Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/269401567

Diamond

Article · January 2014

CITATIONS READS
5 2,563

1 author:

Thomas Stachel
University of Alberta
198 PUBLICATIONS   5,704 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Unconventional settings for diamond deposits View project

MSc Thesis Project View project

All content following this page was uploaded by Thomas Stachel on 11 December 2014.

The user has requested enhancement of the downloaded file.


T. STACHEL – DIAMOND

CHAPTER 1: DIAMONDS

T. Stachel
Department of Earth and Atmospheric Sciences
University of Alberta, 1-26 Earth Sciences Building
Edmonton, Alberta
Canada, T6G 2E3
Email: tstachel@ualberta.ca

INTRODUCTION
Mineralogically, diamond is the cubic high- followed by Russia, Canada, South Africa and
pressure phase of elemental carbon characterized by Angola (Fig. 1-1). Although only about one-fourth
unique Mohs hardness (10), perfect octahedral of the world’s diamond production is of gem
cleavage, and very high density (3.51 g/cm3), quality, retail sales of diamond jewelry exceed US$
refractive index (2.417), dispersion (0.044) and 70 billion.
thermal conductivity. At the same time, history Chemical inertness, deep origin and a formation
shows that diamond is the most precious and desired probably dating back as far as the Paleoarchean also
of all gems. In modern times this status is further make diamond the object of intense research efforts.
promoted through diamond’s extreme durability, its Most of this research focuses not on diamond itself
formation deep in the Earth’s interior (>105 km) and but on minute mineral inclusions – unwanted
during Earth’s earliest history, together with blemishes in the gemstone trade – that diamond
ingenious advertising, such as the theme “a diamond incorporated during growth and then shielded from
is forever”. chemical alteration through subsequent mantle
India was the world’s only supplier of storage and exhumation from depth. It is these
diamonds for at least two thousand years until the inclusions and the phase assemblages they sample
discovery of diamonds in Brazil (1725). Diamonds that provide the only pristine samples of the deep
from the alluvial Kalimantan fields of Borneo likely mantle, while also informing about the origin of
were mined since the 7th century but either only diamond. Because of the small size of inclusions in
traded regionally or included into Indian production diamond (usually < 0.02 mm) early analytical work
before being exported to Europe (Ogden 2005). (e.g., Mitchell & Giardini 1953) was limited to
Following new discoveries in South Africa (from X-ray diffraction studies. Criteria for the visual
1866 onward) it was eventually recognized that identification of mineral inclusions in diamond were
diamond is brought to Earth's surface in volcanic first developed by Harris (1968a, 1968b). With the
rocks subsequently named “kimberlite” after the introduction of the electron probe micro-analyzer in
town Kimberley, where the new diamond-bearing the 1960s, chemical analysis of inclusions
rock was first found. Today diamond has become a eventually became possible, and early pioneering
key mineral resource with annual production in work was simultaneously undertaken at the
2011 reaching 124 million carats equivalent to Carnegie Institution, Washington (Meyer 1968,
US$14.4 billion. In the order of carats extracted, the Meyer & Boyd 1969) and at the Russian Academy
principal diamond-producing countries are Russia, of Science in Novosibirsk (Sobolev et al. 1969,
Botswana, Democratic Republic of the Congo, 1970). Milestone reviews of the early research on
Canada, Zimbabwe, Angola, South Africa and inclusions in diamond were given by Meyer & Boyd
Australia (Fig. 1-1). A factor that distinguishes (1972), Sobolev (1977), Meyer (1987) and Gurney
diamond from non-gem mineral resources is the (1989). These and subsequent studies on mineral
variability of carat price depending on production inclusions in diamond have fundamentally shaped
quality; end-members in the extreme fluctuation our understanding of the geochemical and
about an average production price of US$ 116 per mineralogical environment of diamond formation in
carat extend from US$ 9 (Democratic Republic of the Earth’s mantle and, thereby, laid the
the Congo) to US$ 1602 per carat (Kingdom of groundwork for a modern interpretation of indicator
Lesotho). Based on the value of production in 2011, mineral analyses during diamond exploration.
Botswana is the world’s leading diamond producer,

Mineralogical Association of Canada Short Course 44, Tucson AZ, February 2014, p. 1-28.

1
T. STACHEL – DIAMOND

35.14

30 
Million carats 22.90
19.25
20 

10.80
7.83 8.21 8.33 8.50
10 

0.36 1.26

eo ra
Na e

fri th

ia
e

a
da
Au a
lia

go
la
ca
n

an
bw
i

ss
   L ier

ib

go
   A u
ra

na

on
So

Ru
w
m

ba
An
st
S

Ca

 C

ts
m

DR

Bo
Zi
1,602
1,500 
US$ per carat

1,000 
695

500  348
170 211 236
76 140
9 28 56

la
fri th 

eo ra 

o
Na ne
o

ia

ia
e
Zi alia

C ca

go
an

ad

th
bw
ng

   L ier
   Aou

ib
ss

so
r

an

m
o

Ru

S
ba

A
st
 C

ts

Le
Au

m
DR

Bo

3,902
4,000 

3,000  2,551 2,675


Million US$

2,000  1,730
1,163
873
1,000 
359 476
180 221

o
fri th 

ng
o

ia

a
e
lia

ia

la

da
Ca a

an
bw
th

c
ib

ss
go

   A u
ra

o
na
So
so

w
m

Ru

 C
An
ba
st

ts
Le

DR
Na
Au

Bo
Zi

FIG.1-1. Top: the world’s top ten diamond-producing countries in terms of million carats extracted. Middle: average
production price (US$ per carat) for selected countries. Bottom: the world’s top ten diamond-producing countries in
terms of value (million US$). The Kingdom of Lesotho, which only ranks #14 in production measured by weight,
jumps to #8 in production measured by value due to an extremely high carat price. Data source: Kimberley Process
Rough Diamond Production Statistics for 2011 (https://kimberleyprocessstatistics.org/public_statistics).

2
T. STACHEL – DIAMOND

ORIGIN OF DIAMOND IN EARTH’S presumably related to redox fronts. A strong


MANTLE association between diamonds and harzburgite/
Primary diamond deposits are hosted in dunite (Gurney 1984) suggests that highly depleted
volcanic diatremes (“pipes”) and small subvolcanic lithospheric mantle effectively acts as a trap for
intrusions (dykes and sills), as opposed to secondary carbon in upwelling fluids or melts. In accordance
deposits that result from erosion of primary deposits with the original proposal of Haggerty (1986),
and transport and re-sedimentation of diamonds in recent evidence for a reduced character of deep
rivers and along oceanic coastlines. The growth of cratonic peridotite (Woodland & Koch 2003,
diamonds is unrelated to their volcanic host rocks. McCammon & Kopylova 2004, Creighton et al.
The essential role of magmatism in the formation of 2009, 2010) is consistent with an oxidized character
primary diamond deposits lies instead in the (i.e., carbonate-bearing) of diamond-precipitating
sampling of diamonds pre-existing at depth in fluids. Such oxidized fluids cannot be derived from
Earth’s mantle and their rapid transport to Earth’s deep convecting upper mantle as it is even more
surface. This restricts the occurrence of primary reduced than the cratonic lithosphere (Ballhaus
diamond deposits to only a few types of ultramafic 1992) but could relate to devolatilization of
magmatic rocks originating beneath the graphite– subducting oceanic slabs. The alternative model,
diamond transition (> 105–135 km, depending on oxidation of methane-bearing asthenospheric fluids
the local geotherm): Group I (“basaltic”) kimberlite, to water and diamond during ascent along a gradient
as well as less common Group II (“micaceous”) of increasing oxygen fugacity, is, however, equally
kimberlite (also named orangeite) and olivine viable (Taylor & Green 1989, Huizenga et al. 2012).
lamproite. In addition, volcaniclastic komatiite is The second principal diamond source rock in the
described as the source rock for an unusual suite of cratonic lithosphere is eclogite. It has been
diamonds at Dachine, French Guiana (Capdevila et suggested that eclogite and peridotite in the
al. 1999). subcratonic lithospheric mantle are similarly
The observed close association between reduced (Simakov 2006), which may imply a similar
diamondiferous kimberlite and Archean cratons mechanism of diamond formation. However,
(known as “Clifford’s Rule”, as revised by Janse accurate estimates of oxygen fugacity in eclogite are
1994) was related by Boyd & Gurney (1986) and still lacking since their mineralogical composition
Haggerty (1986) to deep reaching (down to ca. 200 precludes constraining silica activity (an essential
km) and chemically highly depleted lithospheric parameter for available oxybarometers).
keels that are generally absent beneath younger “Super-deep” diamonds derived from the
crust. Rigid and largely stripped of heat-producing convecting mantle (asthenosphere to lower mantle)
elements (such as potassium), subcratonic underlying the common diamond sources in ancient
lithospheric mantle is cooler than convecting and rigid cratonic lithosphere appear to be generally
asthenospheric mantle at similar depth (e.g., associated with recycled oceanic lithosphere
Rudnick & Nyblade 1999), causing the graphite– transporting carbon (as organic matter or carbonate)
diamond transition to rise to shallower depth (ca. back into Earth’s deep mantle and possibly also
105–135 km, see Fig. 1-2) beneath cratons. The providing steep redox gradients (Stachel 2001). The
region where lithospheric mantle reaches into the need for rapid exhumation appears to constrain
diamond stability field establishes a “diamond occurrences of sublithospheric diamonds to regions
window”, with the exact thickness of this window with deep-reaching cratonic keels as well. Special
varying among and within cratons depending on cases of (non-economic) primary diamond
both lithospheric thickness and local geothermal occurrences unrelated to cratons and volcanic hosts
gradient. The restriction of primary diamond are established through the presence of micro-
deposits to lithosphere extending into the diamond diamonds in UHP terranes (e.g., Sobolev & Shatsky
stability field relates to a need for rapid ascent of 1990, Zhang et al. 1997), the occurrence of graphite
magmas from the diamond window, facilitated pseudomorphs after diamond in peridotite massifs
through the brittle nature of lithospheric mantle, to (Pearson et al. 1995), and a common association of
preserve diamonds (e.g., Eggler 1989). In addition, nano- to micro-diamonds with astroblemes (e.g.,
carbon is only a trace element in Earth’s mantle Carlisle & Braman 1991).
(Trull et al. 1993) and formation of macro-diamonds
hence requires local enrichment of carbon

3
T. STACHEL – DIAMOND

30
100

Mantle Adiabat
(TP=1300°C)
40

Gr
Diamond a
Dia phit 150
P 50 Window mo e Depth
nd
[kbar] [km]

60
200

70
600 800 1000 1200 1400
T [ºC]
FIG. 1-2. Conditions of formation and storage of diamond in deep-reaching cratonic keels. Data points (grey squares)
represent pressure–temperature estimates for garnet lherzolite xenoliths from the central Slave Craton (Lac de Gras area;
Grütter 2009) using the single clinopyroxene geothermobarometer of Nimis & Taylor (2000). A linear regression of the
data set (“local geotherm”) is shown as a solid grey line. The graphite–diamond transition (dashed line) is from Day (2012).
Compared to the graphite–diamond transition originally proposed by Kennedy & Kennedy (1976), the steeper slope of the
revised phase boundary (Day 2012) extends the stability of diamond in Earth's mantle to shallower depth, with the
difference being particularly large for areas with a “cool” local geotherm (about 30 km difference for the example shown
here). The adiabat for a mantle potential temperature (TP) of 1300°C (dotted line) is from Rudnick & Nyblade (1999) and
represents the lower end of a range of temperatures (TP of 1300–1400°C) considered by these authors; a TP of ~1300°C is,
however, likely realistic for convecting mantle beneath cratons not affected by diapirs carrying excess heat from the lower
mantle (e.g., McKenzie & O’Nions 1991). Lithospheric mantle reaches its maximum possible depth where the local
geotherm intersects the mantle adiabat, implying that for geothermal gradients not significantly below that derived for the
Slave Craton, lithospheric thickness cannot considerably exceed 200 km depth (cf. Rudnick & Nyblade 1999). For the
central Slave Craton, the lithosphere extends into the diamond stability field between ca. 105–195 km depth, establishing a
diamond window of approximately 90 km width. An average geotherm for Precambrian shields worldwide (40 mW/m2
surface heat flow; Hasterok & Capman 2011) results in a diamond window of 65 km width (130–195 km depth, see Fig.
1.15).

