonough and Sun (1995) as a
Fig. 8. Global dataset of garnet peridotites with measured Fe3+/ΣFe (garnet). Data at less
starting point, they varied the modal mineralogy (and hence the
than 2.5 GPa (mostly Vitim, southern Siberia data from Goncharov and Ionov, 2012) are bulk composition) and found – as anticipated – decreasing the
not plotted. The 35, 40, and 45 mW/m2 model conductive geotherms for lithospheric man- amount of clinopyroxene and garnet decreased the change in
tle and the 1300 °C mantle adiabat are from Hasterok and Chapman (2011). Temperatures Fe3 +/ΣFe required to move the fO 2 of a sample from IW to EMOD.
and pressures were taken from the original papers except for those of Luth et al. (1990)
For example, at ~200 km depth (approximating the base of the litho-
and Canil and O'Neill (1996), which were re-calculated as per Creighton et al. (2009),
i.e., using PBKN (Brey and Köhler, 1990) in combination with TO'Neill (O'Neill, 1980; O'Neill sphere), the Fe3+/ΣFe of pyrolite increases from 0.017 at IW to 0.065 at
and Wood, 1979). Blue circles are diamond-bearing peridotites (Canil and O’Neill, 1996; EMOD (i.e., 382%). For comparison, the Fe3+/ΣFe of a peridotite with 5%
Lazarov et al., 2009). garnet and 5% clinopyroxene would change from 0.009 to 0.026 (289%)
210 T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220

a b

EMOG
IW IW
3 100

120
4

Gr Gr 140

Depth [km]
P [GPa]

Dia Dia
5 160

GW95 180
6 Calibration
Kaapvaal
200

EMOD

EMOD
Siberia

SF10
SF10
Slave
S13 Tanzania
7 220
Calibration Fennoscan

c d
EMOG

EMOG
IW IW
3 100

120
4
SF10

SF10
Gr Gr 140

Depth [km]
P [GPa]

Dia Dia
EMOD

5 160

Kimberley 180
Lesotho
6 Frank Smith
EMOD

Monastery
200

EMOD
Cullinan
(Premier)
Gibeon Diavik
7 Finsch Gahcho Kue 220
(Group II) Jericho

-6 -4 -2 0 -6 -4 -2 0
Δlog fO2 (FMQ) Δlog fO2 (FMQ)

Fig. 9. Top figures: Comparison of the oxidation states of 225 garnet peridotites calculated with the Stagno et al. (2013) calibration (a) and with the Gudmundsson and Wood (1995)
calibration (b). Data for samples with pressures b2.5 GPa are omitted. The curves shown are calculated at the pressure–temperature conditions along a 40 mW/m2 geotherm
(Hasterok and Chapman, 2011). Blue circles in (a) are diamond-bearing peridotites (see text). Bottom figures: Comparison of the oxidation states of garnet peridotites from the Kaapvaal
(c) and Slave (d) cratons, both calculated with the Stagno et al. (2013) oxybarometer calibration. In (c), the solid symbols are “coarse” textured samples, open symbols are “sheared”
samples. Darker blue symbols are diamond-bearing samples from Finsch. Abbreviations: IW is the iron–wüstite buffer reaction, Gr–Dia is the graphite–diamond reaction; EMOG/
EMOD is the enstatite + magnesite = olivine + graphite or diamond reaction; SF10 is the calculated location of the analogous reaction involving carbonate melt rather than crystalline
carbonate. IW from Ballhaus et al. (1991). Gr-Dia, EMOG, EMOD reactions calculated from the thermodynamic dataset of Holland and Powell (2011). SF10 curve from Stagno and Frost
(2010).

over the same range in fO2, and that of a peridotite with 2% garnet and 2% Fe3+/ΣFe — and therefore to changes in oxygen content that could be
clinopyroxene would increase from 0.007 to 0.016 (228%). They then produced by interaction with a melt or fluid, either during the formation
modeled some real samples for which modal mineralogy, mineral compo- of diamond or subsequently. In order to evaluate the effectiveness of
sitions, and Fe3+/ΣFe for garnet are available. Their results are shown for such an interaction, however, we need to turn our attention to what
three natural samples and for the model pyrolite composition in Fig. 12. we know about possible fluids and melts.
All three of the natural samples contain less garnet and clinopyroxene
than pyrolite, which shifts their curves on the Δlog fO2 (FMQ)–Fe3+/ΣFe
diagram to the left. The curves for the natural samples are also steeper 5.4. Fluids in the upper mantle
than pyrolite, so that they require less increase in Fe3+/ΣFe (and therefore
less O2 added to the sample) to change their fO2 from IW to EMOD. Fluids in Earth's upper mantle can be modeled in the CHO system,
Applying their model to a larger set of samples, Luth and Stachel although evidence from natural samples suggests that chlorine, sulfur,
(2014) addressed a simple question: how much O2 is required to shift and nitrogen need to be included in a comprehensive model of mantle
the fO2 of cratonic peridotites from IW to EMOD? Luth and Stachel fluids. In addition, the solubility of silicates in fluids, especially those
(2014) demonstrated that only ~ 400 ppm O2 is required to shift the that are water-rich, increases with increasing pressure to the extent
oxidation state of primitive mantle pyrolite from IW to EMOD. For that a second critical endpoint on the (hydrous) fluid-saturated solidus
depleted peridotites, less than 200 ppm is needed — and for four samples of model eclogite has been found at ~6 GPa (Kessel et al., 2005; Fig. 3).
from Finsch from the study of Lazarov et al. (2009), less than 50 ppm O2 is The second critical endpoint is where there can no longer be any distinc-
needed (Fig. 13). tion between a solute-rich fluid and a fluid-rich melt — so the solidus
The study of Luth and Stachel (2014) quantified the extreme sensi- would terminate at this point. The existence of a second critical end-
tivity of the fO2 of depleted lithospheric peridotites to changes in bulk point on the fluid-saturated solidus of either harzburgite or lherzolite
T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220 211

EMOD
2
-2

Δlog fO2 (FMQ)


P [GPa]

4 -3

5
Kaapvaal MS95 pyrolite
Siberia PHN 5267
-4
6 Slave Finsch F-15
Tanzania Finsch 695
Fennoscan
7 IW
0.00 0.02 0.04 0.06 0.08
-1.0 -0.5 0.0 0.5 1.0 1.5
Δlog fO2 (FMQ)S13 - Δlog fO2 (FMQ)GW95 Fe3+/ΣFe (bulk)

Fig. 12. Δlog fO2 (FMQ) versus Fe3+/ΣFe trends calculated for samples from Finsch and
Fig. 10. Differences between the values of Δlog fO2 (FMQ) calculated with the Stagno et al. Cullinan (Premier) compared to the trend for primitive mantle pyrolite. Data sources:
(2013) and the Gudmundsson and Wood (1995) oxybarometers plotted as a function of MS95 pyrolite composition of McDonough and Sun (1995), PHN 5267 (Cullinan)
pressure for the global dataset. data from Boyd and Mertzman (1987) and Canil and O'Neill (1996), and Finsch data
from Lazarov et al. (2009).

has proven to be more difficult to constrain (indeed, as has the location


of solidi in either lithology — see discussion in Luth, 2014). example, a fluid at IW contains ~8 mol% oxygen, whereas one at EMOD
As a starting point for our discussion, we will use the CHO system, contains ~34 mol% O. Because H and C have much lower molar weights
given that there are quantitative models for fluids in this system that than O, the change is more dramatic in mass units; if the fluids are
allow calculation of composition and speciation under both C-saturated expressed in terms of wt. % (of C, H2, and O2), the composition of the
and C-undersaturated conditions (e.g., Huizenga et al., 2012). The most IW fluid is 33 wt.% O2 versus 87 wt.% O2 in the EMOD fluid.
recent thermodynamic model for CHO fluids is that of Zhang and Duan The shape of the carbon–saturation curve also shows how the carbon
(2009, 2010). The compositions of these fluids are usually depicted in content of the fluid changes, and hence how the saturation in diamond
the CHO ternary (Fig. 14), and the speciation in the fluid can be plotted changes, with oxidation state. With increasing fO2, the carbon content
as a function of fO2 (Supplementary Fig. 6). In both cases, pressure and decreases from 20 mol% C in pure CH4 to its minimum value near the
temperature are fixed; these two example figures show the results for O–H sideline at the composition of H2O. Further oxidation increases the
conditions at the base of the lithosphere along a 40 mW/m2 geotherm. carbon content to its maximum value of 33 mol% at pure CO2. The impli-
From these diagrams, we see that fluids coexisting with diamond would cations for diamond formation are straightforward: oxidation of a
be CH4-rich at reduced conditions (IW), nearly pure H2O at the water reduced fluid with a starting fO2 at IW will precipitate diamond until con-
maximum (labeled “m” on Supplementary Fig. 6), and water-rich ditions of the water maximum are reached. Conceptually, this precipita-
H2O + CO2 at the maximum fO2 at which diamond would be stable in tion can be represented by the reaction CH4 + O2 → C + 2 H2O
olivine-bearing mantle (EMOD). In the absence of olivine, diamond
could coexist with more CO2-rich fluids (dominantly H2O–CO2 mixtures)
O2 required for IW to EMOD shift (ppm)

at higher fO2. 200


The carbon–saturation curve (red curve in Fig. 14) shows the change

carats diamond per metric ton


in fluid composition from CH4-rich to CO2-rich with increasing fO2. For
60 300
150 ppm C
2

4
5
0.06
7

IW
3
0.0

0.0

0.0
0.0
0.0

100 40 200
100

120
Depth [km]

4
P [GPa]

50 20 100
Gr 140
Dia PHN 5267
5 Finsch samples
160
0 0 0
0 2 4 6 8 10 12 14
180
EMOD

6 Modal garnet (wt.%)


200 Fig. 13. Amount of O2 required to shift a sample from IW to EMOD, plotted against the modal
-6 -4 -2 0 garnet in the sample. The right hand axis indicates the amount of elemental carbon (diamond
Δlog fO2 (FMQ) or graphite in ppm and carats per ton) that is precipitated during reduction of CO2 (C:O =
1:2) in the fluid. Finsch samples are cpx-bearing peridotites from Lazarov et al. (2009)
(their samples 695, F-1, F-3, F-5, F-6, F-8, F-11, F-12, F-14, F-15, F-16). Calculated at stated
Fig. 11. Calculated oxygen fugacities for a primitive mantle composition (McDonough and P–T of equilibration for each sample. For comparison to the curves in Fig. 12, samples 695
Sun, 1995) as a function of bulk Fe3+/ΣFe along a 40 mW/m2 geotherm (Hasterok and and F-15 are the most garnet-poor and most garnet-rich points, respectively, of the “Finsch
Chapman, 2011). Numbers at top of curves are the bulk Fe3+/ΣFe of the mantle composition. sample” on this diagram. MS95 pyrolite requires ~400 ppm O2 to shift from IW to EMOD
Oxygen fugacities are calculated with the Stagno et al. (2013) oxybarometer. For reference, (Luth and Stachel, 2014), causing ~150 ppm C (equivalent to 750 carats diamond per metric
the global xenolith database and reaction boundaries from Fig. 9 are shown as well. ton) to precipitate.
212 T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220

C The location of the carbon saturation curve in the CHO ternary


moves with changing pressure and temperature (as shown earlier by
Huizenga et al., 2012). It is difficult to show these changes at the scale
of the whole ternary diagram; a more useful approach is to zoom in
by plotting the concentration of carbon in the fluid directly. The concen-
tration of carbon in a carbon-saturated fluid is a strong function of bulk
composition, which can be expressed as molar O/(O + H), equivalent to
the O–H base of the CHO ternary (Fig. 14). Changing temperature has
Diamond the largest influence on composition near the water maximum
+ (Fig. 15a). At the water maximum, the solubility of carbon (as CO2 and
Fluid -0.31 CO2 CH4) increases most strongly with increasing temperature, almost
doubling in amount with a 200 °C increase in temperature. At higher
CH4 -5 IW and lower O/(O + H) than those shown in Fig. 15a, the changes in C
-4 content of the fluid are negligible. The effect of pressure is greatest at
-3 -1
-2 the water maximum as well, with the content of C decreasing with
EMOD
increasing pressure. In reducing fluids, however, the C content increases
H H2O O with increasing pressure.
Because both pressure and temperature increase with depth along
Fig. 14. Molar ternary diagram showing the compositions of fluids (red curves) in equilibri-
a geotherm, the opposing pressure and temperature effects on the
um with diamond at the base of the lithosphere along a 40 mW/m2 geotherm (192 km,
6.1 GPa, 1356 °C; see Huizenga et al. (2012) for the composition of fluids in equilibrium solubility of C in fluids near the water maximum means that we cannot
with diamond at 5.0 GPa, 1227 °C). The numbers along the curve are values of Δlog fO2 predict a priori how the C-saturation surface near the water maximum
(FMQ). For reference, the location of the Δlog fO2 (FMQ) values for IW and EMOD are will change. Calculating the compositions of the fluids with changing
shown as well. Calculated with GFluid (Zhang and Duan, 2010).
O/(O + H) at two depths along a 40 mW/m2 geotherm, however,
shows that the carbon content decreases with decreasing depth in
both the water maximum fluid and reduced fluids (Fig. 15b). The effect
drops off rapidly as the fluids become more oxidized than the water
(Huizenga et al., 2012; Taylor and Green, 1989). With further oxidation, maximum, such that the two curves are coincident by EMOD-like
the fluid will consume diamond rather than precipitate it. Conversely, conditions (O/(O + H) ~ 0.35). The decrease in maximum C content
an oxidized fluid coexisting with diamond would have a starting fO2 no with ascent in reduced fluids reflects a pressure effect.
higher than EMOD in a peridotitic mantle, and could only precipitate Another way to view the change in composition of the fluid along
diamond upon reduction via CO2 → C + O2 until reaching the water max- the geotherm is to look specifically at the water-maximum fluid, and
imum. Further reduction of such a fluid would resorb diamond as the fluid to see how the solubility of carbon species in that fluid changes. This
became more C-rich. fluid has O/(O + H) = 0.333, but the C content in the fluid increases
The amounts of oxygen involved are not trivial; 119 g O2 have to dramatically, from 0.32 mol% at 3 GPa to 0.91 mol% at 6 GPa. This change
be added to 100 g starting IW fluid, and 45 g diamond has to precip- in the C content of the fluid essentially results from the oxygen-
itate for the fluid to evolve to the water maximum composition at the conservative reaction CO2 + CH4 → C + 2H2O.
P and T conditions of Fig. 14. The amounts of oxygen and carbon As well as composition, the speciation in the fluid will change with
removed from the fluid in reducing an EMOD fluid to the water max- depth along the geotherm (Fig. 16). Comparing the curves for X(H2O)
imum composition are smaller (5.7 g O2 and 1.0 g C per 100 g original and X(CH4) in the fluid with the global xenolith dataset, most of the
EMOD fluid), because of the compositional proximity of the two samples plot to the reduced side of the water maximum, and would
fluids. coexist with water-rich H2O–CH4 fluids.

10 20
a b
mole % C in C-saturated fluid

8
15

6
10
4

5
2 1556°C
1456°C 192 km
1356°C 120 km
0
0.20 0.25 0.30 0.35 0.40 0.1 0.2 0.3 0.4
O/(O + H) (molar) O/(O + H) (molar)

Fig. 15. (a) Change in the carbon content of diamond-saturated fluid with temperature at 192 km depth as a function of the O/(O + H) of the fluid. The black curve shows the composition of the
fluid at the temperature appropriate for a 40 mW/m2 geotherm at that depth. The other two curves illustrate the increasing solubility of C with increasing temperature. (b) Change in the carbon
content of C-saturated fluid with depth along a 40 mW/m2 geotherm as a function of the O/(O + H) of the fluid. For reference, the O/(O + H) for a C-saturated fluid with fO2 equal to IW is
0.10 at 192 km and 0.03 at 120 km depth. The O/(O + H) for a C-saturated fluid with fO2 equal to EMOD is 0.35 at 192 km and 0.34 at 120 km depth. Calculated with GFluid (Zhang and
Duan, 2010).
T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220 213

0 .0 1

0 .0 5
Because the fluid has to change composition to maintain equilibrium
with a constant-composition mantle as it ascends along the geotherm, a
3 IW source of oxygen (or sink for hydrogen) would be required, in addition
100
to the precipitation of carbon that takes place. If the surrounding mantle
were to act as this source or sink, it follows that the “isochemical
120
4 mantle” assumption would no longer hold, and the mantle would
have to become more reduced with decreasing depth. The ability of
0.8

140 the mantle to do so will be addressed presently, but we will first consid-

Depth [km]
Gr
P [GPa]

Dia
er the case of a fluid cooling isobarically or ascending along the
5 160 geotherm without exchanging O (or H) with the surrounding mantle.

180 5.6. “Isochemical” ascent or isobaric cooling of fluids


0.6

6
0.4

200 What happens to a C-saturated CHO fluid that ascends along the
EMOD
0.2

geotherm “isochemically” — that is, without exchanging either oxygen


7 220 or hydrogen with the surrounding mantle? Given the expansion of the
max

0.8
0.2

0.4

0.2
0.6

0.8

C-saturation curve with ascent along a geotherm (Fig. 15b), an ascent


-6 -4 -2 0 in which the O/(O + H) of the fluid remains constant would oversatu-
Δlog fO2 (FMQ) rate a C-saturated fluid and cause precipitation of diamond. This effect
can be seen from the relative locations of the two C-saturation surfaces
in Fig. 15b — a fluid composition that is C-saturated at 192 km depth
Fig. 16. Pressure-Δlog fO2 (FMQ) diagram calculated along a 40 mW/m2 geotherm illustrat-
ing how the speciation of CHO fluids changes along the geotherm. Solid gray lines show the
would contain more C than would be stable in the fluid with the same
mole fraction of H2O, long dashed gray lines show the mole fraction of CH4, and short dashed O/(O + H) at 120 km depth (at least for O/(O + H) ≲ 0.35; that is, for
black lines are the mole fraction of CO2 for 1 and 5 mol% CO2; at more oxidized conditions, the fluids with fO2 of EMOD or lower).
fluid lacks CH4 and is essentially H2O + CO2, so the X(CO2) = 0.2 curve coincides with the Along with precipitating diamond, CHO fluids ascending along the
curve for X(H2O) = 0.8. Calculated with GFluid (Zhang and Duan, 2010). For reference, the
geotherm at constant O/(O + H) will shift to more oxidized values
global xenolith database and reaction boundaries from Fig. 9 are shown as well.
relative to FMQ (Fig. 18a). Fig. 18a also shows the amount of carbon
that would precipitate from four distinct fluids as a percentage of the
amount of carbon in the original fluid. Arguably, expressing the carbon
5.5. Ascent along lithospheric PT–fO2 path precipitation in this fashion fails to account for the effect of the decreas-
ing carbon content in the fluid with increasing O/(O + H). In terms of
By comparing the change in fO2 of an isochemical mantle composi- the mass of carbon precipitated, more reduced fluids precipitate more
tion and the fluid speciation along the geotherm (Figs. 11 and 16), it is carbon (Fig. 18b), but at a lower percentage simply because they are
apparent that a fluid in equilibrium with an isochemical mantle would more carbon-rich.
become more oxidized at shallower depths. This effect can be illustrated The change in carbon content of a diamond saturated fluid with tem-
quantitatively for two values of bulk Fe3+/ΣFe of the primitive mantle perature only (i.e., isobaric cooling) suggests an alternative mode of
(0.02 and 0.03). In both cases, the fluids coexisting with the mantle diamond formation. Fig. 15a shows that the decrease in maximum car-
become more oxidized and lower in carbon content with decreasing bon content with temperature is particularly prominent for fluids near
depth as they shift to higher O/(O + H) and move along the C- the water maximum. This implies that even fluids that initially were
saturation surface towards the water maximum (Fig. 17). not saturated in carbon may begin to precipitate diamond during
cooling. A possible scenario for the infiltration of CHO fluids hotter
than ambient mantle is the release of fluid from crystallizing magmas
or mantle plumes. As shown in Section 3.4, diamond formation general-
14 ly takes place below the wet solidus of harzburgite but above the wet
200
solidi of lherzolite and eclogite. This effectively limits diamond precipi-
12 tation from cooling water-rich CHO fluids to harzburgitic substrates as
mole % C in C-saturated fluid