DIAMOND MORPHOLOGY the two (cubo-octahedra), aggregates and twins. The


The following summary of diamond most common twins are flat triangles called macles
morphology and surface features is built on (spinel-law twins). Variations in growth regime are
comprehensive reviews presented by Robinson reflected by the occurrence of (i) single crystalline
(1979), Sunagawa (1984), Harris (1987), Gurney diamonds, including twins or aggregates of a few
(1989), and Wilks & Wilks (1991). Photographic crystals, (ii) coated diamonds, i.e., single crystalline
examples, largely based on the De Beers Canada diamonds with fibrous overgrowths, (iii) fibrous
Rough Diamond Display at the University of cuboid (no flat {001} faces) diamonds, and (iv)
Alberta, can be found in Tappert & Tappert (2011). polycrystalline diamond aggregates. Polycrystalline
diamond derived from primary volcanic hosts is
Primary and secondary shapes generally referred to as bort (or boart); a fine-
Primary growth forms of diamond are octahedra grained whitish variety (common at Orapa and
and cuboids (Fig. 1-3), including combinations of Venetia mines) is called framesite. Carbonado is a

4
T. STACHEL – DIAMOND

FIG. 1-3. Primary diamond shapes. From left to right: (i) octahedron, (ii) macle (spinel twin), (iii) cube with minor
dodecahedral resorption, (iv) re-entrant cube with concave {001} faces (after Welbourn et al. 1989).
porous polycrystalline form of diamond that cannot surfaces (giving 24 faces in all), based on the
(yet?) be linked to volcanic sources and may have presence of a curved crest line. In all cases of
an impact origin. resorption, true crystal shapes are not produced
Experimental work suggests that the primary because the resulting ‘faces’ are curved; planar
diamond shape changes from cube through cubo- crystal faces are, however, necessary to define a
octahedron to octahedron with increasing crystal shape crystallographically. Observations of
crystallization temperature (Yamaoka et al. 1977). pseudo-hemimorphic (“half resorbed”) diamonds –
The transition from smooth dislocation growth via with their octahedral portion being protected inside
smooth hopper growth (producing concave crystal mantle xenoliths and the dodecahedral portion being
faces) for single crystal diamonds to rough growth exposed on the surface – link the development of
for fibrous and polycrystalline diamonds reflects rounded dodecahedra to resorption in the
increasing supersaturation and growth rate transporting kimberlite or lamproite magma.
(Sunagawa 1984). Consequently, the thin fibrous Detailed studies on the internal growth structure of
coats over gem quality, single crystal diamond diamonds have also documented earlier stages of
typical of the productions from Ekati, Diavik and resorption followed by re-growth (e.g., Harte et al.
Snap Lake, for example, represent two distinct 1999a), implying that dissolution of diamond may
stages of diamond precipitation, early stable already occur during mantle residence.
dislocation growth and a second stage of rough
overgrowth at high levels of supersaturation. Common surface features
Resorption of diamond leads to the Stacked growth layers in the form of triangular
development of rounded dodecahedral faces at the plates on octahedral surfaces (both octahedra and
expense of primary cube and octahedral faces (Fig. macles) are the only growth-related surface features
1-4). With cubes a distinct tetrahexahedral shape observed on diamonds (Fig. 1-5). Incipient
results, whereas with octahedra the resulting form is resorption produces a similar secondary feature on
generally referred to as “rounded dodecahedra” (or octahedral faces called shield-shaped laminae.
“dodecahedroids”). Robinson (1979) also suggested There, resorption exhumes steps in growth layering,
the term “tetrahexahedroid” because each rounded with the resulting stacked laminae being generally
dodecahedral face may be subdivided into two thinner than primary triangular plates and showing

FIG. 1-4. Secondary shapes through increasing degrees of resorption (after Robinson 1979). From left to right: (i) sharp-edged
octahedron (100% mass preserved), (ii) octahedron with dodecahedral faces, (iii) rounded dodecahedron with residual
octahedral faces, (iv) rounded dodecahedron (less than 55% of original mass preserved).

5
T. STACHEL – DIAMOND

Plastic
Stacked deformation Edge
Hexagons
growth lines Terraces abrasion
layers
Corrosion
sculpture
Microdisk
patterns Scratch
mark
Positive Elongate
Negative Shallow hillocks Percussion Abraded
trigons surface
trigons depressions marks

FIG. 1-5. Surface features on diamond. Left: surface features on octahedral faces. Center: surface features on rounded
dodecahedral faces. Right: surface features typically associated with diamond from alluvial deposits (unrestricted to a
particular type of face).
rounded edges and blunt points. If points are sharp relate to the wear of diamonds during sedimentary
and jagged, the term serrate laminae is used instead. transport.
“Ribbing” describes the imbrication of exhumed
growth layers along rounded edges of octahedral OPTICAL CENTERS
crystals. As a common secondary etch feature, Because of similar ionic radius and charge,
octahedral faces may show triangular pits, called nitrogen substitutes for carbon in the diamond
trigons (Fig. 1-5) that are either flat-bottomed or lattice and is the most abundant impurity in
point-bottomed with inclined sides. Rare positive diamond. Nitrogen contents are generally low
trigons parallel the orientation of the octahedral (median value of 160 at.ppm (atomic ppm); <1% of
face, whereas common negative trigons show an samples exceed 1400 at.ppm) and tend to be higher
inverse orientation. Hexagonal etch pits are a in diamonds with eclogitic inclusions (median of
combination form of positive and negative trigons 494 at.ppm, see Figure 1-6) compared to peridotitic
(Fig. 1-5). The etch features equivalent to trigons on samples (median of 82 at.ppm). Both the overall
cube faces are tetragons, with rare positive differences between eclogitic and peridotitic
orientations parallel to cube edges and common diamond and the distinct zoning in N content seen
negative orientations, set at 45° to those edges. on the level of individual diamond crystals may be
Shape and orientation of trigons and tetragons due to kinetic factors related to the speed of
reflect the combined effect of temperature, oxygen diamond growth (Boyd et al. 1994) or may reflect
fugacity and chemical composition of the fluid variations in the availability of N acting as a
during etching (Fedortchouk & Canil 2009, compatible trace element during diamond
Fedortchouk & Zhang 2011). precipitation (Stachel et al. 2009). Optical centers
Dodecahedral resorption may exhume due to N in diamond are reviewed in detail by
octahedral growth layering in the form of terraces Harris (1987) and Wilks & Wilks (1991). A simple
and as discontinuous ridges called “elongate guide to the determination of the content and
hillocks” (Fig. 1-5). Closely spaced subtle sharp aggregation state of N in diamond using Fourier
parallel ridges (“plastic deformation lines”) reflect transform infrared (FTIR) spectroscopy may be
strain during diamond’s residence in the mantle. found in Taylor (2004).
“Microdisk patterns” (circular shallow positive Nitrogen enters the diamond lattice as a single
relief features), deep elliptical to irregular depress- substitutional impurity (C center) which under high
ions (“corrosion sculpture”), and less pronounced ambient temperatures in Earth’s mantle rapidly
“shallow depressions” (representing incipient aggregates to form pairs of N atoms (A center).
corrosion sculpture) likely reflect late stage etching, During continued mantle residence the N pairs may
since they have been linked to diamonds recovered aggregate further into groups of four atoms
from hypabyssal (i.e., magmatic) kimberlite in surrounding a vacancy (B center). Increasing
diatreme root zones (e.g., Robinson et al. 1989). aggregation of N in diamond from the A to the B
For alluvial diamonds, resorbed center is usually accompanied by a simultaneous
dodecahedral morphologies generally dominate over increase in N3 (groups of three N atoms surrounding
octahedral shapes. Percussion marks, ground edges a vacancy) and associated N2 centers, and B' centers
and surfaces (Fig. 1-5), and similar abrasion marks (platelets, nanometre- to micrometre-sized disc

6
T. STACHEL – DIAMOND

0 500 1000 1500 relationship between platelets and B centers likely


does not exist (Howell et al. 2012); catastrophic
Eclogitic platelet degradation may thus be erroneously
80 N=598 assigned to externally octahedral diamond
Median=494 containing early stages of cuboid growth.
Type II The thermodynamic parameters governing the
60 interrelationships of N content, N aggregation,
mantle residence time, and mantle residence
temperature were derived by Evans & Harris (1989)
40 and Taylor et al. (1990). Under mantle conditions
aggregation of single substitutional N atoms into A
20 centers occurs so quickly (Taylor et al. 1996) that it
may be ignored in a thermo-chronometric analysis.
Frequency

The second step, the aggregation from A to B


0 centers occurs at a much lower rate and thus, in
principle, it is possible to constrain either the mantle
400 Peridotitic residence time or the time-averaged residence
N=986 temperature of a diamond from the concentration
Median=82 and aggregation of N impurities. Estimates of
300 mantle residence temperatures are fairly insensitive
Type II to the choice of residence time (based on values
generally >1 Ga, see Fig. 1-7) and normally show
200 satisfactory agreement with independent estimates
based on inclusion geothermobarometry (e.g., Leahy
& Taylor 1997). Conversely, N-based geochrono-
100 metry is extremely temperature-dependent and
hence may yield very crude estimates of mantle
residence times only (e.g., billions versus millions
0 of years). Plastic deformation is common in
0 500 1000 1500 diamond and has been suggested to accelerate N
N [atomic ppm]
aggregation (Evans 1992). Considering the
FIG. 1-6. Nitrogen content in lithospheric diamonds dominant control of temperature on N aggregation,
containing mineral inclusions of peridotitic (bottom) the additional error on residence temperature
and eclogitic (top) paragenesis. Twelve samples estimates introduced through possible unaccounted
(three peridotitic, nine eclogitic) fall outside the effects of plastic deformation probably lies within
diagram with N >1500 at.ppm (maximum content is the accuracy limits of conventional, inclusion-based
3830 at.ppm). Class 0–50 at.ppm includes Type II geothermobarometry.
diamonds (i.e., diamonds with N below the limit of Other important optical centers in diamond
detection at ~10 at.ppm) shown in black.
include: (i) substitution of boron, responsible for
shaped defects, which are composed of carbon blue coloration and electrical conductivity (p-type
interstitials that originate during the formation of semiconductivity) of diamond; (ii) lattice defects
vacancies associated with nitrogen B centers). For (clusters of around 60 vacancies) due to plastic
“regular” octahedral diamonds, infrared spectra deformation giving rise to brown coloration (Fisher
show a linear relationship between the integrated 2009). For a review of color in diamond, see
strength of the platelet peak (I[B']) and the intensity Fritsch (1998), and for brown diamonds, see
of absorption (measured as absorption coefficient at Hainschwang et al. (2005) and Jones (2009).
1282 cm–1) due to the B center (Woods 1986). For
“irregular” octahedral specimens this relationship Diamond “types” and their relative abundances
breaks down due to “catastrophic” platelet Nitrogen-bearing diamonds (i.e., Type I
degradation likely related to intense deformation diamonds) may be classified based on their N
(Woods 1986), short-lived heating events (Evans et aggregation state into Type Ib (single substitutional
al. 1995), or a combination of both during mantle N) and Type Ia (aggregated). Type Ia diamonds may
residence. For cuboid growth sectors a “regular” be further categorized as Type IaA (90–100% A

7
T. STACHEL – DIAMOND

10000
300 IaA Type IaAB IaB a Peridotitic

10 a
°C
3G

°C

0°C
1G
00

50
Eclogitic
11

115
°C
00
11
Eclogitic (N=549)

0°C
1000
Frequency

200

120
Peridotitic (N=737) N

100 100

C
°C
50

00°
12

13
0 10
0 20 40 60 80 100 0 20 40 60 80 100
%B [100B/(A+B)] %B [100B/(A+B)]
FIG. 1-7. Nitrogen aggregation expressed as the relative proportion of N in the B center (assuming that all N is present in
either A or B centers). Diamonds with N <10 ppm (Type II) are excluded since their N aggregation state is unknown. The
histogram on the left shows the relative abundance of Type IaA, IaAB and IaB diamonds. The graph on the right depicts
the relationship between N content, N aggregation, mantle residence time (1 Ga and 3 Ga) and time-averaged residence
temperature (isotherms calculated after Leahy & Taylor 1997). It is apparent that the precise choice of a mantle residence
time is almost irrelevant as long as ages much older (>> 100 Ma) than kimberlite emplacement are assumed. Diamonds
cover a large range in residence temperatures from <1050 to >1300°C.
center) and Type IaB (90–100% B center) with Type Ib
intermediate types being designated as Type IaAB C center (single N)
(see Fig. 1-6 and 1-8). Type II diamonds are defined (deep yellow to brown)
as being nitrogen-“free”, i.e., having N contents
below the limit of detection of FTIR spectroscopy at Type Ib-IaA
about 5–10 at.ppm. There is a further subdivision of
Type II diamonds into boron-free (Type IIa) and Type IaA
boron-bearing samples (Type IIb). A detailed study A center
of blue Type IIb diamonds (Gaillou et al. 2012) (colourless)
showed color intensity to correlate only loosely with
B content (ranging from <0.08 to 8.4 ppm in ToF– Type IaAB
SIMS spot analyses). The B content of Type I +
diamonds is not well studied since N compensates B’ center (platelets)
the B-related centers which therefore cannot be (colourless)
detected using optical absorption spectroscopy. A +
single Type IaAB diamond included in the study of N2 and N3 centers
Gaillou et al. (2012) had no detectable B (limit of (pale to bright yellow)
detection during ToF–SIMS analysis was 0.3 ppm).
Considering the natural abundance of the
various diamond types it is useful to separate
“normal” diamonds derived from the subcratonic Type IaB
lithospheric mantle from “super deep” B center
sublithospheric diamonds. For diamonds containing (colourless)
lithospheric inclusions a world-wide data-base FIG. 1-8. Flow diagram illustrating increasing aggreg-
(Table 1-1) shows 19% of the samples to be Type II, ation of N in Type I diamonds. Nitrogen C and N2
a much higher proportion than the commonly quoted centers cause various shades (depending on
number of 2% (Dyer et al. 1965). Peridotitic concentration) of yellow body colors.