150
fluid dilution through a melt component would occur in lherzolitic
10
and eclogitic lithologies (except along conductive model geotherms
b40 mW/m2 at pressures b5 GPa; see Fig. 1). This may explain the
8
observed close relationship between harzburgitic (“G10”) garnets and
6 diamond (Gurney, 1984) employed as the principal tool of indicator
200 mineral based diamond exploration. More reducing fluids may also
4 penetrate lherzolitic and eclogitic lithologies due to the higher solidus
temperature associated with such fluids, but as seen in Fig. 15a, the
100 150 change in maximum carbon content of such fluids with temperature is
2
Fe3+/ΣFe = 0.02 negligible.
0 Fe3+/ΣFe = 0.03 100
5.7. Diamond precipitation from ascending or isobarically cooling melts
0.10 0.15 0.20 0.25 0.30 0.35
In the presence of methane-rich fluids, the high solidus temperature
O/(O + H) (molar) of peridotite and to a lesser degree eclogite largely precludes percola-
tion of strongly reducing melts through the lithospheric mantle. There-
Fig. 17. Change in carbon content of a C-saturated fluid in equilibrium with primitive mantle
fore, for melt associated diamond precipitation, likely to dominate in
compositions with constant bulk Fe3+/ΣFe of 0.02 and 0.03 (compare Fig. 11). Numbers by
symbols are depth in km. In both cases, the fluids become more oxidized with decreasing eclogitic and lherzolitic substrates (see Section 3.4), only carbonate-
depth and consequently liberate elemental carbon upon ascent. Calculated with GFluid bearing melts need to be considered. With decreasing carbonate con-
(Zhang and Duan, 2010). tent such melts become stable to fO2 conditions far below the EMOD
214 T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220

100

01

0
a

0 .0

0 .0

0 .1

0 .5
4% 6% 49% 2% 3 IW
120 100

EMOG
4
IW 120
140 4

Depth [km]
Gr
P [GPa]

Dia 140

Depth [km]
Gr

P [GPa]
5 Dia
160
5 160

180 180
max

EMOD 6
6
200

EMOD
200
-4 -2 0 7 220

max
Δlog fO2 (FMQ)
2.5 -6 -4 -2 0
g C precipitated per 100 g original fluid

b Δlog fO2 (FMQ)

2.0
Fig. 19. Pressure-Δlog fO2 (FMQ) diagram calculated along a 40 mW/m2 geotherm. Light blue
curves show the maximum-fO2 stability of carbonate-bearing melts according to the model
of Stagno and Frost (2010). Numbers at the top of the curves indicate the molar fraction of
1.5 CO2 (Xmelt(CO2)) in the melt (pure carbonate melt has Xmelt(CO2) = 0.5). Unlabeled curves
(from left to right) are for Xmelt(CO2) = 0.005, 0.05, 0.2, 0.3, and 0.4.

1.0 The high abundance of eclogitic suite diamonds (1/3 of all inclusion-
bearing diamonds) derived from a volumetrically very minor component
of lithospheric mantle (b1% to 5%; Dawson and Stephens, 1975; Schulze,
0.5 1989; McLean et al., 2007) implies that a fundamental difference exists
between diamond precipitation in lherzolite and eclogite. The extension
of diamond stability in eclogite, compared to peridotite, by at least one
0.0
0.1 0.2 0.3 0.4 log unit to more oxidizing conditions (shift from EMOD to the DCDD
[dolomite + coesite = diopside + diamond] equilibrium; Luth, 1993)
O/(O + H) (molar)
potentially establishes such a key difference. Relative to pure diopside,
the omphacitic nature of eclogitic clinopyroxene shifts the DCDD equilib-
Fig. 18. (a) Pressure-Δlog fO2 (FMQ) diagram calculated along a 40 mW/m2 geotherm.
Yellow-orange lines show the trajectory of four CHO fluid compositions ascending along
rium even further to more oxidizing conditions (see Fig. 2 in Luth, 1993).
the geotherm at constant O/(O + H). The numbers at the top of each arrow provide the This expansion of diamond stability to higher fO2 values may strongly
percentage of carbon in the original fluid that precipitates during the ascent of each enhance diamond precipitation directly from cooling or upward migrat-
fluid from 192 to 120 km. The trajectory for the most oxidized fluid, as it is entirely ing carbonate-bearing melts. Alternatively, the redox buffering capacity
above the fO2 of EMOD, could only occur in an olivine-free lithology. Curves calculated
of eclogite, which is much more Fe-rich than peridotite, far exceeds that
with GFluid (Zhang and Duan, 2010). For reference, the global xenolith database
(gray symbols) and reaction boundaries from Fig. 9 are shown as well. (b) Amount of car- of cratonic peridotites and, consequently, the possible extent of diamond
bon precipitated from fluids during ascent along a 40 mW/m2 geotherm, as a function of precipitation during carbonate reduction is greatly enhanced. Olivine-
O/(O + H). The amount of carbon is expressed as the mass of carbon precipitated from free garnet pyroxenite layers with original diamond contents of up to
100 g of the fluid that was present at 192 km. Calculated with GFluid (Zhang and Duan, 15 vol.% (preserved as graphite pseudomorphs) embedded in non-
2010).
diamondiferous peridotite at Beni Bousera (Pearson et al., 1989) strong-
ly support localized diamond precipitation from melts more oxidizing
than EMOD in olivine-free lithologies. Extreme diamond contents,
buffer (Fig. 19) and consequently, precipitation of diamond from the such as in the Beni Bousera websterites but also in diamondiferous
melt is possible. The precipitation of diamond directly from carbonate- eclogite xenoliths (e.g., up to 20% diamond in eclogite xenoliths
bearing melts would involve an internal redistribution of oxygen and from the Jericho kimberlite; Smart et al., 2009) cannot conceivably
thus have an oxidizing effect on the residual melt, which ultimately result from wall rock buffered redox reactions and, thus, support
may stall further diamond precipitation. Percolating carbonate-bearing an “isochemical” mode of diamond precipitation from cooling or
melts are in equilibrium with the surrounding peridotitic mantle ascending carbonate-bearing melts. The redistribution of oxygen
(e.g., Hiraga and Kohlstedt, 2009; Watson et al., 1990) and, consequent- within the melt and consequent increase in fO 2 associated with
ly, have the full peridotitic mineral assemblage as liquidus phases. In the this “isochemical” mode of diamond precipitation may be alleviated
absence of adequate thermodynamic data for carbonate-bearing melts when continuous flow of percolating melt into eclogite/pyroxenite
we cannot evaluate if they would experience a similar decrease in max- involving incremental diamond growth is considered. When such
imum carbon content as CHO fluids during cooling or ascent along a melts cross from ubiquitous peridotitic lithologies into eclogite,
geotherm. The possible appearance of diamond on the liquidus of they will also be out of equilibrium with their new eclogitic wall
cooling and evolving carbonate-bearing melts at fO2 conditions below rocks, which at fO 2 conditions below the DCDD equilibrium may
the EMOD buffer has not been investigated experimentally and, there- enhance diamond precipitation. In any case, that diamond is so
fore, is speculative. The absence of a clear association between diamond much more abundant in eclogite compared to lherzolite must relate to
and lherzolitic garnets, however, suggests that this either is not the case the more oxidizing conditions permissive of diamond precipitation in
or only rarely occurs. olivine-free lithologies (Luth, 1993).
T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220 215

5.8. The redox buffering capacity of cratonic peridotite Consequently, diamond stable lithospheric mantle sampled, e.g., by
Mesoproterozoic to Cretaceous kimberlites on the Kaapvaal Craton
As outlined above, redox reactions such as the oxidation of methane last interacted with fluids that invariably were more reducing than
or the reduction of carbonate are thought to be essential aspects of the EMOD buffer (Fig. 9c). In contrast, peridotite xenoliths in the Eocene
diamond formation in Earth's upper mantle. Expressing these two A154 kimberlite at Diavik (Central Slave Craton) show metasomatic
mechanisms by the simple reactions CH4 + O2 = C + 2 H2O and overprint by fluids/melts that are more variable in character, ranging
CO23 − = C + O2 − + O2 demonstrates the fundamental role of O2 in from highly reducing to oxidizing (Fig. 9d).
mass-balancing this process as either a reactant or product. That is not
to say that molecular oxygen per se is involved, but there has to be 6. Diamond forming processes based on co-variations in δ13C–N and
another coupled reaction to either produce or consume oxygen to on fluid inclusions
allow redox precipitation of diamond to occur. A logical candidate for
such a coupled reaction is the iron redox couple, in which reduced The isotopic fractionation factor of carbon (ΔCdiamond-fluid) for
ferrous iron reacts with oxygen to produce oxidized ferric iron. This precipitation of diamond from reduced (methane-bearing) and oxidized
reaction, expressed as 4FeO + O2 = 2Fe2O3, could proceed in this (carbonate- or CO2-bearing) fluids or melts has opposite signs and, on
form in iron-bearing melts, but in the assemblage of solid silicates that this basis, it should be relatively straightforward to derive the
make up peridotitic mantle, charge-balance and crystal–chemical con- mode(s) of natural diamond formation based on population density
straints require more complex reactions to provide or consume oxygen. plots of diamond carbon isotopic compositions (Deines, 1980) and
These reactions are actually the same reactions previously discussed as through evaluation of co-variations between δ13C and the compati-
the basis for oxybarometry in mantle peridotites (Eqs. (1)–(5) above). ble trace element nitrogen (Stachel et al., 2009; Thomassot et al.,
A key issue, therefore, is the ability of these reactions in the litho- 2007). Diamonds in a typical kimberlite deposit, however, originate
spheric mantle to act as the necessary source or sink of oxygen to over a depth range that in some cases encompasses the entire diamond
allow redox precipitation of diamond to occur. In order to address this stable lithospheric mantle and reflect multiple growth episodes (Harte
question, it is necessary to quantify both the amount of oxygen required et al., 1999a; Palot et al., 2013; Wiggers de Vries et al., 2013a, 2013b) rath-
for the redox precipitation of diamond, and the capacity of the mantle to er than single diamond forming events. Variation in the bulk distribution
act as a source or sink for oxygen. The former is reasonably straightfor- coefficients for carbon and nitrogen (e.g., through co-precipitation of
ward, in that the oxygen required can be scaled to the amount of dia- other phases) combined with large differences in the concentration of
mond precipitated, assuming that the redox reactions involve either carbon (major element) and nitrogen (trace element) in the diamond
methane or carbonate. In either case, the molar ratio of O2 to C is 1:1, forming fluids/melts introduce further complexity and may lead to an
which translates into a mass ratio of 2.67 g of O2 per gram of C. apparently decoupled behavior of δ13C and nitrogen concentration. On
As seen from the previous discussion of the study of Luth and Stachel this background, discernible correlations between carbon isotopic com-
(2014), the capacity of mantle peridotite to act as an effective source or position and nitrogen content, indicative of a particular mode of forma-
sink of O2 is limited by the sensitivity of its redox state to small changes tion, are unlikely to be observed based on bulk diamond analyses even
in O2 content, at least relative to the much larger amount of O2 required on the level of individual deposits. The character of diamond forming
to change the oxidation state of a CHO fluid. As shown above, b50 ppm fluids/melts, therefore, is difficult to constrain.
O2 is needed to move the oxidation state of a strongly depleted perido- Thomassot et al. (2007) presented evidence for diamond formation
tite from IW to EMOD (i.e., b5 mg O2 per 100 g rock) compared to the from a reduced, methane-bearing fluid, based on a suite of diamonds
119 g O2 that must be added to 100 g IW fluid to move it to the water recovered from a single garnet–lherzolite xenolith from the Cullinan
maximum composition. (Premier) mine. Nitrogen concentration and aggregation state charac-
It is also worth noting that the amount of diamond that would precip- teristics suggested that the diamonds in this xenolith are cogenetic; a
itate as a result of the shift of the oxidation state of even primitive mantle linear positive correlation between log N and δ13C for these diamonds,
from IW to EMOD is relatively small (~150 ppm — see Fig. 13). For therefore, appeared to indicate operation of a single diamond forming
comparison, diamond-bearing peridotite xenoliths can have 5500 ppm process that could be modeled as precipitation from a methane-
diamond (Viljoen et al., 2004). bearing fluid or melt. Thomassot et al. (2007) also presented nitrogen
Luth and Stachel (2014) modeled various scenarios of oxidized isotopic data (δ15N) for this suite of diamonds and documented a posi-
fluids interacting with reduced peridotite and vice versa; their conclu- tive linear correlation with δ13C, implying (based on a methane precip-
sion was that b50 ppm fluid would be required to reset the oxidation itation model) a positive nitrogen isotope fractionation factor. The
state of depleted peridotite to that of the fluid. Thus, relative to a CHO subsequent determination of a negative sign of ΔNdiamond-fluid (Petts
fluid, the oxygen buffering capacity of cratonic (depleted) peridotites et al., 2014, submitted for publication) is incompatible with the model
is very small, and the ability of such mantle to act as the sink or source of Thomassot et al. (2007) and indicates that a more complex growth
for oxygen necessary for redox precipitation of diamond is quite low. event (e.g., fluid mixing or co-precipitation of nitrogen free phases)
To allow for the growth of common commercial sized diamonds occurred. Consequently, to date there is no compelling case of diamond
through wall rock buffered redox reactions, diffusive transport of precipitation from a reducing fluid documented in the literature that is
oxygen would need to occur over large distances (10s of cm); diffuse built on the stable isotope and nitrogen characteristics of diamond itself.
exchange over such length scales during metasomatic events is incon- Indirect evidence for diamond precipitation from highly reducing fluids
sistent with common evidence for mineral compositional and isotopic is, however, provided by reduced inclusions in diamond (e.g., native
heterogeneity observed on the hand-sample scale in cratonic peridotite iron; see Section 5.2).
xenoliths (e.g., Burgess and Harte, 1999; Schmidberger et al., 2003 and Ion microprobes – for high precision analyses equipped with a large
references therein). magnetic sector and a multi-collection system – enable the study of
From the degree of secondary LREE enrichment in cratonic garnet possible co-variations between δ13C and nitrogen content on the level
peridotites, Luth and Stachel (2014) estimated bulk addition of metaso- of individual diamond growth zones. On this spatially highly resolved
matic fluid/melt approximately in the range of 0.1–5 wt.% — that is 20 to level, strongly correlated variations in δ13C and N-content are occasion-
1000 times the maximum amount of fluid required to reset the oxida- ally observed. To date, the best documented example is diamond 25-S1
tion state of these rocks. This implies that redox profiles through recovered from a low-Mg eclogite xenolith from the Jericho kimberlite
subcratonic lithospheric mantle (Fig. 9) have no bearing on Archean (Smart et al., 2011): in its outer core zone, characterized by homogenous
mantle fO2 but merely are a reflection of the redox state of the last blue CL response, this diamond shows an outward increase in δ13C (and
metasomatic fluid/melt passing through a section of lithosphere. δ15N; Petts et al., 2014; Petts et al., submitted for publication) coupled
216 T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220