8
T. STACHEL – DIAMOND

diamonds are three times as likely to be Type II as STABLE ISOTOPIC COMPOSITION OF


eclogitic diamonds (Table 1-1). Nitrogen-bearing DIAMONDS
diamonds are dominated by intermediate Apart from minor elemental and molecular
aggregation states (Type IaAB, see Fig. 1-7). impurities at the trace level, diamond is composed
Compared to the eclogitic suite, peridotitic of pure carbon. Hence, the C stable isotopic
diamonds have a higher proportion of the lesser signature of diamond may be expected to provide
aggregated Type IaA diamonds, which is a function the most robust fingerprint for its origin (see Kirkley
of lower average N contents (Fig. 1-6) leading to et al. 1991, Cartigny 2005 and Stachel et al. 2009
slower aggregation at comparable residence for full reviews). Carbon has two stable isotopes,
12
temperatures. Fully aggregated Type IaB diamonds C and 13C and the ratio of the two is
only form a minor portion of the world-wide data conventionally expressed as a normalized difference
base (Table 1-1). Diamonds containing some non- (δ13C measured in ‰) with respect to the
aggregated N (Type Ib–IaA) are extremely rare but international VPDB (Vienna Pee Dee Belemnite)
make up about 5% of the production at the Helam standard.
Mine (Swartruggens dyke swarm, South Africa; Figure 1-9 shows the C isotopic composition of
McKenna et al. 2004). Bluish Type IIb diamonds diamonds carrying inclusions that are derived from
are very rare and inclusion studies have not been the subcratonic lithospheric mantle (peridotitic,
completed on this kind of diamond. World-wide, eclogitic and websteritic suites), from the
Type IIb diamonds are mainly derived from the underlying asthenosphere and transition zone
Cullinan Mine (“Premier”) in South Africa, and a (eclogitic and websteritic garnet inclusions with a
likely lithospheric origin is inferred by association significant majorite component) and from the lower
with the general diamond production at that mine. mantle (various associations involving the mineral
Diamonds containing inclusion parageneses ferropericlase). Peridotitic, eclogitic, websteritic and
indicative of derivation from the lowermost lower mantle diamonds all have modes that coincide
transition zone and the lower mantle typically are with the accepted mantle value of about –5‰. This
Type II (Hutchison et al. 1999). Rare exceptions are concurrence, however, cannot be used to
several ferropericlase-bearing diamonds from the conclusively establish a mantle origin for most
Lac de Gras area (Davies et al. 1999, Tappert et al. diamond C, as the C isotopic composition of
2005a, Donnelly et al. 2007) containing 16 to 1950 average altered oceanic crust is indistinct (about –
at.ppm N. This may indicate that these diamonds 4.7‰, Shilobreeva et al. 2011). For eclogitic and
actually represent ferropericlase-bearing dunitic websteritic diamond the distributions are strongly
sources in the lithosphere (Brey et al. 2004) rather skewed towards more negative values (down to –
than a sublithospheric origin. Majorite garnet- 41.1 and –41.3‰, respectively, De Stefano et al.
bearing diamonds derived from the asthenosphere 2009). An apparent secondary mode for eclogitic
and transition zone are either Type II or have low N diamond at –13 to –9‰ (Fig. 1-9) relates to
contents at intermediate to high aggregation states abundant isotopic data for a single diamond
(Type IaAB to IaB). For example, eleven diamonds occurrence (Argyle Mine in Western Australia). For
with eclogitic majorite garnet inclusions from asthenospheric and transition zone diamonds
Jagersfontein, South Africa, have an average N Tappert et al. (2005c) pointed out that each locality
content of 24 at.ppm (Tappert et al. 2005c). studied so far has yielded a distinctive and narrow
range in C isotopic compositions, with values below
TABLE 1-1: PERCENTAGES OF TYPES (BASED ON N –18‰ representing diamonds from the former
CONTENT AND AGGREGATION STATE) FOR Jagersfontein mine (South Africa) only.
LITHOSPHERIC DIAMOND WITH PERIDOTITIC AND The association of isotopically light C with
ECLOGITIC INCLUSIONS eclogitic (and comparatively rare websteritic)
diamond was first noticed by Sobolev et al. (1979).
Peridotitic Eclogitic Combined Based on similarities with the C isotopic
Type n=986 n=598 n=1584 composition of organic matter, these strongly
II 25 8 19 negative values have since been widely used as
IaA 19 16 18 evidence for a subduction origin of eclogitic
IaAB 52 72 59 diamonds (e.g., Sobolev & Sobolev 1980, Kirkley et
IaB 3 4 4 al. 1991, McCandless & Gurney 1997). Various
arguments have, however, been forwarded against a

9
T. STACHEL – DIAMOND

-40 -30 -20 -10 0 direct relationship between diamonds with light C
isotopic compositions and recycled organic matter
Lower Mantle
60 N=146 (see e.g., Deines 2002). Clearly the strongest
Median=-4.5‰ challenge comes from the isotopic analysis of N
impurities in diamonds with a C isotopic
Frequency

40 “subduction signature” (see Cartigny 2005 for a


review). Nitrogen in oceanic sediments has a
positive isotopic signature (average δ15N of about
20 +6‰) and metamorphic overprint in a descending
slab will only increase this value (Cartigny et al.
1998). The N isotopic composition of Earth’s
0 mantle is estimated to be about –5 ±3‰ (Cartigny
Asthenosphere and Transition Zone 2005). Consequently, if diamond with strongly
5 N=46
Median=-13.6‰ negative C isotopic composition had formed from
4
subducted organic matter, then C isotopic
Frequency

compositions below about –10‰ should be


3 accompanied by positive δ15N values. As shown in
multiple papers by Cartigny (op. cit. and references
2 therein) this is not the case and ca. 70% of eclogitic
diamonds have negative δ15N values instead.
1
To reconcile the N and C isotopic composition
0
of diamond, Cartigny et al. (2001) favor
Eclogitic and Websteritic metasomatic diamond formation involving open
N=923 Webst. system isotopic fractionation. Considering the tight
Median eclog.=-6.3‰
150
Median webst.=-7.3‰ distribution of peridotitic diamonds around the
Eclog.
mantle value, this fractionation process must operate
Frequency

in eclogitic environments only. At the high


100
temperature conditions of the deep lithospheric
mantle strong fractionation of stable isotopes
50 generally involves open systems and for C isotopes
1 Webst. requires separation of a free CO2 fluid. At high
1 Eclog.
pressure in a peridotitic environment CO2 reacts
0 with forsteritic olivine to form magnesite and
Peridotitic enstatite, hence precluding the presence of a free
N=1259
300
Median=-4.9‰ CO2 fluid (Wyllie & Huang 1976). In eclogite,
olivine is absent and possible carbonation reactions
Frequency

involve clinopyroxene (Luth 1995) or garnet


200 (Knoche et al. 1999). For natural mineral
compositions and in the likely presence of H2O in a
CO2-bearing fluid, these buffer reactions may well
100
only occur at depth beneath the subcratonic
lithospheric mantle. The model of Cartigny et al.
0 (2001) thus provides an elegant explanation for the
-40 -30
13
-20 -10 0 (almost complete) absence of peridotitic diamonds
δ C [‰] with C isotopic compositions that are significantly
FIG. 1-9. Carbon isotopic composition (δ13C) of diamond lighter than the mantle value. Additional evidence
derived from the subcratonic lithosphere (with for a formation of eclogitic diamonds during
peridotitic, eclogitic and websteritic inclusions), the fluid/melt infiltration processes, rather than via
asthenosphere and transition zone (containing eclogitic direct conversion of subducted organic matter,
and websteritic garnet inclusions with a majorite comes from observations of complex growth
component), and the lower mantle (ferropericlase- patterns, non-systematic isotopic heterogeneity and
bearing inclusion parageneses). Arrow indicates the strong chemical disequilibrium among multiple
assumed mantle value for δ13C of about –5‰ (e.g.,
inclusions in individual diamonds (e.g., Rickard et
Cartigny 2005).

10
T. STACHEL – DIAMOND

al. 1989, Sobolev et al. 1998), and significant epigenetic inclusions. Protogenetic inclusions are
variations in the stable isotopic composition of pre-existing minerals that are passively overgrown
multiple diamonds recovered from individual by newly formed diamond. Consequently, proto-
eclogite xenoliths (e.g., Taylor et al. 2000). genetic inclusions are recognized by an absence of
However, as pointed out by Smart et al. (2011), if a both epitaxy and external morphologies imposed by
mantle-like C isotopic composition of –5‰ is the host diamond. Protogenetic inclusions are
assumed for the primary metasomatic fluid/melt, presumed to be rare but have been described as
then the CO2 separation model cannot account for seeds for diamond precipitation (Bulanova 1995).
eclogitic diamonds with δ13C much below –14‰ as Syngenetic inclusions co-precipitated with their
this would require unrealistically high fractionation diamond hosts either from a fluid/melt or during a
values (separation of >99% CO2). The Cartigny et metamorphic recrystallization event. Because of a
al. (2001) model, therefore, cannot be used to relate very high crystal form energy diamond imposes its
the origin of diamonds with very low δ13C to own symmetry on simultaneously growing inclusion
mantle-derived C and the need for input of a minerals, resulting in cubo-octahedral external
strongly 13C depleted component via subduction morphologies for the vast majority of inclusions
persists. (including non-cubic minerals). Epigenetic
Irrespective of the cause, as light C isotopic inclusions commonly are alteration products of pre-
compositions are fairly common among eclogitic existing inclusions formed when fluids enter
diamonds but rare among peridotitic diamonds (only inclusion cavities along cracks and are usually
1.3% of peridotitic diamonds have δ13C below – related to ascent and cooling of the transporting
10‰) they may be employed as a useful indicator kimberlite magma or subsequent sedimentary and
for the presence of eclogitic diamonds in population metamorphic events. Epigenetic inclusions are
studies where mineral inclusions in diamond cannot commonly cryptocrystalline or represent phases that
be recovered to establish their paragenesis. clearly are not of mantle origin (e.g., crystallized
during metamorphic events). However, the presence
INCLUSIONS IN DIAMOND of annealed cracks in diamonds, revealed in
In the gemmology of diamond, inclusions are cathodoluminescence studies, suggests that in rare
negative features that detract from value; studies on cases epigenetic modification of inclusions may
mineral inclusions in diamond have, however, already occur in the Earth’s mantle (e.g., Kopylova
played a key role in establishing the xenocrystic et al. 1997).
nature of diamonds with respect to their volcanic For studies on the physical and chemical
hosts and in deriving the mineralogical and environment of diamond formation common
chemical composition of diamond source rocks in syngenetic inclusions are targeted, although, in
Earth’s mantle. Because these inclusions can only principle, protogenetic phases that re-equilibrated
be accessed by breaking or cutting of the host with their environment until the moment of
diamonds, the majority of studies have focused on encapsulation – which is not unlikely considering
small, low values stones (typically in the 2–3 mm their small (micrometre) size and high ambient
size range). It is, however, believed that the temperatures – will yield identical information.
information retrieved from the smaller diamonds As a first order division, mineral inclusions in
extends to larger gem-quality stones. The concept of diamond can be assigned to sources within the
using Cr-rich harzburgitic garnet (G10) as a key subcratonic lithospheric mantle or may have a
indicator for diamond potential during kimberlite sublithospheric origin. Because of large scientific
exploration is based on its prevalence among interest, diamonds with sublithospheric inclusions
peridotitic garnet inclusions in diamond (Gurney & are highly over-represented in the literature and in
Switzer 1973, Gurney 1984, see below). Similarly, reality probably account for less than 1% (by mass)
spinel recovered during indicator mineral sampling of the current world-wide diamond production.
is employed to assess diamond potential by Lithospheric diamonds are subdivided into three
comparing how closely it matches the composition main suites representing peridotitic, eclogitic and
of Mg-chromite inclusions in diamond with respect rare websteritic sources (Fig. 1-10). Sublithospheric
to Mg, Cr and Ti content (e.g., Fipke et al. 1995). diamonds may be broadly split into two groups: (i)
Depending on when inclusion minerals formed asthenospheric and transition zone diamonds with
relative to their diamond hosts, a distinction is made “eclogitic” and “websteritic” inclusions, character-
between (i) protogenetic, (ii) syngenetic and (iii) ized by garnet containing variable amounts of a