with a decrease in nitrogen content. Such a relationship can only be Comparison of the geothermobarometric data with hydrous
modeled based on precipitation from an oxidized (carbonate- or (-carbonated) solidi indicates that diamond formation in lherzolitic
CO2-bearing) fluid or melt (Smart et al., 2011). This conclusion is sup- and eclogitic substrates typically occurs in the presence of a melt, whilst
ported by the occurrence of (rare) carbonate inclusions in smooth- harzburgitic paragenesis diamonds generally form under sub-solidus
surfaced monocrystalline diamond (see Section 5.2). For non-gem fi- conditions. The increase in solidus temperature associated with reduced
brous diamonds, a clear link to carbonatitic melts (“high density fluids”) fluids is permissive of melt-absent diamond formation in lherzolite and
has been established through the study of finely dispersed, micrometer eclogite under such conditions. Consequently, for the dominant
to nanometer-sized original melt inclusions (e.g., Klein-BenDavid et al., harzburgitic paragenesis, diamond formation can only relate to CHO
2009; Kopylova et al., 2010; Navon et al., 1988). Similar carbonatitic fluids (without geothermobarometric constraints on fO2, i.e., within
melt inclusions have now also been documented within smooth- the range from H2O–carbonate to H2O–CH4–H2) whilst for diamonds
surfaced gem-type diamonds (Weiss et al., 2014), providing strong hosted by lherzolite and eclogite both melts (carbonatites or silicate
evidence that carbonate-bearing melts are an important diamond melts with dissolved OH− and CO23 −; e.g., Green and Falloon, 1998)
forming medium. With decreasing carbonate content, such melts may and reducing fluids need to be considered.
represent a spectrum of fO2 conditions from oxidizing (carbonatites) Diamond formation is generally believed to relate to redox reactions
to reducing (possible for carbonated silicate melts; Fig. 19). In view of involving either oxidized (carbonate- or CO2-bearing) fluids/melts
the currently very limited number of in situ studies documenting inter- interacting with reduced wall rocks or reduced (methane-bearing)
nal variations in stable isotope composition and nitrogen content dur- fluids/melts in contact with more oxidized wall rocks. Modeling the
ing gem-diamond growth, the current absence of clear evidence for interaction of fluids with depleted cratonic peridotites indicates, how-
methane-driven diamond precipitation can, however, not be used to ever, that only less than 50 ppm fluid are required to completely reset
discount its operation in nature. Given the limited buffering capacity the oxidation state of depleted peridotite to that of the fluid. This
of cratonic peridotites (see Section 5.8), their overall reduced character extremely limited buffering capacity of cratonic peridotites implies
(Fig. 9a) clearly implies that they last interacted with a reduced fluid or that redox reactions between fluid and wall rock are not an efficient
melt, documenting upward migration of chiefly reduced fluids through way to precipitate diamond and extremely unlikely to produce large
diamond stable lithospheric mantle. Whilst in the case of CHO fluids this macro-diamonds. An additional implication of the limited buffering
implies methane as an important component, in melts the carbon capacity of cratonic peridotites is that fO2 studies on garnet peridotite
speciation is less straightforward to predict and even at the reducing xenoliths generally only determine the redox state of the last metaso-
conditions of typical diamond stable mantle may still involve carbonate matic fluid/melt that interacted with these rocks. The observation that
(see Fig. 19). the bulk of peridotite xenoliths derived from the diamond stability
Based on these results, natural diamond formation involving field yields fO2 conditions more reducing than the EMOD buffer, conse-
methane- and carbonate-bearing fluids/melts are both possible although quently, implies that the last metasomatic event in most sections of
actual diamond based evidence so far points to a strong predominance of deep cratonic lithosphere was reducing in character.
carbonate-bearing diamond forming media. We propose that a much more efficient mode of diamond precipita-
tion is “isochemical” (not involving oxygen exchange with the wall
7. Conclusions rock) cooling of CHO fluids or their “isochemical” ascent along a geo-
thermal gradient. In particular for fluids with compositions close to
Inclusion based geothermobarometry indicates that peridotitic suite the water maximum both scenarios are associated with a distinct
diamonds typically (1 sigma range about the average) originate from a decrease in the solubility of carbon species (CH4, CO2 or CO2−3 ), leading
140–190 km depth at temperatures between 1040–1250 °C. Regression to diamond precipitation in peridotite from fluids more reducing than
of the entire dataset results in a “diamond geotherm” that is slightly the EMOD buffer and in eclogite from fluids more reducing than the
oblique to the model geotherms of Hasterok and Chapman (2011), DCDD equilibrium. Incorporating the derived relationships between
increasing from 36 mW/m2 at the graphite–diamond transition to diamond forming conditions and hydrous solidi, this mode of diamond
39 mW/m2 at the intersection with the mantle adiabat (base of the formation will be largely restricted to harzburgitic substrates and thus
lithosphere). The resulting “diamond window” (diamond stable region may explain the strong association between diamond and harzburgitic
of the lithospheric mantle) extends from 110–205 km depth. In the ab- garnets observed worldwide. For melt-associated diamond precipita-
sence of a reliable barometer for garnet–clinopyroxene assemblages, tion, likely to dominate in eclogitic and lherzolitic substrates, the effect
projection of thermometric data for diamonds hosted by eclogite on of cooling or ascent along a geotherm cannot be assessed in the absence
the peridotitic suite “diamond geotherm” indicates derivation mainly of adequate thermodynamic data. The extreme overabundance of
(1 sigma range about the average) from 155–200 km depth and eclogitic suite diamonds (1/3 of all inclusion bearing diamonds belong
1060–1340 °C. Projection on the 40 mW/m2 Precambrian shield to the eclogitic suite but eclogite constitutes only a very minor portion
geotherm of Hasterok and Chapman (2011) results in a 1 sigma range of lithospheric mantle) indicates, however, that the extension of dia-
of 135–190 km depth at temperatures of 1040–1330 °C. Compared to mond stability by at least one log unit to more oxidizing conditions
the peridotitic suite, typical eclogitic suite diamonds thus show a larger (shift from EMOD to DCDD for olivine-free lithologies) plays a critical
spread in temperatures extending to higher values. role. This expansion of diamond stability in fO2 space may strongly
A detailed comparison of temperature estimates based on non- enhance diamond precipitation either during redox reactions with such
touching (diamond formation) and touching (final residence tempera- far more Fe-rich bulk rock compositions or directly from “isochemically”
ture) inclusion pairs of peridotitic and eclogitic paragenesis indicates cooling or upward migrating carbonate-bearing melts. The occurrence
that diamond formation in some cases is associated with elevated tem- of originally highly diamondiferous garnet pyroxenite layers surrounded
peratures (by about 100–180 °C), whilst in other cases temperatures of by carbon-free orogenic peridotites (Pearson et al., 1989) strongly
formation and final residence are comparable or final residence may supports preferential carbon capture in olivine-free pyroxenite or
even occur at slightly elevated (40 °C) temperatures. For peridotitic eclogite, specifically in settings more oxidized than EMOD. Extreme
suite diamonds from the Panda kimberlite (Central Slave Craton), com- (15–20 vol.%) diamond contents in these garnet websterite layers and
parison with nitrogen-in-diamond based mantle residence temperature in some kimberlite born eclogite xenoliths, and the occurrence of very
estimates indicates that cooling after diamond formation was slow large (cm-sized) diamonds hosted by eclogite appear to be incompatible
(100 s of millions to billion year time scale), suggesting lithosphere- with formation during wall rock buffered redox reactions and again favor
scale thermal relaxation (e.g., after impact of a plume) rather than an “isochemical” mode of precipitation directly from a fluid or melt more
cooling following a local advective heating event. reducing than DCDD.
T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220 217

Stable isotope and nitrogen content studies on diamond itself so far in three dimensions using quantitative ERDA microscopy. Diamond and Related
have not provided a conclusive answer as to what the redox state of the Materials 7 (11–12), 1714–1718.
Creighton, S., Stachel, T., Matveev, S., Höfer, H., McCammon, C., Luth, R.W., 2009. Oxidation of
dominant diamond forming fluid/melt may be. Combining the currently the Kaapvaal lithospheric mantle driven by metasomatism. Contributions to Mineralogy
available data from in situ (ion microprobe) studies with the recent and Petrology 157, 491–504.
finding of broadly carbonatitic high density fluids (melt inclusions) in Creighton, S., Stachel, T., Eichenberg, D., Luth, R.W., 2010. Oxidation state of the lithospheric
mantle beneath Diavik diamond mine, central Slave craton, NWT, Canada. Contributions
gem diamonds, then diamond formation from carbonate-bearing fluids/ to Mineralogy and Petrology 159, 645–657.
melts can be considered as being well established, but given the current Daniels, L.R.M., Gurney, J.J., 1991. Oxygen fugacity constraints on the southern African
paucity of conclusive datasets, diamond formation from reducing, lithosphere. Contributions to Mineralogy and Petrology 108 (1–2), 154–161.
Dasgupta, R., Hirschmann, M.M., Withers, A.C., 2004. Deep global cycling of carbon
methane-bearing fluids cannot be discounted. constrained by the solidus of anhydrous, carbonated eclogite under upper mantle
Supplementary data to this article can be found online at http://dx. conditions. Earth and Planetary Science Letters 227 (1–2), 73–85.
doi.org/10.1016/j.lithos.2015.01.028. Dawson, J.B., Stephens, W.E., 1975. Statistical classification of garnets from kimberlite and
associated xenoliths. Journal of Geology 83 (5), 589–607.
Day, H.W., 2012. A revised diamond–graphite transition curve. American Mineralogist 97
(1), 52–62.
Acknowledgments De Vries, R.C., 1975. Plastic deformation and “work-hardening” of diamond.
Materials Research Bulletin 10, 1193–1200.
Marco Scambeluri is thanked for inviting this review paper and Deines, P., 1980. The carbon isotopic composition of diamonds: relationship to diamond
shape, color, occurrence and vapor composition. Geochimica et Cosmochimica Acta
showing both persistence and patience over the two years of writing.
44, 943–961.
The new concepts presented here matured during fruitful discussions Di Pierro, S., Gnos, E., Grobety, B.H., Armbruster, T., Bernasconi, S.M., Ulmer, P., 2003.
in particular with Herman Grütter, Jeff Harris, Gerhard Brey and Tom Rock-forming moissanite (natural alpha-silicon carbide). American Mineralogist 88,
Chacko. Herman Grütter and Fanus Viljoen are also thanked for their 1817–1821.
Eggler, D.H., Baker, D.R., 1982. Reduced volatiles in the system C–O–H: implications for
detailed input during formal reviews. George Read is thanked for show- mantle melting, fluid formation and diamond genesis. In: Akimoto, S., Manghnani,
ing TS a diamond vein cutting through an eclogite xenolith from Fort a la M.H. (Eds.), High pressure research in geophysics. Center for Academic Publications,
Corne, changing TS' thinking about diamond forming processes. TS and Tokyo, pp. 237–250.
Eugster, H.P., 1957. Heterogeneous reactions involving oxidation and reduction at high
RL both acknowledge funding of their research programs through the pressures and temperatures. Journal of Chemical Physics 26, 1760–1761.
Natural Sciences and Engineering Research Council of Canada (NSERC) Foley, S.F., 2011. A reappraisal of redox melting in the Earth's mantle as a function of
Discovery grants. TS receives additional funding through the Canada tectonic setting and time. Journal of Petrology 52 (7–8), 1363–1391.
Foley, S.F., Yaxley, G.M., Rosenthal, A., Buhre, S., Kiseeva, E.S., Rapp, R.P., Jacob, D.E., 2009.
Research Chairs program. The composition of near-solidus melts of peridotite in the presence of CO2 and H2O
between 40 and 60 kbar. Lithos 112, 274–283.
French, B.M., 1966. Some geological implications of equilibrium between graphite and a
References C–H–O gas phase at high temperatures and pressures. Reviews of Geophysics 4 (2),
223–253.
Aulbach, S., Stachel, T., Creaser, R.A., Heaman, L.M., Shirey, S.B., Muehlenbachs, K., Frey, F.A., Green, D.H., 1974. The mineralogy, geochemistry and origin of lherzolite
Eichenberg, D., Harris, J.W., 2009. Sulphide survival and diamond genesis during inclusions in Victorian basanites. Geochimica et Cosmochimica Acta 38, 1023–1059.
formation and evolution of Archaean subcontinental lithosphere: a comparison Frost, D.J., McCammon, C.A., 2008. The redox state of Earth's mantle. Annual Review of
between the Slave and Kaapvaal cratons. Lithos 112 (Suppl. 2), 747–757. Earth and Planetary Sciences 36, 389–420.
Ballhaus, C., Frost, B.R., 1994. The generation of oxidized CO2-bearing basaltic melts from Goncharov, A.G., Ionov, D.A., 2012. Redox state of deep off-craton lithospheric mantle: new
reduced CH4-bearing upper mantle sources. Geochimica et Cosmochimica Acta 58, data from garnet and spinel peridotites from Vitim, southern Siberia. Contributions to
4931–4940. Mineralogy and Petrology 164, 731–745.
Ballhaus, C., Berry, R.F., Green, D.H., 1991. High-pressure experimental calibration of the Goncharov, A.G., Ionov, D.A., Doucet, L.S., Pokhilenko, L.N., 2012. Thermal state, oxygen
olivine–orthopyroxene–spinel oxygen geobarometer — implications for the oxidation fugacity and C–O–H fluid speciation in cratonic lithospheric mantle: new data on
state of the upper mantle. Contributions to Mineralogy and Petrology 107, 27–40. peridotite xenoliths from the Udachnaya kimberlite, Siberia. Earth and Planetary
Bonney, T., 1899. The parent-rock of the diamond in South Africa. Proceedings of the Science Letters 357, 99–110.
Royal Society of London 65, 223–236. Green, D.H., 1990. The role of oxidation–reduction and C–H–O fluids in determining
Boyd, F.R., Gurney, J.J., 1986. Diamonds and the African lithosphere. Science 232 (4749), melting conditions and magma compositions in the upper mantle. Proceedings of
472–477. the Indian Academy of Sciences — Earth and Planetary Sciences 99 (1), 153–165.
Boyd, F.R., Mertzman, S.A., 1987. Composition and structure of the Kaapvaal lithosphere, Green, D.H., Falloon, T.J., 1998. Pyrolite: a Ringwood concept and its current expression.
Southern Africa. In: Mysen, B.O. (Ed.), Magmatic Processes: Physicochemical Principles. In: Jackson, I. (Ed.), The Earth's mantle. Cambridge University Press, pp. 311–378.
Geochemical Society, University Park, PA, USA, pp. 13–24. Griffin, W.L., Ryan, C.G., 1995. Trace-elements in indicator minerals — area selection and
Boyd, S.R., Mattey, D.P., Pillinger, C.T., Milledge, H.J., Mendelssohn, M., Seal, M., 1987. Multiple target evaluation in diamond exploration. Journal of Geochemical Exploration 53
growth events during diamond genesis: an integrated study of carbon and nitrogen (1–3), 311–337.
isotopes and nitrogen aggregation state in coated stones. Earth and Planetary Science Griffin, W.L., Jaques, A.L., Sie, S.H., Ryan, C.G., Cousens, D.R., Suter, G.F., 1988.
Letters 86 (2–4), 341–353. Conditions of diamond growth — a proton microprobe study of inclusions in
Boyd, S.R., Pillinger, C.T., Milledge, H.J., Mendelssohn, M.J., Seal, M., 1992. C-isotopic and west Australian diamonds. Contributions to Mineralogy and Petrology 99
N-isotopic composition and the infrared-absorption spectra of coated diamonds — (2), 143–158.
evidence for the regional uniformity of CO2–H2O rich fluids in lithospheric mantle. Griffin, W.L., O'Reilly, S.Y., Natapov, L.M., Ryan, C.G., 2003. The evolution of lithospheric
Earth and Planetary Science Letters 109 (3–4), 633–644. mantle beneath the Kalahari Craton and its margins. Lithos 71 (2–4), 215–241.
Brey, G.P., Köhler, T., 1990. Geothermobarometry in four-phase lherzolites II. New Grütter, H.S., 2009. Pyroxene xenocryst geotherms: techniques and application. Lithos
thermobarometers, and practical assessment of existing thermobarometers. Journal 112 (S2), 1167–1178.
of Petrology 31, 1353–1378. Grütter, H.S., Gurney, J.J., Menzies, A.H., Winter, F., 2004. An updated classification scheme
Brey, G., Brice, W.R., Ellis, D.J., Green, D.H., Harris, K.L., Ryabchikov, I.D., 1983. for mantle-derived garnet, for use by diamond explorers. Lithos 77 (1–4), 841–857.
Pyroxene–carbonate reactions in the upper mantle. Earth and Planetary Science Gudmundsson, G., Wood, B.J., 1995. Experimental tests of garnet peridotite oxygen
Letters 62, 63–74. barometry. Contributions to Mineralogy and Petrology 119, 56–67.
Bulanova, G.P., 1995. The formation of diamond. Journal of Geochemical Exploration 53 Gurney, J.J., 1984. A correlation between garnets and diamonds in kimberlites. In: Glover,
(1–3), 1–23. J.E., Harris, P.G. (Eds.), Kimberlite Occurrence and Origin: A Basis for Conceptual
Burgess, S.R., Harte, B., 1999. Tracing lithospheric evolution trough the analysis of hetero- Models in Exploration. Publs Geol. Dept. & Univ. Extension, Univ. West. Aust., Perth,
geneous G9/G10 garnets in peridotite xenoliths, I: Major element chemistry. In: pp. 143–166.
Gurney, J.J., Gurney, J.L., Pascoe, M.D., Richardson, S.H. (Eds.), The J.B. Dawson Volume, Gurney, J.J., 1989. Diamonds. In: Ross, J., et al. (Eds.), Kimberlites and Related Rocks. GSA
Proceedings of the VIIth International Kimberlite Conference. Red Roof Design, Cape Spec Publ 14. Blackwell, Carlton, pp. 935–965.
Town, pp. 66–80. Gurney, J.J., Hildebrand, P.R., Carlson, J.A., Fedortchouk, Y., Dyck, D.R., 2004. The morpho-
Burgess, R., Layzelle, E., Turner, G., Harris, J.W., 2002. Constraints on the age and halogen logical characteristics of diamonds from the Ekati property, Northwest Territories,
composition of mantle fluids in Siberian coated diamonds. Earth and Planetary Science Canada. Lithos 77 (1–4), 21–38.
Letters 197 (3–4), 193–203. Gurney, J.J., Helmstaedt, H.H., Richardson, S.H., Shirey, S.B., 2010. Diamonds through time.
Canil, D., O'Neill, H.S.C., 1996. Distribution of ferric iron in some upper-mantle assemblages. Economic Geology 105 (3), 689–712.
Journal of Petrology 37 (3), 609–635. Haggerty, S.E., 1986. Diamond genesis in a multiply-constrained model. Nature 320
Chinn, I., 1995. A study of Unusual diamonds From the George Creek K1 Kimberlite Dyke, (6057), 34–37.
Colorado. (PhD thesis), University of Cape Town, RSA, p. 95 (146 pp.). Harley, S.L., 1984. An experimental study of the partitioning of iron and magnesium
Connell, S.H., Sellschop, J.P.F., Butler, J.E., Maclear, R.D., Doyle, B.P., Machi, I.Z., 1998. A between garnet and orthopyroxene. Contributions to Mineralogy and Petrology 86,
study of the mobility and trapping of minor hydrogen concentrations in diamond 359–373.
218 T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220