11
T. STACHEL – DIAMOND

Eclogitic 
33%
Wehrlitic
Harzburgitic 1%
Peridotitic 86% Lherzolitic
65% 13%
Websteritic
2%

FIG. 1-10. Relative proportions of source parageneses for inclusion-bearing diamond derived from the subcratonic
lithospheric mantle. Diamond from the sublithospheric mantle is excluded and likely accounts for less than 1% (by mass)
of the world-wide diamond production. The pie chart on the left is based on 2844 inclusion-bearing diamonds (data-base
of Stachel & Harris 2008). The pie chart on the right splits the peridotitic suite into harzburgitic (-dunitic), lherzolitic and
wehrlitic parageneses. Percentages are derived from compositional classification of 685 peridotitic garnet inclusions (data-
base of Stachel & Harris 2008, corrected for an over-representation of inclusion-bearing diamonds from DeBeers Pool
mines, South Africa). In combination, the two graphs imply that about 56% of inclusion-bearing diamonds derive from
harzburgitic sources.
majorite component indicative of an ultra-high diamond inclusion garnet with Cr2O3 >4 wt.% from
pressure origin and (ii) diamonds from the lower- the lherzolitic field and, with reference to the
most transition zone and the lower mantle typically classification scheme of Dawson & Stephens
recognized by the presence of ferropericlase as part (1975), designated the garnet on the Ca-poor side of
of the inclusion assemblage. the line “G10” and on the Ca-rich side “G9”. The
The classical review on lithospheric inclusions two concepts of a mineralogically defined lherzolitic
in diamond was given by Meyer (1987), a com- field and a statistically based G10/G9 division were
prehensive recent update may be found in Stachel & recently merged by Grütter et al. (2004) who
Harris (2008). Inclusions in sublithospheric defined the exact boundaries of the garnet
diamond were reviewed by Stachel (2001), Stachel classification diagram shown in Figure 1-11. Based
et al. (2005), Harte (2010) and Kaminsky (2012). on garnet compositions, it is not possible to detect
the absence or presence of orthopyroxene in
Peridotitic inclusion suite clinopyroxene-free sources and hence to distinguish
Peridotitic diamond sources may be split into between harzburgitic (orthopyroxene-bearing) and
three parageneses: harzburgitic–dunitic (clino- dunitic (pyroxene-free) parageneses. In Figure 1-11
pyroxene absent), lherzolitic (orthopyroxene and dunitic garnet will plot together with highly
clinopyroxene present) and wehrlitic (orthopyroxene depleted harzburgitic garnet at CaO contents below
absent, clinopyroxene present). Classification of about 1.8 wt.% (cf. Grütter et al. 1999).
inclusions into one of the three parageneses is Mineral inclusion phases observed in peridotitic
commonly established using garnet chemistry. diamonds are listed in Table 1-2 and the relative
Based on xenolith studies, Sobolev et al. (1973) abundance of the various parageneses can be taken
were the first to define a lherzolitic field in Cr–Ca from Figure 1-10. Common inclusions phases are
space delineating garnet in equilibrium with both Cr-bearing (mainly as a knorringite component)
clinopyroxene and orthopyroxene. Garnet plotting pyrope garnet (29% of peridotitic inclusions),
on the Ca-poor side (“sub-calcic garnet”) of the olivine (28%), Mg-chromite (27%), orthopyroxene
lherzolitic field (Fig. 1-11) is derived from (12%) and clinopyroxene (4%). Very common Fe-
clinopyroxene-free harzburgitic–dunitic sources, Ni sulfide minerals are not included in calculated
whereas garnet falling on the Ca-rich side of the abundances because they are not systematically
lherzolite field cannot have been in equilibrium with recovered during inclusion studies (because of small
orthopyroxene and thus reflects wehrlitic mineral size and a common association with extensive
assemblages (Sobolev et al. 1969, 1973). Gurney fracture systems). The extreme overabundance of
(1984) used an “arbitrary line” to separate 85% of sulfide minerals, garnet and Mg-chromite relative to

12
T. STACHEL – DIAMOND

G10 G9 G12
20 FIG. 1-11. Cr–Ca plot conventionally
used for classification of garnet in
inclusion studies and for indicator
mineral assessment. Divisions
(dashed lines) are from Grütter et

c
15 al. (2004). Compositional fields for

zoliti
Harzburgitic garnet inclusions in diamond are
shown as shaded areas and are

Lher
based on 1830 individual
Cr2O3 inclusions. Note that the com-
10 positional field for wehrlitic garnet
is derived from seven inclusions
only.

5
Wehrlitic

Eclogitic
0
0 5 10 15 20
Websteritic CaO
normal peridotitic mineral modes indicates in source composition, eclogitic inclusions can
preferential “sampling” of these minerals by generally be readily distinguished from their
diamond, possibly related to surface energy peridotitic counterparts, e.g., on the basis of lower
minimization (Meyer & Boyd 1972) or to frequent Mg-numbers and Cr concentrations, and elevated
co-precipitation with diamond during melt/fluid Na, Ca and Al contents. Besides Fe sulfide minerals,
infiltration events. The sources of peridotitic common mineral inclusions in eclogitic diamond are
diamond are chemically highly depleted as a garnet (56% of inclusions, excluding sulfide
consequence of severe melt extraction during the minerals) and clinopyroxene (39%), and less
Archean (e.g., Boyd et al. 1993). This depleted abundant rutile (3%) and coesite (2%). Eclogitic
character is reflected by inclusions in diamond garnet inclusions are variable mixtures of the three
through overall high Mg-numbers (commonly >91 end-members grossular, almandine and pyrope.
for olivine and orthopyroxene, and >90 for Their most characteristic feature is the virtual
clinopyroxene, with median values of 92.9, 94.5 and absence of Cr: a cut-off at 1 wt.% Cr2O3 is used to
92.9, respectively), high Cr in garnet (typically >4 separate eclogitic from peridotitic garnet in Figure
wt.% Cr2O3, median value is 7.9 wt.%), and 1-11, but in reality 79% of eclogitic garnet contains
presence of very Cr-rich (molar 100Cr/[Cr+Al] >80) <0.1 wt.% Cr2O3. Eclogitic clinopyroxene typically
and Ti-poor (<1 wt.% TiO2) spinel. Lherzolitic is omphacitic (i.e., contains several wt.% Na2O and
inclusions are generally less magnesian than their Al2O3) and has Mg-numbers <80 (median value is
harzburgitic counterparts (Fig. 1-12); however, 76.8 wt.%). In rare cases where eclogitic
sharp divisions cannot be established. clinopyroxene reaches or exceeds Mg-numbers of
90 (typical for the peridotitic inclusion suite), a
Eclogitic inclusions suite distinction can still be made based on its low Cr–Al
The source of eclogitic diamonds, based on ratio (molar 100Cr/[Cr+Al] <10).
studies of inclusions in diamond and eclogite
xenoliths, has been shown to have a broadly basaltic Websteritic suite
bulk composition, chemically depleted relative to Diamond from websteritic sources is relatively
present day MORB tholeiite and more akin to uncommon and accounts for only about 2% of
picritic or komatiitic basalt (e.g., Rudnick 1995, inclusion-bearing diamonds (Fig. 1-10). In the
Jacob 2004). Based on this fundamental difference petrographic sense the terms websterite and olivine

13
T. STACHEL – DIAMOND

TABLE 1-2: MINERAL INCLUSIONS IN DIAMONDS.


Peridotitic Eclogitic Websteritic Sublithospheric Uncertain
Common Common Common Common Common
Cr-pyrope Grossular-almandine- Almandine-pyrope Majoritic garnet Graphite
Olivine pyrope Diopside-augite Omphacitic cpx
Enstatite Omphacitic cpx Enstatite Mg-Si-perovskite* Rare
Cr-diopside Fe sulfides Ca-Si-perovskite* Diamond
Mg-chromite Occasional Ferropericlase Calcite
Fe-Ni sulfides Occasional Coesite* Dolomite
Rutile Olivine Rare Perovskite
Rare Coesite* Stishovite* Amphibole
Coesite* Rare Ilmenite Moissanite
Mg-ilmenite Rare Phlogopite Corundum Apatite
Magnesite Kyanite Chromite Eskolaite
Calcite Corundum Spinel Sr-titanate
Native Fe Ilmenite Native Fe
Zircon Magnetite Native Ni
Phlogopite Fe-Mg-chromite Calcite
Yimengite Phlogopite Siderite
K-feldspar “Ca-feldspar”
Sphene “K-feldspar” (former
Staurolite hollandite?)
Zircon NAL (new aluminum
Moissanite silicate phase)*
Calcite Calcium ferrite*
Dolomite Ringwoodite*
Sphene
Sulfides

Retrograde(?)
Larnite (Ca2SiO4)
CaSi2O5
(titanite structure)
CaSiO3
(walstromite structure)
Type-III cpx (originally
garnet structured)
TAPP

Epigenetic phases are not listed. “Common”, “occasional” and “rare” refers to the relative abundance of a mineral within a
particular inclusion suite, with sublithospheric and websteritic inclusions being rare themselves. * refers to mineral inclusions
where an original structure is inferred based on pressure–temperature constraints. In the case of NAL and calcium ferrite, both
composition and structure are inferred for inclusions that underwent unmixing during exhumation from the lower mantle
(Walter et al. 2011). Inclusions in quotation marks have the composition of the mineral listed, but likely formed in a different
structure that cannot be inferred unambiguously. Compiled from a data-base of inclusions in diamonds from world-wide
sources (see Stachel & Harris 2008). Phases only occurring as submicroscopic (nanometre-sized) inclusions in sublithospheric
diamond (see Kaminsky 2012 for a review) are not included.

websterite refer to pyroxenite containing both Compositional criteria for assigning inclusions to
clinopyroxene and orthopyroxene. For inclusions in the websteritic rather than the eclogitic suite include
diamond the term websteritic suite is not well elevated Mg- and Cr-numbers for garnet (Fig. 1-11)
defined and has traditionally been applied to garnet and pyroxene inclusions and, for garnet, low Ca
and pyroxene inclusions with compositions contents (<6 wt.%, Grütter et al. 2004). Common
transitional between the peridotitic and eclogitic inclusions in websteritic diamond are almandine–
suites. Hence the websteritic suite encompasses both pyrope garnet (38% of inclusions), clinopyroxene
“peridotitic” inclusions with unusually low Mg- (with a limited omphacite component compared to
number (Gurney et al. 1984) and “eclogitic” garnet eclogitic inclusions) (37%), orthopyroxene (21%),
and clinopyroxene inclusions which coexist – or are with additional occurrence of coesite (3%) and
inferred to coexist – with orthopyroxene. olivine (1%).