Harris, J.W., Stachel, T., Leost, I., Brey, G.P., 2004. Peridotitic diamonds from Namibia: Malaspina, N., Poli, S., Fumagalli, P., 2009. The oxidation state of metasomatized mantle
constraints on the composition and evolution of their mantle source. Lithos 77 wedge: insights from C–O–H-bearing garnet peridotite. Journal of Petrology 50,
(1–4), 209–223. 1533–1552.
Harte, B., Harris, J.W., 1994. Lower mantle associations preserved in diamonds. Mineralogical Malaspina, N., Scambelluri, M., Poli, S., Van Roermund, H.L.M., Langenhorst, F., 2010. The
Magazine 58A, 384–385. oxidation state of mantle wedge majoritic garnet websterites metasomatised by C-
Harte, B., Fitzsimons, I.C.W., Harris, J.W., Otter, M.L., 1999a. Carbon isotope ratios and nitrogen bearing subduction fluids. Earth and Planetary Science Letters 298, 417–426.
abundances in relation to cathodoluminescence characteristics for some diamonds from McCammon, C., Kopylova, M.G., 2004. A redox profile of the Slave mantle and oxygen
the Kaapvaal Province, S-Africa. Mineralogical Magazine 63 (6), 829–856. fugacity control in the cratonic mantle. Contributions to Mineralogy and Petrology
Harte, B., Harris, J.W., Hutchison, M.T., Watt, G.R., Wilding, M.C., 1999b. Lower mantle 148, 55–68.
mineral associations in diamonds from Sao Luiz, Brazil. In: Fei, Y., Bertka, C.M., McDonough, W.F., Sun, S.S., 1995. The composition of the Earth. Chemical Geology 120,
Mysen, B.O. (Eds.), Mantle Petrology: Field Observations and High Pressure Experi- 223–253.
mentation: A Tribute to Francis R. (Joe) Boyd. Special Publication. The Geochemical McLean, H., Banas, A., Creighton, S., Whiteford, S., Luth, R.W., Stachel, T., 2007. Garnet
Society, Houston, pp. 125–153. xenocrysts from the diavik mine, NWT, Canada: composition, colour, and paragenesis.
Hasterok, D., Chapman, D.S., 2011. Heat production and geotherms for the continental Canadian Mineralogist 45, 1131–1145.
lithosphere. Earth and Planetary Science Letters 307 (1–2), 59–70. Meyer, H.O.A., 1968. Chrome pyrope — an inclusion in natural diamond. Science 160
Hiraga, T., Kohlstedt, D.L., 2009. Systematic distribution of incompatible elements in (3835), 1446–1447.
mantle peridotite: importance of intra- and inter-granular melt-like components. Meyer, H.O.A., 1987. Inclusions in diamond. In: Nixon, P.H. (Ed.), Mantle Xenoliths. John
Contributions to Mineralogy and Petrology 158 (2), 149–167. Wiley & Sons Ltd., Chichester, pp. 501–522.
Holland, T.J.B., Powell, R., 2011. An improved and extended internally consistent thermo- Meyer, H.O.A., Boyd, F.R., 1972. Composition and origin of crystalline inclusions in natural
dynamic dataset for phases of petrological interest, involving a new equation of state diamonds. Geochimica et Cosmochimica Acta 36 (11), 1255–1273.
for solids. Journal of Metamorphic Geology 29, 333–383. Meyer, H.O.A., McCallum, M.E., 1986. Mineral inclusions in diamonds from the Sloan
Huizenga, J.M., Crossingham, A., Viljoen, F., 2012. Diamond precipitation from ascending Kimberlites, Colorado. Journal of Geology 94 (4), 600–612.
reduced fluids in the Kaapvaal lithosphere: thermodynamic constraints. Comptes Meyer, H.O.A., Tsai, H.M., 1976. Mineral inclusions in diamond — temperature and pressure
Rendus Geoscience 344 (2), 67–76. of equilibration. Science 191 (4229), 849–851.
Ickert, R.B., Stachel, T., Stern, R.A., Harris, J.W., 2013. Diamond from recycled crustal Moore, R.O., Gurney, J.J., 1985. Pyroxene solid solution in garnets included in diamonds.
carbon documented by coupled δ18O — δ13C measurements of diamonds and their Nature 318 (6046), 553–555.
inclusions. Earth and Planetary Science Letters 364, 85–97. Moore, R.O., Gurney, J.J., 1989. Mineral inclusions in diamond from Monastery kimberlite,
Irifune, T., Kurio, A., Sakamoto, S., Inoue, T., Sumiya, H., Funakoshi, K., 2004. Formation of South Africa. In: Ross, J., et al. (Eds.), Kimberlites and Related Rocks. GSA Spec Publ 14.
pure polycrystalline diamond by direct conversion of graphite at high pressure and Blackwell, Carlton, pp. 1029–1041.
high temperature. Physics of the Earth and Planetary Interiors 143–44, 593–600. Navon, O., Hutcheon, I.D., Rossman, G.R., Wasserburg, G.J., 1988. Mantle-derived fluids in
Jaques, A.L., Hall, A.E., Sheraton, J.W., Smith, C.B., Sun, S.-S., Drew, R.M., Foudoulis, C., diamond micro-inclusions. Nature 335 (6193), 784–789.
Ellingsen, K., 1989. Composition of crystalline inclusions and C-isotopic composition Newton, R.C., Sharp, W.E., 1975. Stability of forsterite + CO2 and its bearing on the role of
of Argyle and Ellendale diamonds. In: Ross, J., et al. (Eds.), Kimberlites and Related CO2 in the mantle. Earth and Planetary Science Letters 26, 239–244.
Rocks. GSA Spec Publ 14. Blackwell, Carlton, pp. 966–989. Nimis, P., 2002. The pressures and temperatures of formation of diamond based on
Keller, R., Taylor, L., Snyder, G., Sobolev, V., Carlson, W., Sobolev, N., Pokhilenko, N., 1998. 3-D thermobarometry of chromian diopside inclusions. Canadian Mineralogist 40 (Part 3),
petrography of diamondiferous eclogite from Udachnaya, Siberia. Seventh International 871–884.
Kimberlite Conference, Extended Abstracts, pp. 405–407. Nimis, P., Grütter, H., 2010. Internally consistent geothermometers for garnet peridotites
Kessel, R., Ulmer, P., Pettke, T., Schmidt, M.W., Thompson, A.B., 2005. The water–basalt and pyroxenites. Contributions to Mineralogy and Petrology 159 (3), 411–427.
system at 4 to 6 GPa: Phase relations and second critical endpoint in a K-free eclogite Nimis, P., Taylor, W.R., 2000. Single clinopyroxene thermobarometry for garnet
at 700 to 1400 °C. Earth and Planetary Science Letters 237 (3–4), 873–892. peridotites. Part I. Calibration and testing of a Cr-in-Cpx barometer and an
Klein-BenDavid, O., Logvinova, A.M., Schrauder, M., Spetius, Z.V., Weiss, Y., Hauri, E.H., enstatite-in-Cpx thermometer. Contributions to Mineralogy and Petrology 139
Kaminsky, F.V., Sobolev, N.V., Navon, O., 2009. High-Mg carbonatitic microinclusions (5), 541–554.
in some Yakutian diamonds — a new type of diamond-forming fluid. Lithos 112 O'Neill, H.S.C., 1980. An experimental study of the iron–magnesium partitioning between
(Suppl. 2), 648–659. garnet and olivine and its calibration as a geothermometer: corrections. Contribu-
Knoche, R., Sweeney, R.J., Luth, R.W., 1999. Carbonation and decarbonation of tions to Mineralogy and Petrology 72, 337.
eclogites: the role of garnet. Contributions to Mineralogy and Petrology 135 O'Neill, H.S.C., Wall, V.J., 1987. The olivine–orthopyroxene–spinel oxygen-geobarometer,
(4), 332–339. the nickel precipitation curve and the oxygen fugacity of the Earth's upper mantle.
Kopylova, M.G., Gurney, J.J., Daniels, L.R.M., 1997. Mineral inclusions in diamonds from Journal of Petrology 28, 1169–1191.
the River Ranch kimberlite, Zimbabwe. Contributions to Mineralogy and Petrology O'Neill, H.S.C., Wood, B.J., 1979. An experimental study of the Fe–Mg partitioning between
129 (4), 366–384. garnet and olivine and its calibration as a geothermometer. Contributions to Mineralogy
Kopylova, M., Navon, O., Dubrovinsky, L., Khachatryan, G., 2010. Carbonatitic mineralogy and Petrology 70, 59–70.
of natural diamond-forming fluids. Earth and Planetary Science Letters 291 (1–4), Otter, M.L., Gurney, J.J., 1989. Mineral inclusions in diamond from the Sloan diatremes,
126–137. Colorado–Wyoming State Line kimberlite district, North America. In: Ross, J., et al.
Kramers, J.D., 1979. Lead, uranium, strontium, potassium and rubidium in inclusion- (Eds.), Kimberlites and Related Rocks. GSA Spec. Publ. 14. Blackwell, Carlton,
bearing diamonds and mantle-derived xenoliths from southern Africa. Earth and pp. 1042–1053.
Planetary Science Letters 42 (1), 58–70. Palot, M., Pearson, D.G., Stern, R.A., Stachel, T., Harris, J.W., 2013. Multiple growth events,
Krogh, E., 1988. The garnet–clinopyroxene iron–magnesium geothermometer — a processes and fluid sources involved in diamond genesis: a micro-analytical study of
reinterpretation of existing experimental data. Contributions to Mineralogy sulphide-bearing diamonds from Finsch mine, RSA. Geochimica et Cosmochimica
and Petrology 99, 44–48. Acta 106, 51–70.
Kushiro, I., Syono, Y., Akimoto, S.I., 1968. Melting of a peridotite nodule at high Pearson, D.G., Shirey, S.B., 1999. Isotopic dating of diamonds. In: Ruiz, J., Lambert, D.D. (Eds.),
pressures and high water pressures. Journal of Geophysical Research 73 Applications of Radiogenic Isotopes to Ore Deposit Research. Economic Geology special
(18), 6023–6029. publication: SEG Reviews in Economic Geology, pp. 143–171.
Lazarov, M., Woodland, A.B., Brey, G.P., 2009. Thermal state and redox conditions of the Pearson, D.G., Davies, G.R., Nixon, P.H., Milledge, H.J., 1989. Graphitized diamonds from a
Kaapvaal mantle: a study of xenoliths from the Finsch mine, South Africa. Lithos peridotite massif in Morocco and implications for anomalous diamond occurrences.
112, 913–923. Nature 338 (6210), 60–62.
Leahy, K., Taylor, W.R., 1997. The influence of the Glennie domain deep structure on the Pearson, D.G., Shirey, S.B., Harris, J.W., Carlson, R.W., 1998. Sulphide inclusions in diamonds
diamonds in Saskatchewan kimberlites. Russian Geology and Geophysics 38 (2), from the Koffiefontein kimberlite, S Africa: constraints on diamond ages and mantle Re-
451–460. Os systematics. Earth and Planetary Science Letters 160 (3–4), 311–326.
Leost, I., Stachel, T., Brey, G.P., Harris, J.W., Ryabchikov, I.D., 2003. Diamond formation and Peats, J., Stachel, T., Stern, R.A., Muehlenbachs, K., Armstrong, J., 2012. Aviat diamonds: a
source carbonation: mineral associations in diamonds from Namibia. Contributions window into the deep lithospheric mantle beneath the northern Churchill Province,
to Mineralogy and Petrology 145 (1), 15–24. Melville Peninsula, Canada. Canadian Mineralogist 50 (3), 611–624.
Lewis, H.C., 1887. On a diamondiferous peridotite, and the genesis of the diamond. Petts, D.C., Stern, R.A., Stachel, T., Chacko, T., Heaman, L., 2014. A nitrogen isotope fraction-
Geological Magazine 4, 22–24. ation factor between diamond and fluid derived from detailed SIMS analysis of an
Litasov, K.D., Shatskiy, A., Ohtani, E., 2014. Melting and subsolidus phase relations in eclogitic diamond. V. M. Goldschmidt Conference, Sacramento, June 2014, Abstract
peridotite and eclogite systems with reduced COH fluid at 3–16 GPa. Earth and Planetary #1953.
Science Letters 391, 87–99. Petts, D.C., Stern, R.A., Chacko, T., Stachel, T., Heaman, L.M., 2015. A Nitrogen Isotope
Luth, R.W., 1993. Diamonds, eclogites, and the oxidation state of the Earth's mantle. Science Fractionation Factor Between Diamond and Its Parental Fluid Derived From Detailed
261 (5117), 66–68. SIMS Analysis of an Eclogitic Diamond (submitted for publication).
Luth, R.W., 2014. 3.9 — Volatiles in Earth's mantle. In: Holland, H.D., Turekian, K.K. (Eds.), Phillips, D., Harris, J.W., 1995. Geothermobarometry of diamond inclusions from the De
Treatise on Geochemistry, Second edition Elsevier, Oxford, pp. 355–391. Beers Pool Mines, Kimberley. Sixth International Kimberlite Conference, Novosibirsk,
Luth, R.W., Stachel, T., 2014. The buffering capacity of lithospheric mantle: implications Extended Abstracts, pp. 441–443.
for diamond formation. Contributions to Mineralogy and Petrology 168. http://dx. Phillips, D., Harris, J.W., Viljoen, K.S., 2004. Mineral chemistry and thermobarometry of
doi.org/10.1007/s00410-014-1083-6. inclusions from De Beers Pool diamonds, Kimberley, South Africa. Lithos 77 (1–4),
Luth, R.W., Virgo, D., Boyd, F.R., Wood, B.J., 1990. Ferric iron in mantle-derived garnets — 155–179.
implications for thermobarometry and for the oxidation state of the mantle. Prinz, M., Manson, D.V., Hlava, P.F., Keil, K., 1975. Inclusions in diamonds: garnet lherzolite
Contributions to Mineralogy and Petrology 104, 56–72. and eclogite assemblages. Physics and Chemistry of the Earth 9, 797–815.
T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220 219