14
T. STACHEL – DIAMOND

86 88 90 92 94 96 98 moderate majorite component likely derived from


14
extremely deep (>250 km) lithospheric roots
Lherzolitic (Pokhilenko et al. 2004, Stachel et al. 2006).
12
“Normal” sublithospheric inclusions of majoritic
N=49 garnet are generally eclogitic to websteritic in
10 Median=92.1
composition and, accordingly, sometimes accom-
Mean=92.1 ±1.0
Frequency

panied by omphacitic clinopyroxene inclusions.


8
Diamond from the lowermost transition zone
and the lower mantle is primarily described from
6 three locations: the Juina area in Brazil (including
Rio Sao Luiz; Harte et al. 1999b), the DO-27 pipe at
4 Lac de Gras, Canada (Davies et al. 1999) and the
Kankan area in Guinea (Stachel et al. 2000). The
2 crystal structure of true ultra-high pressure phases
(e.g., Ca–Si- and Mg–Si-perovskite or stishovite) is
0 not preserved in such inclusions. Evidence for a
lower mantle origin hence is based on coexistence
Harzburgitic
of the oxide ferropericlase ([Mg,Fe]O) with silica-
N=173 rich phases ([Mg,Fe]SiO3, CaSiO3 and SiO2
30
Median=93.2 compositions); such parageneses cannot have
Mean=93.2 ±1.0 formed outside the lower mantle as at lower
Frequency

pressures ferropericlase would react with these


20 phases to form olivine or pyroxene compositions.
Because of a very strong preference of Ni for
ferropericlase, original formation of inclusions with
orthopyroxene compositions as Mg–Si-perovskite in
10 the lower mantle may be inferred from low Ni
contents (NiO <300 ppm), with typical ortho-
pyroxene from the upper mantle containing >1000
0 ppm NiO. Additionally, a number of “exotic” ultra-
86 88 90 92 94 96 98 high pressure phases from the deep transition zone
Mg# and the lower mantle have been described as
inclusions in diamond, mainly from the Juina area
FIG. 1-12. Mg-number of olivine inclusions in diamond. (e.g., Harris et al. 1997, Hutchison et al. 2001,
Olivine of lherzolitic and harzburgitic paragenesis has
Kaminsky et al. 2001, Walter et al. 2011, Kaminsky
different modes, medians and means but overlaps
within one standard deviation and hence cannot be 2012; see Table 1-2).
separated based on Mg-number.
Mineral inclusion-based geothermobarometry
Sublithospheric inclusions The presence of multiple inclusion phases in
Garnet inclusions containing a majorite the same diamond allows determination of the
component – with an end-member composition pressure–temperature conditions of last equilibration
M3(AlM0.5Si0.5)[VI]Si3[IV]O12 – were first discovered via mineral exchange geothermobarometry. The
by Moore & Gurney (1985) in diamond from the most commonly applied thermometers are based on
Monastery mine (South Africa) and have since been the coexistence of garnet with olivine (TO’Neill 79,
recognized in diamond from several other deposits O’Neill & Wood 1979, O’Neill 1980) and
(see Stachel 2001 for a review). In analogy to garnet clinopyroxene (TKrogh 88, Krogh 1988), whereas pairs
compositions observed in high pressure experi- of garnet and orthopyroxene may be employed to
ments, inclusions of majoritic garnet are inferred to derive both pressure (PBKN, Brey & Köhler 1990)
indicate diamond formation beneath the subcratonic and temperature (THarley, Harley 1984). In cases
lithosphere (i.e., deeper than ca. 200 km) and where the inclusion pairs are not in contact, the
extending into the transition zone down to at least resulting P–T estimates represent the ambient
500 km depth (e.g., Tappert et al. 2005b). Rare conditions at the time of diamond formation. At
exceptions are Cr-rich garnet inclusions with a high temperatures in the Earth’s mantle, touching or

15
T. STACHEL – DIAMOND

intergrown mineral pairs will over time re- PNT 00 [kbar]


equilibrate in response to changes in ambient 30 40 50 60 70 80
pressure and/or temperature and hence reflect the 30
conditions of mantle storage prior to kimberlite N = 79
magmatism. Single phase geothermometers (Ni-in- Median=51
garnet: TNi, Griffin et al. 1989, Canil 1999; Zn-in- [kbar]
spinel: TRyan, Ryan et al. 1996; Mn-in-garnet: TMn, Mean = 50
20 ±8 [kbar] Lherzolitic

Frequency
Grütter et al. 1999, Creighton 2009) and
thermobarometers (clinopyroxene: PNT 00 and TNT 00,
Nimis & Taylor 2000) allow determination of the
conditions of formation for diamond containing only
one inclusion species. A comprehensive review of 10
the P–T conditions of diamond formation and
mantle storage may be found in Stachel & Harris
(2008) and only a short summary is given here.
PBKN, calculated iteratively in combination with
0
THarley, and PNT 00 in combination with TNT 00, are the
two thermobarometers that can be applied to the N = 88
largest number of inclusions in diamond (Fig. 1-13). 20 Median=55
Although only a limited number of lherzolitic [kbar] Harzb.
Mean = 55 Lherz.
garnet–orthopyroxene pairs is available, the overall Frequency ±8 [kbar]
agreement in PBKN with inclusion pairs of
harzburgitic paragenesis (average lherzolitic: 51
±10 kbar; average harzburgitic: 55 ±8 kbar)
indicates formation at similar conditions. Therefore, 10
despite being based on lherzolitic inclusions only,
the more accurate (Nimis & Grütter 2010) pressure
values derived via the PNT 00–TNT 00 combination
may be used to derive 50 ±8 kbar (equivalent to 160
±25km) as the best estimate for average conditions 0
of diamond formation. Assuming equilibration at 30 40 50 60 70 80
this average pressure (50 kbar), the olivine–garnet PBKN [kbar]
(TO’Neill 79) and the clinopyroxene–garnet (TKrogh 88)
thermometers can be applied to a large number of FIG. 1-13. Pressure of origin for peridotitic (harzburgitic
and lherzolitic parageneses) inclusions in diamond
peridotitic and eclogitic suite inclusion pairs. The
based on simultaneous estimates (iterative calculations)
two thermometers yield matching thermal of pressure and temperature. Garnet–orthopyroxene
conditions for peridotitic and eclogitic inclusion pairs (bottom histogram) are present in diamond of
(average of 1160–1170 ±110°C, modes in class both harzburgitic and lherzolitic parageneses and allow
1150–1200°C, see Fig. 1-14). Temperature for simultaneous derivation of PBKN (Brey & Köhler
estimates for harzburgitic and lherzolitic garnet– 1990) and THarley (Harley 1984). Peridotitic clinopyrox-
olivine inclusion pairs are also indistinct (Fig. 1-14, ene inclusions (top) are, by definition, absent in
averages agree within 10°C). diamond of harzburgitic paragenesis and, therefore,
Figure 1-15 shows pressure–temperature originate from lherzolitic sources only (ignoring the
estimates (PBKN–THarley and PNT 00–TNT 00) for possible presence of rare wehrlitic clinopyroxene
peridotitic inclusions relative to the continental inclusions). Before calculation of PNT 00 and TNT 00, the
filters of Grütter (2009) were applied to exclude
geotherms of Hasterok & Chapman (2011). The data
clinopyroxene compositions outside the calibration
strongly scatter, with a number of points falling range of Nimis & Taylor (2000). The two
outside the diamond stability field, and cover a geothermobarometer combinations indicate similar
range of model geotherms between 35 and 45 average conditions of diamond formation, at pressures
mW/m2 surface heat flow. Linear regression of all of 50–55 (±8) kbar. However, because of the lower
data points (PBKN–THarley and PNT 00–TNT 00, Fig. accuracy of THarley compared to TNT 00, the pressure
1-15) yields a trend approximately equivalent to a estimates derived through the combination of PNT 00
38 mW/m2 model geotherm. A large comparative and T NT 00 are considered more reliable.

16
T. STACHEL – DIAMOND

TKrogh 88 [°C] study of non-touching and touching garnet–


800 900 1000 1100 1200 1300 1400 orthopyroxene pairs in diamond from the De Beers
40 Pool kimberlite pipes (South Africa) produced
strong evidence that diamond formation (average
Eclogitic: N=144
non-touching inclusions: 1200°C) was followed by
Median=1160 °C
cooling (average touching: 1080°C) by about 120°C
30 Mean=1170 (Phillips & Harris 1995, Phillips et al. 2004; for this
±110 °C
relative comparison the 2 sigma uncertainty is
Frequency

estimated to be ±25°C). Subsequently, indications


20
for a similar decrease in ambient temperature soon
after diamond formation were also found elsewhere
(Stachel et al. 2003, 2004), suggesting that diamond
formation may relate to short-lived heating events.
10 Garnet–orthopyroxene pairs falling below a 37
mW/m2 Hasterok & Chapman (2011) model
geotherm thus likely represent conditions of mantle
storage rather than diamond formation.
0
Peridotitic: N=164 Harzb. Fluid and melt inclusions in diamond
Median=1140 °C Fibrous and clouded diamond usually contains
Lherz.
innumerable micrometre to nanometre-sized melt
30 Mean=1160
±110 °C (or “high density fluid”) inclusions. Electron
microprobe-based studies of the bulk composition of
Frequency

such inclusions, pioneered by Navon and co-


20 workers (e.g., Navon et al. 1988, Schrauder &
Navon 1994, Izraeli et al. 2001, Klein-BenDavid et
al. 2009), revealed the presence of four
compositional end-members: (i) hydrous chloride
10 melt, (ii) high-Mg carbonatitic melt and an array
between (iii) hydrous-silicic and (iv) low-Mg
carbonatitic melts. Studies of the daughter minerals
in the inclusions (e.g., Guthrie et al. 1991, Klein-
0 BenDavid et al. 2006, Weiss et al. 2010, Kopylova
800 900 1000 1100 1200 1300 1400
et al. 2010) revealed multi-phase assemblages of
TO’Neill 79 [°C] carbonate phases, mica, apatite and quartz, with
FIG. 1-14. Temperature estimates for peridotitic possible occurrence of additional phases. The trace
(harzburgitic and lherzolitic parageneses; based on element abundance patterns of the inclusions (e.g.,
Mg–Fe exchange between garnet–olivine; TO’Neill 79) Schrauder et al. 1996, Tomlinson et al. 2005,
and eclogitic (Mg–Fe exchange between garnet– Zedgenizov et al. 2007, Rege et al. 2010, Weiss et
clinopyroxene; TKrogh 88) inclusions in diamond. For the al. 2011) are very high in incompatible elements
calculations a fixed pressure of 50 kbar (see Fig. 1-13) and show broad similarity to those of kimberlite
was assumed. Results exceeding the maximum (e.g., Tomlinson et al. 2005, Weiss et al. 2011).
temperature of lithospheric mantle (1400°C, see Fig. The genetic and temporal relationship between
1-2) likely represent disequilibrium between inclusion fibrous/clouded and clear monocrystalline (gem-
phases and were disregarded. For eclogitic diamonds
type) diamond is not fully understood. Detectable
the inclusion suite from Argyle (average temperature
of 1240°C) was considered to represent unusually hot
concentrations (at the ppm to ppt level) of lithophile
conditions and hence excluded. elements in clear monocrystalline diamond have
been related to submicroscopic fluid and melt
inclusions (Fesq et al. 1973, McNeill et al. 2009,
Rege et al. 2010, Melton et al. 2012), but may
equally be explained through nanometre-sized
mineral inclusions (Melton et al. 2012), implying
that the origin of gem diamonds is more complex.

17
T. STACHEL – DIAMOND

TP=1300°C
Adiabat
40
FIG. 1-15. Simultaneous estimates of pressure and
temperature of origin using the PBKN–THarley
(garnet–orthopyroxene) and PNT 00–TNT 00
4 (clinopyroxene) thermobarometer combinations.
Gra 5
50 p The uncertainty of individual data points is equal
Dia hite
P [kbar]

mo to or better than 5 kbar and 100°C. Mantle


nd
adiabat (for a mantle potential temperature of
1300˚C and a gradient of 0.3˚C/km) and model
geotherms equivalent to 35–45 mW/m2 surface
60
PNT 00-PNT 00 40 heat flow are plotted after Hasterok & Chapman
(2011), the graphite–diamond transition is from
Lherzolitic Day (2012). The average (± one standard
PBKN-THarley deviation) and a linear regression (r2=0.68) for all
70 Lherzolitic data points are indicated as a black cross and a
Harzburgit. solid grey line, respectively.
35

900 1000 1100 1200 1300 1400


T [°C]

TIMING OF DIAMOND FORMATION Ma) possibly related to kimberlite magmatism


Nitrogen aggregation and the age of diamonds (Boyd et al. 1992, Gurney et al. 2010).
The general antiquity of diamonds relative to
their kimberlite hosts can be established based on Radiometric dating of inclusions
the aggregation state and concentration of N An indirect way of obtaining diamond ages is
impurities in the diamond lattice (Fig. 1-16). Taking radiometric dating of silicate and sulfide inclusions.
the combined average of the geothermometry data Despite challenges to the applicability of inclusion-
for peridotitic and eclogitic inclusions (1170 derived ages to the timing of diamond formation
±110°C, see above) and assuming that diamond (e.g., Navon 1999, Spetsius et al. 2002), the
storage temperatures did not dramatically fluctuate approach is generally considered valid (e.g., Pearson
over time, it is possible to derive crude relative & Shirey 1999, Richardson et al. 2004). Some of
diamond ages by applying N thermochronometry the arguments brought forward as to why radiogenic
(see section “Optical Centers”, p. 6) to match these dating of inclusions indeed yields diamond
temperatures. Assuming mantle residence times of formation ages include (i) closure temperatures for
3, 1 and 0.1 Ga (resulting in average TNitrogen of Re and Os in sulfide minerals well below ambient
1130, 1150 and 1210°C for diamonds with temperatures in Earth’s mantle at depths beneath the
peridotitic and eclogitic inclusions) provides a graphite–diamond transition and (ii) an agreement
reasonable bracket around the inclusion-based of sulfide-based Re–Os and silicate-based Sm–Nd
temperatures. Diamond formation only 10 Ma ages – as shown for eclogitic inclusions from
before kimberlite eruption (average TNitrogen of Jwaneng (Richardson et al. 1999, 2004) and Orapa
1280°C, Fig. 1-16), however, would require mantle (Richardson 1986, Shirey et al. 2001) – that is too
storage at temperatures well above (by about 110°C) close to be fortuitous (see Fig. 1-17).
those indicated by inclusion thermometry. Although A comprehensive review on inclusion-based
only crude age constraints can be established, N dating of diamonds was presented by Pearson &
characteristics nevertheless provide clear evidence Shirey (1999) with recent updates in Stachel &
that diamond generally formed at least hundreds of Harris (2008) and Gurney et al. (2010). Sm–Nd
millions of years prior to transport to Earth’s ages, generally based on garnet (± clinopyroxene)
surface. composites, and Re–Os ages, determined through
However, exceptions to this generalization are analysis of individual sulfide inclusions, are
established e.g., through fibrous cubes and fibrous summarized in Figure 1-17. With the exception of a
coats on gem diamonds, with poorly aggregated N Mesozoic isochron age obtained from two sulfide
in the overgrowths pointing to a late stage of inclusions in a single diamond from Koffiefontein
diamond formation (mantle residence of less than 5 (Pearson et al. 1998) and an inferred similar age for

18
T. STACHEL – DIAMOND

1000 1200 1400 1600 a peridotitic sulfide inclusion from Jagersfontein


250
(Aulbach et al. 2009), all dated peridotitic diamonds
1 stdev
0.01 Ga formed between 3.6 and 1.4 Ga. Formation of
mineral
200 Mean=1280
thermom.
±80 °C
diamonds containing harzburgitic (3.6–3.2 Ga) and
lherzolitic inclusions (about 2 Ga for Premier,
Venetia and Udachnaya, about 1.4 Ga for Ellendale)
Frequency

150 Eclogitic
occurred during two discrete periods in the
Paleoarchean and in the Paleoproterozoic to
100 Peridotitic Mesoproterozoic, respectively. This has been related
to early formation of harzburgitic diamonds during
50
initial craton stabilization and root formation, and
later addition of lherzolitic diamonds during
Proterozoic craton modification (accretion along
0 craton rims and magmatic mantle re-fertilization;
0.1 Ga
Shirey et al. 2004, Gurney et al. 2010). Formation
Mean=1210 of eclogitic diamonds is documented between 2.9
200
±80 °C and 0.99 Ga (Fig. 1-17). Mesoarchean ages for
eclogitic sulfide inclusions are interpreted to mark
Frequency

150 the onset of modern style plate tectonics at around 3


Ga (Helmstaedt et al. 2010, Shirey & Richardson
2011). With the exception of Koffiefontein
100
(peridotitic) and Premier (eclogitic), inclusion-based
ages of diamond formation predate the eruption of
50 their kimberlite and lamproite hosts by hundreds of
millions to billions of years (Fig. 1-17), consistent
with the inferences made above based on N
0
aggregation states.
1 Ga
200 Mean=1150
±70 °C
Frequency

150

100

50

FIG. 1-16. Time-averaged mantle residence temperatures


0 for 1285 diamonds with peridotitic and eclogitic
inclusions (data-base of Stachel et al. 2009), based on
3 Ga the N thermometer of Taylor et al. (1990) and Leahy
200 Mean=1130 and Taylor (1997). TNitogen is shown for mantle
±70 °C
residence times of 3, 1, 0.1 and 0.01 Ga. TNitogen cannot
Frequency

be calculated for diamonds with 0 (N=172) or 100%


150 (N=17) N in B centers; therefore, aggregation states of
0.1 and 99.9 %B were instead assumed for such
100
samples, making the calculated residence temperatures
maximum and minimum estimates, respectively. For a
given residence time, average residence temperatures
50 for diamonds with peridotitic and eclogitic inclusions
differ by less than 20 °C. The temperature range (1
standard deviation about the mean) for conventional
0 mineral exchange thermometry (see Fig. 1-14) is
1000 1200 1400 1600 shown as a shaded area.
TNitrogen [°C]

19
T. STACHEL – DIAMOND

4.0
Peridotitic
Kaapvaal Slave
SE
3.0 Daldyn Australia
Age [Ga]

Limpopo (alluv.)

2.0
Kimberley

1.0

A154-S
Koffiefont.

Wellington
Finsch

Ellendale
Venetia
Kimberley

Panda
Premier

Udach-
naya

Model ages
Eclogitic Sm-Nd
4.0 Re-Os
Kaapvaal
Isochrons
Zimbabwe Sm-Nd
3.0
Age [Ga]

Re-Os
Limpopo Slave Kimberlite
2.0

1.0
Kimberley

0
Koffiefont.
Jagersfont.

Argyle
Finsch

Orapa
Venetia

A154-S
Jwaneng
Kimberley

Premier
Klipspringer

FIG. 1-17. Compilation of diamond ages based on the Sm–Nd (garnet ± clinopyroxene) and Re–Os (sulfide minerals) isotopic
systems. Results for peridotitic and eclogitic diamonds are shown separately. Eruption ages of the respective kimberlite or
lamproite hosts are given as well (triangles). The host kimberlite bodies are listed along the X-axis, the respective host
cratons/areas are indicated inside the graphs. For a complete list of references see Stachel & Harris (2008, their Table 9)
and Gurney et al. (2010, their Tables 1–4).

20
T. STACHEL – DIAMOND

CONCLUSIONS BOYD, F.R. & GURNEY, J.J. (1986): Diamonds and


Based on their mineral inclusion content gem the African lithosphere. Science 232, 472–477.
diamonds form in peridotitic, eclogitic and BOYD, S.R., PILLINGER, C.T., MILLEDGE, H.J.,
websteritic sources located in the subcratonic MENDELSSOHN, M.J. & SEAL, M. (1992): C-
lithospheric mantle. Rare exceptions are isotopic and N-isotopic composition and the
sublithospheric diamonds derived from the infrared-absorption spectra of coated diamonds –
asthenosphere, the transition zone and the lower Evidence for the regional uniformity of CO2–H2O
mantle. Common peridotitic and eclogitic diamonds rich fluids in lithospheric Mantle. Earth. Planet.
generally are ancient (ca. 3.6–1.0 Ga) and reside in Sci. Lett. 109, 633–644.
relatively cool (bracketed by 35 and 45 mW/m2
BOYD, F.R., PEARSON, D.G., NIXON, P.H. &
surface heat flow model geotherms of Hasterok &
MERTZMAN, S.A. (1993): Low-calcium garnet
Chapman 2011) lithospheric sections ranging in
harzburgites from southern Africa – Their
depth between about 105 and 200 km. From there,
relations to craton structure and diamond
diamonds are transported to Earth’s surface by
crystallization. Contrib. Mineral. Petrol. 113,
rapidly ascending kimberlite or lamproite magmas.
352–366.
Resorption by the transporting magma may convert
primary octahedral or cuboid diamonds into BOYD, S.R., PINEAU, F. & JAVOY, M. (1994):
secondary dodecahedral shapes, resulting in loss of Modelling the growth of natural diamonds. Chem.
diamond mass. Color centers in diamond originate Geol. 116, 29–42.
from molecular impurities giving rise to yellow BREY, G.P. & KÖHLER, T. (1990):
(nitrogen C and N2 centers) and blue (boron) Geothermobarometry in four-phase lherzolites II.
colouration or are due to plastic deformation (brown New thermobarometers, and practical assessment
and pink). A type classification of diamonds is of existing thermobarometers. J. Petrol. 31,
based on N contents and aggregation states. 1353–1378.
Dissolution, etching, plastic deformation, BREY, G.P., BULATOV, V., GIRNIS, A., HARRIS, J.W.
mechanical wear and irradiation produce a range of & STACHEL, T. (2004): Ferropericlase – a lower
surface features that may be used to further mantle phase in the upper mantle. Lithos 77, 655–
characterize diamonds from individual deposits. 663.
ACKNOWLEDGEMENTS BULANOVA, G.P. (1995): The formation of
Jeff W. Harris and K. S. (Fanus) Viljoen are diamond. J. Geochem. Explor. 53, 1–23.
thanked for their valuable input into the diamond CANIL, D. (1999): The Ni-in-garnet geothermo-
morphology section. Heidi Höfer provided advice meter: calibration at natural abundances. Contrib.
on crystal growth, Oded Navon on fluid inclusions Mineral. Petrol. 136, 240–246.
in diamond. Anetta Banas and Steven Creighton CAPDEVILA, R., ARNDT, N., LETENDRE, J. &
helped during data collection and writing. George SAUVAGE, J.F. (1999): Diamonds in volcaniclastic
Harlow and Tom McCandless are thanked for komatiite from French Guiana. Nature 399, 456–
formal reviews of the manuscript. Funding by 458.
NSERC and the CRC program is gratefully
CARLISLE, D.B. & BRAMAN, D.R. (1991) Nano-
acknowledged.
meter-size diamonds in the Cretaceous/Tertiary
boundary clay of Alberta. Nature 352, 708–709.
REFERENCES
CARTIGNY, P. (2005): Stable isotopes and the origin
AULBACH, S., SHIREY, S.B., STACHEL, T.,
of diamond. Elements 1, 79–84.
CREIGHTON, S., MUEHLENBACHS, K. & HARRIS,
J.W. (2009): Diamond formation episodes at the CARTIGNY, P., HARRIS, J.W. & JAVOY, M. (1998):
southern margin of the Kaapvaal Craton: Re–Os Eclogitic diamond formation at Jwaneng: No
systematics of sulfide inclusions from the room for a recycled component. Science 280,
Jagersfontein Mine. Contrib. Mineral. Petrol. 1421–1424.
157, 525-540. CARTIGNY, P., HARRIS, J.W. & JAVOY, M. (2001):
BALLHAUS, C. (1992): Redox states of lithospheric Diamond genesis, mantle fractionations and
and asthenospheric upper mantle. Contrib. mantle nitrogen content: a study of δ13C–N
Mineral. Petrol. 114, 331–348. concentrations in diamonds. Earth. Planet. Sci.
Lett. 185, 85–98.