Purwin, H., Lauterbach, S., Brey, G., Woodland, A., Kleebe, H.-J., 2013. An experimen- Stagno, V., Ojwang, D.O., McCammon, C.A., Frost, D.J., 2013. The oxidation state of
tal study of the Fe oxidation states in garnet and clinopyroxene as a function the mantle and the extraction of carbon from Earth's interior. Nature 493,
of temperature in the system CaO–FeO–Fe 2 O 3 –MgO–Al2 O 3 –SiO 2 : implications 84–88.
for garnet–clinopyroxene geothermometry. Contributions to Mineralogy and Stosch, H.G., Lugmair, G.W., 1986. Trace element and Sr and Nd isotope geochemistry of
Petrology 165 (4), 623–639. peridotite xenoliths from the Eifel (West Germany) and their bearing on the
Richardson, S.H., Harris, J.W., 1997. Antiquity of peridotitic diamonds from the Siberian evolution of the subcontinental lithosphere. Earth and Planetary Science Letters 80,
craton. Earth and Planetary Science Letters 151 (3–4), 271–277. 281–298.
Richardson, S.H., Gurney, J.J., Erlank, A.J., Harris, J.W., 1984. Origin of diamonds in old Sunagawa, I., 1984. Morphology of natural and synthetic diamond crystals. In:
enriched mantle. Nature 310 (5974), 198–202. Sunagawa, I. (Ed.), Materials Science of the Earth's Interior. Terra Scientific,
Rickard, R.S., Harris, J.W., Gurney, J.J., Cardoso, P., 1989. Mineral inclusions from Tokyo, pp. 303–330.
Koffiefontein mine. In: Ross, J., et al. (Eds.), Kimberlites and Related Rocks. GSA Tappert, R., Stachel, T., Harris, J.W., Muehlenbachs, K., Ludwig, T., Brey, G.P., 2005.
Spec. Publ. 14. Blackwell, Carlton, pp. 1054–1062. Diamonds from Jagersfontein (South Africa): messengers from the sublithospheric
Rohrbach, A., Ballhaus, C., Golla Schindler, U., Ulmer, P., Kamenetsky, V.S., Kuzmin, D.V., mantle. Contributions to Mineralogy and Petrology 150 (5), 505–522.
2007. Metal saturation in the upper mantle. Nature 449, 456–458. Taylor, W.R., Green, D.H., 1988. Measurement of reduced peridotite-C–O–H
Rohrbach, A., Ballhaus, C., Ulmer, P., Golla Schindler, U., Schönbohm, D., 2011. Experimen- solidus and implications for redox melting of the mantle. Nature 332 (6162),
tal evidence for a reduced metal-saturated upper mantle. Journal of Petrology 52, 349–352.
717–731. Taylor, W.R., Green, D.H., 1989. The role of reduced C–O–H fluids in mantle partial
Rosenhauer, M., Woermann, E., Knecht, B., Ulmer, C.G., 1977. The stability of graphite and melting. In: Ross, J., et al. (Eds.), Kimberlites and Related Rocks: Their Composition,
diamond as a function of the oxygen fugacity in the mantle. 2nd International Kimberlite Occurrence, Origin and Emplacement. GSA Spec Publ No 14. Blackwell, Carlton,
Conference, Santa Fe, p. 297. pp. 592–602.
Saguy, C., 2004. Diffusion of light elements in diamond. Defects and Diffusion in Ceramics — Taylor, L.A., Milledge, H.J., Bulanova, G.P., Snyder, G.A., Keller, R.A., 1998. Metasomatic
an Annual Retrospective VI, 226–228, pp. 31–48. eclogitic diamond growth: evidence from multiple diamond inclusions. International
Schmidberger, S.S., Simonetti, A., Francis, D., 2003. Small-scale Sr isotope investigation of Geology Review 40 (8), 663–676.
clinopyroxenes from peridotite xenoliths by laser ablation MC-ICP-MS — implications Thomassot, E., Cartigny, P., Harris, J.W., Viljoen, K.S., 2007. Methane-related diamond
for mantle metasomatism. Chemical Geology 199 (3–4), 317–329. crystallization in the Earth's mantle: stable isotope evidences from a single
Schulze, D.J., 1989. Constraints on the abundance of eclogite in the upper mantle. Journal diamond-bearing xenolith. Earth and Planetary Science Letters 257 (3–4),
of Geophysical Research-Solid Earth and Planets 94 (B4), 4205–4212. 362–371.
Scott Smith, B.H., Danchin, R.V., Harris, J.W., Stracke, K.J., 1984. Kimberlites near Orroroo, Ulmer, G.C., Grandstaff, D.E., Woermann, E., Gobbels, M., Schonitz, M., Woodland, A.B.,
South Australia. In: Kornprobst, J. (Ed.), Kimberlites I: Kimberlites and Related Rocks. 1998. The redox stability of moissanite (SiC) compared with metal–metal oxide
Elsevier, Amsterdam, pp. 121–142. buffers at 1773 K and at pressures up to 90 kbar. Neues Jahrbuch Fur Mineralogie-
Shirey, S.B., Harris, J.W., Richardson, S.H., Fouch, M.J., James, D.E., Cartigny, P., Deines, P., Abhandlungen 172, 279–307.
Viljoen, K.S., 2002. Diamond genesis, seismic structure, and evolution of the Viljoen, K.S., Dobbe, R., Smit, B., Thomassot, E., Cartigny, P., 2004. Petrology and geochem-
Kaapvaal – Zimbabwe craton. Science 297 (5587), 1683–1686. istry of a diamondiferous lherzolite from the Premier diamond mine, South Africa.
Shirey, S.B., Cartigny, P., Frost, D.J., Keshav, S., Nestola, F., Nimis, P., Pearson, D.G., Sobolev, Lithos 77 (1–4), 539–552.
N.V., Walter, M.J., 2013. Diamonds and the geology of mantle carbon. In: Hazen, R.M., Walter, M.J., Kohn, S.C., Araujo, D., Bulanova, G.P., Smith, C.B., Gaillou, E., Wang, J., Steele,
Jones, A.P., Baross, J.A. (Eds.), Carbon in EarthReviews in Mineralogy & Geochemistry A., Shirey, S.B., 2011. Deep mantle cycling of oceanic crust: evidence from diamonds
355–421. and their mineral inclusions. Science 333 (6052), 54–57.
Shiryaev, A.A., Griffin, W.L., Stoyanov, E., 2011. Moissanite (SiC) from kimberlites: Wang, A., Pasteris, J.D., Meyer, H.O.A., Deleduboi, M.L., 1996. Magnesite-bearing inclusion
polytypes, trace elements, inclusions and speculations on origin. Lithos 122, assemblage in natural diamond. Earth and Planetary Science Letters 141 (1–4),
152–164. 293–306.
Smart, K.A., Heaman, L.M., Chacko, T., Simonetti, A., Kopylova, M., Mah, D., Daniels, D., Watson, E.B., Brenan, J.M., Baker, D.R., 1990. Distribution of fluids in the continental
2009. The origin of high-MgO diamond eclogites from the Jericho Kimberlite, mantle. In: Menzies, M.A. (Ed.), Continental MantleOxford Monographs on Geology
Canada. Earth and Planetary Science Letters 284 (3–4), 527–537. and Geophysics 111–125.
Smart, K.A., Chacko, T., Stachel, T., Muehlenbachs, K., Stern, R.A., Heaman, L.M., 2011. Weiss, Y., Kiflawi, I., Davies, N., Navon, O., 2014. High-density fluids and the growth of
Diamond growth from oxidized carbon sources beneath the Northern Slave Craton, monocrystalline diamonds. Geochimica et Cosmochimica Acta 141, 145–159.
Canada: a δ13C–N study of eclogite-hosted diamonds from the Jericho kimberlite. Wiggers de Vries, D.F., Bulanova, G.P., De Corte, K., Pearson, D.G., Craven, J.A., Davies, G.R.,