21
T. STACHEL – DIAMOND

CREIGHTON, S. (2009): A semi-empirical EVANS, T. & HARRIS, J.W. (1989): Nitrogen


manganese-in-garnet single crystal thermometer. aggregation, inclusion equilibration temperatures
Lithos 112 (S1), 177–182. and the age of diamonds. In Kimberlites and
CREIGHTON, S., STACHEL, T., MATVEEV, S., HÖFER, related rocks. (J. Ross et al., eds), Blackwell,
H., MCCAMMON, C. & LUTH, R.W. (2009): Carlton, 1001–1006.
Oxidation of the Kaapvaal lithospheric mantle EVANS, T., KIFLAWI, I., LUYTEN, W.,
driven by metasomatism. Contrib. Mineral. VANTENDELOO, G. & WOODS, G.S. (1995):
Petrol. 157, 491–504. Conversion of platelets into dislocation loops and
CREIGHTON, S., STACHEL, T., EICHENBERG, D. & voidite formation in Type Iab diamonds. Proc. R.
LUTH, R.W. (2010): Oxidation state of the Soc. London Ser. A – Math. Phys. Sci. 449, 295–
lithospheric mantle beneath Diavik diamond 313.
mine, central Slave craton, NWT, Canada. FEDORTCHOUK, Y. & CANIL, D. (2009): Diamond
Contrib. Mineral. Petrol. 159, 645–657. oxidation at atmospheric pressure: development
DAVIES, R.M., GRIFFIN, W.L., PEARSON, N.J., of surface features and the effect of oxygen
ANDREW, A.S., DOYLE, B.J. & O'REILLY, S.Y. fugacity. Eur. J. Mineral. 21, 623–635.
(1999): Diamonds from the deep: pipe DO–27, FEDORTCHOUK, Y. & ZHANG, Z. (2011): Diamond
Slave Craton, Canada. In The J.B. Dawson Resorption: Link to metasomatic events in the
Volume, Proceedings of the VIIth International mantle or record of magmatic fluid in kimberlitic
Kimberlite Conference, (J.J. Gurney, J.L. Gurney, magma? Can. Mineral. 49, 707–719.
M.D. Pascoe & S.H. Richardson, eds.), Red Roof FESQ, H.W., BIBBY, D.M., ERASMUS, C.S., KABLE,
Design, Cape Town, 148–155. E.J.D. & SELLSCHOP, J.P.F. (1973): A compar-
DAWSON, J.B. & STEPHENS, W.E. (1975): Statistical ative trace element study of diamonds from
classification of garnets from kimberlite and Premier, Finsch and Jagersfontein mines, South
associated Xenoliths. J. Geol. 83, 589–607. Africa. Phys. Chem. Earth. 9, 111–114.
DAY, H.W. (2012): A revised diamond–graphite FIPKE, C.E., GURNEY, J.J. & MOORE, R.O. (1995):
transition curve. Am. Miner. 97, 52–62. Diamond exploration techniques emphasising
DE STEFANO, A., KOPYLOVA, M.G., CARTIGNY, P. & indicator mineral geochemistry and Canadian
AFANASIEV, V. (2009): Diamonds and eclogites examples. Geol. Surv. Can., Bull. 423, 24–36
of the Jericho kimberlite (Northern Canada). FISHER, D. (2009): Brown diamonds and high
Contrib. Mineral. Petrol. 158, 295–315. pressure high temperature treatment. Lithos 112
DEINES, P. (2002): The carbon isotope geochemistry (S2), 619–624.
of mantle xenoliths. Earth Science Reviews 58, FRITSCH, E. (1988): The nature of color in
247–278. diamonds. In The Nature of Diamonds, (G.E.
DONNELLY, C.L., STACHEL, T., CREIGHTON, S., Harlow, ed.), Cambridge University Press,
MUEHLENBACHS, K. & WHITEFORD, S. (2007): Cambridge, 23–47.
Diamonds and their mineral inclusions from the GAILLOU, E., POST, J.E., ROST, D. & BUTLER, J.E.
A154 South pipe, Diavik Diamond Mine, (2012): Boron in natural type IIb blue diamonds:
Northwest Territories, Canada. Lithos, (in press). Chemical and spectroscopic measurements. Am.
DYER, H.B., RAAL, F.A., DUPREEZ, L. & LOUBSER, Mineral. 97, 1–18.
J.H.N. (1965): Optical absorption features GRIFFIN W.L., RYAN C.G., COUSENS D.R. & SUTER
associated with paramagnetic nitrogen in G.F. (1989): Ni in chrome pyrope garnets: a new
diamond. Philosophical Magazine 11, 763–774. geothermometer. Contrib. Mineral. Petrol. 103,
EGGLER, D.H. (1989): Kimberlites: how do they 99–202.
form? In Kimberlites and related rocks. (J. Ross GRÜTTER, H.S. (2009): Pyroxene xenocryst
et al., eds), Blackwell, Carlton, 489–504. geotherms: Techniques and application. Lithos
EVANS, T. (1992): Aggregation of nitrogen in 112 (S2), 1167–1178.
diamond. In The Properties of Natural and GRÜTTER, H.S., APTER, D.B. & KONG, J. (1999):
Synthetic Diamond, (J.E. Field, ed.), Academic Crust-mantle coupling: evidence from mantle-
Press, London, 259–290. derived xenocrystic garnets. In The J.B. Dawson
Volume, Proceedings of the VIIth International

22
T. STACHEL – DIAMOND

Kimberlite Conference, (J.J. Gurney, J.L. Gurney, HARRIS, J.W., HUTCHISON, M.T., HURSTHOUSE, M.,
M.D. Pascoe & S.H. Richardson, eds.), Red Roof LIGHT, M. & HARTE, B. (1997): A new tetragonal
Design, Cape Town, 307–313. silicate mineral occurring as inclusions in lower-
GRÜTTER, H.S., GURNEY, J.J., MENZIES, A.H. & mantle diamonds. Nature 387, 486–488.
WINTER, F. (2004): An updated classification HARTE, B. (2010): Diamond formation in the deep
scheme for mantle–derived garnet, for use by mantle: the record of mineral inclusions and their
diamond explorers. Lithos 77, 841-857. distribution in relation to mantle dehydration
GURNEY, J.J. (1984): A correlation between garnets zones. Min. Mag. 74, 189–215.
and diamonds in kimberlites. In Publs. Geol. HARTE, B., FITZSIMONS, I.C.W., HARRIS, J.W. &
Dept. & Univ. Extension, Univ. West. Aust. 8, OTTER, M.L. (1999a): Carbon isotope ratios and
143–166. nitrogen abundances in relation to
GURNEY, J.J. (1989): Diamonds. In Kimberlites and cathodoluminescence characteristics for some
related rocks. (J. Ross et al., eds), Blackwell, diamonds from the Kaapvaal Province, S-Africa.
Carlton, 935–965. Mineral. Mag. 63, 829–856.
GURNEY, J.J. & SWITZER, G.S. (1973): The HARTE, B., HARRIS, J.W., HUTCHISON, M.T., WATT,
discovery of garnets closely related to diamonds G.R. & WILDING, M.C. (1999b): Lower mantle
in the Finsch pipe, South Africa. Contrib. mineral associations in diamonds from Sao Luiz,
Mineral. Petrol. 39, 103–116. Brazil. In Mantle Petrology: Field Observations
and High Pressure Experimentation: A tribute to
GURNEY, J.J., HARRIS, J.W. & RICKARD, R.S.
Francis R. (Joe) Boyd (Y. Fei, C.M. Bertka &
(1984): Silicate and oxide inclusions in diamonds
B.O. Mysen, eds.), The Geochemical Society,
from the Orapa Mine, Botswana. In Kimberlites
Houston, 125–153.
II: the mantle and crust-mantle relationships (J.
Kornprobst, ed.), Elsevier, Amsterdam, 3–9. HASTEROK, D. & CHAPMAN, D.S. (2011): Heat
production and geotherms for the continental
GURNEY, J.J., HELMSTAEDT, H.H., RICHARDSON,
lithosphere. Earth and Planetary Science Letters
S.H. & SHIREY, S.B. (2010): Diamonds through
307, 59–70.
Time. Econ. Geol. 105, 689–712.
HELMSTAEDT, H.H., GURNEY, J.J. & RICHARDSON,
GUTHRIE, G.D., VEBLEN, D.R., NAVON, O., &
S.H. (2010): Ages of cratonic diamond and
ROSSMAN, G.R. (1991): Submicrometer fluid
lithosphere evolution: constraints on Precambrian
inclusions in turbid-diamond coats. Earth Planet.
tectonics and diamond exploration. Can. Mineral.
Sci. Lett. 105, 1–12.
48, 1385–1408.
HAGGERTY, S.E. (1986): Diamond genesis in a
HOWELL, D., O’NEILL, C., GRANT, K., GRIFFIN, W.,
multiply-constrained model. Nature 320, 34–37.
O’REILLY, S., PEARSON, N., STERN, R. &
HAINSCHWANG, T., KATRUSHA, A. & VOLLSTAEDT, STACHEL, T. (2012): Platelet development in
H. (2005): HPHT treatment of different classes of cuboid diamonds: insights from micro-FTIR
type I brown diamonds. J. Gemmol. 29, 261–273. mapping. Contrib. Mineral. Petrol., 1–15.
HARLEY, S.L. (1984): An experimental study of the HUIZENGA, J.M., CROSSINGHAM, A. & VILJOEN, F.
partitioning of iron and magnesium between (2012): Diamond precipitation from ascending
garnet and orthopyroxene. Contrib. Mineral. reduced fluids in the Kaapvaal lithosphere:
Petrol. 86, 359–373. Thermodynamic constraints. Comptes Rendus
HARRIS, J.W. (1968a): The recognition of diamond Geoscience 344, 67–76.
inclusions. Part 1: Syngenetic mineral inclusions. HUTCHISON, M.T., CARTIGNY, P. & HARRIS, J.W.
Industrial Diamond Review 28, 402–410. (1999): Carbon and nitrogen compositions and
HARRIS, J.W. (1968b): The recognition of diamond physical characteristics of transition zone and
inclusions. Part 2: Epigenetic mineral inclusions. lower mantle diamonds from Sao Luiz, Brazil. In
Industrial Diamond Review 28, 558–561. The J.B. Dawson Volume, Proceedings of the
HARRIS, J.W. (1987): Recent physical, chemical, VIIth International Kimberlite Conference, (J.J.
and isotopic research of diamond. In Mantle Gurney, J.L. Gurney, M.D. Pascoe & S.H.
Xenoliths, (P.H. Nixon, ed.), John Wiley & Sons Richardson, eds.), Red Roof Design, Cape Town,
Ltd., Chichester, 477–500. 372–382.