Geochimica et Cosmochimica Acta 75 (20), 6027–6047. 2013a. Micron-scale coupled carbon isotope and nitrogen abundance variations in
Smit, K.V., Shirey, S.B., Richardson, S.H., le Roex, A.P., Gurney, J.J., 2010. Re–Os isotopic com- diamonds: evidence for episodic diamond formation beneath the Siberian Craton.
position of peridotitic sulphide inclusions in diamonds from Ellendale, Australia: age Geochimica et Cosmochimica Acta 100, 176–199.
constraints on Kimberley cratonic lithosphere. Geochimica et Cosmochimica Acta 74 Wiggers de Vries, D.F., Pearson, D.G., Bulanova, G.P., Smelov, A.P., Pavlushin, A.D.,
(11), 3292–3306. Davies, G.R., 2013b. Re–Os dating of sulphide inclusions zonally distributed in
Sobolev, N.V., Lavrent'ev, Y.G., Pospelova, L.N., Sobolev, V.C., 1969. Chrome pyrope from single Yakutian diamonds: evidence for multiple episodes of Proterozoic forma-
Yakutian diamonds. Doklady Akademii Nauk SSSR 189 (1), 133–136. tion and protracted timescales of diamond growth. Geochimica et Cosmochimica
Sobolev, N.V., Efimova, E.S., Pospelova, L.N., 1981. Native iron in Yakutian diamonds and Acta 120, 363–394.
its paragenesis. Geologiya i Geofizika 22, 25–29 (in Russian). Williams, A.S., 1932. The Genesis of the Diamond 2 vol. Ernest Benn Limited, London
Sobolev, N.V., Logvinova, A.M., Efimova, E.S., 2009. Syngenetic phlogopite inclusions in (636 pp.).
kimberlite-hosted diamonds: implications for role of volatiles in diamond formation. Wood, B.J., 1993. Carbon in the core. Earth and Planetary Science Letters 117 (3–4),
Russian Geology and Geophysics 50, 1234–1248. 593–607.
Stachel, T., Harris, J.W., 1997. Diamond precipitation and mantle metasomatism — evidence Wood, B.J., Bryndzia, L.T., Johnson, K.E., 1990. Mantle oxidation state and its
from the trace element chemistry of silicate inclusions in diamonds from Akwatia, relationship to tectonic environment and fluid speciation. Science 248
Ghana. Contributions to Mineralogy and Petrology 129 (2–3), 143–154. (4953), 337–345.
Stachel, T., Harris, J.W., 2008. The origin of cratonic diamonds — constraints from mineral Woodland, A.B., 2009. Ferric iron contents of clinopyroxene from cratonic mantle and
inclusions. Ore Geology Reviews 34 (1–2), 5–32. partitioning behaviour with garnet. Lithos 112 (Suppl. 2), 1143–1149.
Stachel, T., Harris, J.W., Brey, G.P., 1998. Rare and unusual mineral inclusions in Woodland, A.B., Koch, M., 2003. Variation in oxygen fugacity with depth in the upper
diamonds from Mwadui, Tanzania. Contributions to Mineralogy and Petrology mantle beneath the Kaapvaal craton, Southern Africa. Earth and Planetary Science
132 (1), 34–47. Letters 214 (1–2), 295–310.
Stachel, T., Brey, G.P., Harris, J.W., 2000a. Kankan diamonds (Guinea) I: from the lithosphere Woodland, A.B., O'Neill, H.S.C., 1993. Synthesis and stability of Fe2+ 3+ 3
3 Fe2 Si O12 garnet and
down to the transition zone. Contributions to Mineralogy and Petrology 140, 1–15. phase relations with Fe3Al2Si3O12–Fe32+Fe3+ 2 Si3O12 solutions. American Mineralogist
Stachel, T., Harris, J.W., Brey, G.P., Joswig, W., 2000b. Kankan diamonds (Guinea) II: 78 (9–10), 1002–1015.
lower mantle inclusion parageneses. Contributions to Mineralogy and Petrology Woodland, A.B., Peltonen, P., 1999. Ferric iron contents of garnet and clinopyroxene and
140 (1), 16–27. estimated oxygen fugacities of peridotite xenoliths from the Eastern Finland Kimber-
Stachel, T., Harris, J.W., Tappert, R., Brey, G.P., 2003. Peridotitic diamonds from the Slave lite Province. In: Gurney, J.J., Gurney, J.L., Pascoe, M.D., Richardson, S.H. (Eds.), The
and the Kaapvaal cratons — similarities and differences based on a preliminary data P.H. Nixon Volume, Proceedings of the VIIth International Kimberlite Conference.
set. Lithos 71 (2–4), 489–503. Red Roof Design, Cape Town, pp. 904–977.
Stachel, T., Aulbach, S., Brey, G.P., Harris, J.W., Leost, I., Tappert, R., Viljoen, K.S., 2004. The Wyllie, P.J., 1987. Metasomatism and fluid generation in mantle xenoliths. In: Nixon, P.H.
trace element composition of silicate inclusions in diamonds: a review. Lithos 77 (Ed.), Mantle Xenoliths. John Wiley & Sons Ltd., Chichester, pp. 609–621.
(1–4), 1–19. Wyllie, P.J., Huang, W.-L., 1976. Carbonation and melting reactions in the system CaO–
Stachel, T., Harris, J.W., Muehlenbachs, K., 2009. Sources of carbon in inclusion bearing MgO–SiO2–CO2 at mantle pressures with geophysical and petrological applications.
diamonds. Lithos 112 (S2), 625–637. Contributions to Mineralogy and Petrology 54, 79–107.
Stachel, T., Harris, J.W., Hunt, L., Muehlenbachs, K., Kobussen, A., 2015. Argyle diamonds — Wyllie, P.J., Ryabchikov, I.D., 2000. Volatile components, magmas, and critical fluids in
how subduction along the Kimberley Craton edge generated the world's biggest dia- upwelling mantle. Journal of Petrology 41 (7), 1195–1206.
mond deposit. Economic Geology (in press). Wyllie, P.J., Huang, W.L., Otto, J., Byrnes, A.P., 1983. Carbonation of peridotites and decarbon-
Stagno, V., Frost, D.J., 2010. Carbon speciation in the asthenosphere: experimental ation of siliceous dolomites represented in the system CaO–MgO–SiO2–CO2 to 30 kbar.
measurements of the redox conditions at which carbonate-bearing melts coexist Tectonophysics 100, 359–388.
with graphite or diamond in peridotite assemblages. Earth and Planetary Science Letters Yasuda, A., Fujii, T., Kurita, K., 1994. Melting phase-relations of an anhydrous
300, 72–84. midocean ridge basalt from 3 to 20 GPa — implications for the behavior of
220 T. Stachel, R.W. Luth / Lithos 220–223 (2015) 200–220

subducted oceanic-crust in the mantle. Journal of Geophysical Research-Solid Zhang, C., Duan, Z., 2009. A model for C–O–H fluid in the Earth's mantle. Geochimica et
Earth 99 (B5), 9401–9414. Cosmochimica Acta 73, 2089–2102.
Yaxley, G.M., Berry, A.J., Kamenetsky, V.S., Woodland, A.B., Golovin, A.V., 2012. An oxygen Zhang, C., Duan, Z., 2010. GFluid an Excel spreadsheet for investigating C–O–H fluid
fugacity profile through the Siberian Craton — Fe K-edge XANES determinations of composition under high temperatures and pressures. Computers & Geosciences 36,
Fe3+/∑Fe in garnets in peridotite xenoliths from the Udachnaya East kimberlite. 569–572.
Lithos 140–141, 142–151.

You might also like