23
T. STACHEL – DIAMOND

HUTCHISON, M.T., HURSTHOUSE, M.B. & LIGHT, KNOCHE, R., SWEENEY, R.J. & LUTH, R.W. (1999):
M.E. (2001): Mineral inclusions in diamonds: Carbonation and decarbonation of eclogites: the
associations and chemical distinctions around the role of garnet. Contrib. Mineral. Petrol. 135,
670-km discontinuity. Contrib. Mineral. Petrol. 332–339.
142, 119–126. KOPYLOVA, M.G., RICKARD, R.S.,
IZRAELI, E.S., HARRIS, J.W. & NAVON, O. (2001): KLEYENSTUEBER, A., TAYLOR, W.R., GURNEY,
Brine inclusions in diamonds: a new upper mantle J.J. & DANIELS, L.R.M. (1997): First occurrence
fluid. Earth Planet. Sci. Lett. 187, 323–332. of strontian K-Cr-loparite and Cr-chevkinite in
JACOB, D.E. (2004): Nature and origin of eclogite diamonds. Geol. Geofiz. 38, 382–397.
xenoliths from kimberlites. Lithos 77, 295–316. KOPYLOVA, M., NAVON, O., DUBROVINSKY, L. &
JANSE, A.J.A. (1994): Is Clifford's rule still valid? KHACHATRYAN, G. (2010): Carbonatitic
Affirmative examples from around the world. In mineralogy of natural diamond-forming fluids.
Diamonds: characterization, genesis and Earth. Planet. Sci. Lett. 291, 126–137.
exploration, (ed. H.O.A. Meyer & O.H. KROGH, E. (1988): The garnet-clinopyroxene iron-
Leonardos), CPRM Special Publication Jan/94, magnesium geothermometer – a reinterpretation
Brasilia, 215–235. of existing experimental data. Contrib. Mineral.
JONES, R. (2009): Dislocations, vacancies and the Petrol. 99, 44–48.
brown colour of CVD and natural diamond. LEAHY, K. & TAYLOR, W.R. (1997): The influence
Diam. Relat. Mat. 18, 820–826. of the Glennie domain deep structure on the
KAMINSKY, F. (2012): Mineralogy of the lower diamonds in Saskatchewan kimberlites. Geol.
mantle: A review of 'super-deep' mineral Geofiz. 38, 451–460.
inclusions in diamond. Earth Science Reviews LUTH, R.W. (1995): Experimental determination of
110, 127–147. the reaction dolomite plus 2 coesite equals
KAMINSKY, F.V., ZAKHARCHENKO, O.D., DAVIES, diopside plus 2 CO2 to 6 GPa. Contrib. Mineral.
R., GRIFFIN, W.L., KHACHATRYAN-BLINOVA, Petrol. 122, 152–158.
G.K. & SHIRYAEV, A.A. (2001): Superdeep MCCAMMON, C. & KOPYLOVA, M.G. (2004): A
diamonds from the Juina area, Mato Grosso State, redox profile of the Slave mantle and oxygen
Brazil. Contrib. Mineral. Petrol. 140, 734–753. fugacity control in the cratonic mantle. Contrib.
KENNEDY, C.S. & KENNEDY, G.C. (1976): The Mineral. Petrol. 148, 55–68.
equilibrium boundary between graphite and MCCANDLESS, T.E. & GURNEY, J.J. (1997):
diamond. J. Geophys. Res. 81, 2467–2470. Diamond eclogites: Comparison with
KIRKLEY, M.B., GURNEY, J.J., OTTER, M.L., HILL, carbonaceous chondrites, carbonaceous shales,
S.J. & DANIELS, L.R. (1991): The application of and microbial carbon-enriched MORB. Geol.
C isotope measurements to the identification of Geofiz. 38, 371–381.
the sources of C in diamonds: a review. Appl. MCKENNA, N., GURNEY, J.J., KLUMP, J. &
Geochem. 6, 477–494. DAVIDSON, J.M. (2004): Aspects of diamond
KLEIN-BENDAVID, O., WIRTH, R. & NAVON, O. mineralisation and distribution at the Helam
(2006): TEM imaging and analysis of Mine, South Africa. Lithos 77, 193–208.
microinclusions in diamonds: a close look at MCKENZIE, D. & O'NIONS, R.K. (1991): Partial melt
diamond growing fluids. Am. Mineral. 91, 353– distribution from inversion of rare-earth-element
365. concentrations. J. Petrol. 32, 1021–1091.
KLEIN-BENDAVID, O. LOGVINOVA, A.M., MCNEILL, J., PEARSON, D.G., KLEIN-BENDAVID, O.,
SCHRAUDER, M., SPETIUS, Z.V., WEISS, Y., NOWELL, G.M., OTTLEY, C.J. & CHINN, I. (2009).
HAURI, E.H., KAMINSKY, F.V., SOBOLEV, N.V. & Quantitative analysis of trace element
NAVON, O. (2009): High-Mg carbonatitic concentrations in some gem-quality diamonds. J.
microinclusions in some Yakutian diamonds – a Phys.-Cond. Matt. 21(36).
new type of diamond–forming fluid. Lithos 112 MELTON G.L., MCNEILL, J., STACHEL, T., PEARSON,
(S2), 648–659. D.G. & HARRIS, J.W. (2012): Trace elements in
gem diamond from Akwatia, Ghana and DeBeers
Pool, South Africa. Chem. Geol. 314, 1–8.

24
T. STACHEL – DIAMOND

MEYER, H.O.A. (1968): Chrome pyrope – an Mineral. Petrol. 70, 59–70.


inclusion in natural diamond. Science 160, 1446– PEARSON, D.G. & SHIREY, S.B. (1999): Isotopic
1447. dating of diamonds. SEG Reviews in Econ. Geol.
MEYER, H.O.A. (1987): Inclusions in diamond. In 9, 143–171.
Mantle xenoliths, (P.H. Nixon, ed.), John Wiley PEARSON, D.G., DAVIES, G.R. & NIXON, P.H.
& Sons Ltd., Chichester, 501–522. (1995): Orogenic ultramafic rocks of ultra-high
MEYER, H.O.A. & BOYD, F.R. (1969): Mineral pressure (diamond facies) origin. In Ultra-high
inclusions in diamonds. Carnegie Inst. Wash. pressure metamorphism, (R.G. Coleman & X.
Yearb. 67, 1446–1447. Wang, eds.), Cambridge University Press,
MEYER, H.O.A. & BOYD, F.R. (1972): Composition Cambridge, 456–510.
and origin of crystalline inclusions in natural PEARSON, D.G., SHIREY, S.B., HARRIS, J.W. &
diamonds. Geochim. Cosmochim. Acta 36, 1255– CARLSON, R.W. (1998): Sulphide inclusions in
1273. diamonds from the Koffiefontein kimberlite, S
MITCHELL, R.S. & GIARDINI, A.A. (1953): Oriented Africa: constraints on diamond ages and mantle
olivine inclusions in diamond. Amer. Mineral. 38, Re-Os systematics. Earth. Planet. Sci. Lett. 160,
136–138. 311–326.
MOORE, R.O. & GURNEY, J.J. (1985): Pyroxene PHILLIPS, D. & HARRIS, J.W. (1995): Geothermo-
solid solution in garnets included in diamonds. barometry of diamond inclusions from the De
Nature 318, 553–555. Beers Pool Mines, Kimberley, South Africa. Sixth
International Kimberlite Conference,
NAVON, O. (1999): Diamond formation in the
Novosibirsk, Extended Abstracts, 441–443.
Earth's mantle. In The P.H. Nixon Volume,
Proceedings of the VIIth International Kimberlite PHILLIPS, D., HARRIS, J.W. & VILJOEN, K.S. (2004):
Conference, (J.J. Gurney, J.L. Gurney, M.D. Mineral chemistry and thermobarometry of
Pascoe & S.H. Richardson, ed.), Red Roof inclusions from De Beers Pool diamonds,
Design, Cape Town, 584–604. Kimberley, South Africa. Lithos 77, 155–179.
NAVON, O., HUTCHEON, I.D., ROSSMAN, G.R. & POKHILENKO, N.P., SOBOLEV, N.V., REUTSKY,
WASSERBURG, G.J. (1988): Mantle-derived fluids V.N., HALL, A.E. & TAYLOR, L.A. (2004):
in diamond micro-inclusions. Nature 335, 784– Crystalline inclusions and C isotope ratios in
789. diamonds from the Snap Lake/King Lake
kimberlite dyke system: evidence of ultradeep and
NIMIS, P. & GRÜTTER, H. (2010): Internally
enriched lithospheric mantle. Lithos 77, 57–67.
consistent geothermometers for garnet peridotites
and pyroxenites. Contrib. Mineral. Petrol. 159, REGE, S., GRIFFIN, W.L., PEARSON, N.J., ARAUJO,
411–427. D., ZEDGENIZOV, D., O'REILLY, S.Y. (2010):
Trace-element patterns of fibrous and
NIMIS, P. & TAYLOR, W.R. (2000): Single clino-
monocrystalline diamonds: Insights into mantle
pyroxene thermobarometry for garnet peridotites.
fluids. Lithos 118, 313–337
Part I. Calibration and testing of a Cr-in-Cpx
barometer and an enstatite-in-Cpx thermometer. RICHARDSON, S.H. (1986): Latter-day origin of
Contrib. Mineral. Petrol. 139, 541–554. diamonds of eclogitic paragenesis. Nature 322,
623–626.
OGDEN, J. (2005): Diamonds, head hunters and a
prattling fool: The British exploitation of Borneo RICHARDSON, S.H., CHINN, I.L. & HARRIS, J.W.
diamonds. Gems and Jewellery 14, 67–69. (1999): Age and origin of eclogitic diamonds
from the Jwaneng kimberlite, Botswana. In The
O'NEILL, H.S.C. (1980): An experimental study of
P.H. Nixon Volume, Proceedings of the VIIth
the iron-magnesium partitioning between garnet
International Kimberlite Conference, (J.J. Gurney,
and olivine and its calibration as a
J.L. Gurney, M.D. Pascoe & S.H. Richardson,
geothermometer: corrections. Contrib. Mineral.
eds.), Reed Roof Design, Cape Town, 709–713.
Petrol. 72, 337.
RICHARDSON, S.H., SHIREY, S.B. & HARRIS, J.W.
O'NEILL, H.S.C. & WOOD, B.J. (1979): An (2004): Episodic diamond genesis at Jwaneng,
experimental study of the iron-magnesium Botswana, and implications for Kaapvaal craton
partitioning between garnet and olivine and its
evolution. Lithos 77, 143–154.
calibration as a geothermometer. Contrib.

25
T. STACHEL – DIAMOND

RICKARD, R.S., HARRIS, J.W., GURNEY, J.J. & SHIREY, S.B., RICHARDSON, S.H. & HARRIS, J.W.
CARDOSO, P. (1989): Mineral inclusions from (2004): Integrated models of diamond formation
Koffiefontein mine. In Kimberlites and related and craton evolution. Lithos 77, 923–944.
rocks. (J. Ross et al., eds), Blackwell, Carlton, SIMAKOV, S.K. (2006): Redox state of eclogites and
1054–1062. peridotites from sub-cratonic upper mantle and a
ROBINSON, D.N. (1979): Surface textures and other connection with diamond genesis. Contrib.
features of diamonds. Unpubl. PhD thesis, Mineral. Petrol. 151, 282–296.
University of Cape Town, Cape Town, South SMART, K.A., CHACKO, T., STACHEL, T.,
Africa, Vol. 1: pp. 221, Vol. 2: pp. 161. MUEHLENBACHS, K., STERN, R.A. & HEAMAN,
ROBINSON, D.N., SCOTT, J.A., VAN NIEKERK, A. & L.M. (2011): Diamond growth from oxidized
ANDERSON, V.G. (1989): The sequence of events carbon sources beneath the Northern Slave
reflected in the diamonds of some southern Craton, Canada: A δ13C–N study of eclogite-
African kimberlites. In Kimberlites and related hosted diamonds from the Jericho kimberlite.
rocks. (J. Ross et al., eds), Blackwell, Carlton, Geochim. Cosmochim. Acta 75, 6027–6047.
990–1000. SOBOLEV, N.V. (1977): Deep-seated inclusions in
RUDNICK, R.L. (1995): Eclogite xenoliths: samples kimberlites and the problem of the composition of
of Archean sea floor. Sixth International the upper mantle. AGU, Washington (Translated
Kimberlite Conference, Novosibirsk, Extended from the Russian edition, 1974).
Abstracts, 473–475. SOBOLEV, N.V. & SHATSKY, V.S. (1990): Diamond
RUDNICK, R.L. & NYBLADE, A.A. (1999): The inclusions in garnets from metamorphic rocks – a
thickness and heat production of Archean new environment for diamond formation. Nature
lithosphere: constraints from xenolith 343, 742–746.
thermobarometry and surface heat flow. In SOBOLEV, N.V., LAVRENT'EV, Y.G., POSPELOVA,
Mantle Petrology: Field Observations and High L.N. & SOBOLEV, V.C. (1969): Chrome pyrope
Pressure Experimentation: A tribute to Francis R. from Yakutian diamonds. Doklady Akademii
(Joe) Boyd (Y. Fei, C.M. Bertka & B.O. Mysen, Nauk SSSR 189, 133–136.
eds.), The Geochemical Society, Houston, 3–12.
SOBOLEV, N.V., BARTOSHINSKY, Z.V., YEFIMOVA,
RYAN, C.G., GRIFFIN, W.L. & PEARSON, N.J. E.S., LAVRENT'YEV, Y.G. & POSPELOVA, L.N.
(1996): Garnet geotherms – pressure-temperature (1970): Association of olivine, garnet and
data from Cr-pyrope garnet xenocrysts in volcanic chrome-diopside from Yakutian diamonds (in
rocks. J. Geophys. Res. 101, 5611-5625. Russian). Doklady Akademii Nauk SSSR 192,
SCHRAUDER, M. & NAVON, O. (1994): Hydrous and 1349–1353.
carbonatitic fluids in fibrous diamonds from SOBOLEV, N.V., LAVRENT'EV, Y.G., POKHILENKO,
Jwaneng, Botswana. Geochim. Cosmochim. Acta N.P. & USOVA, L.V. (1973): Chrome-rich garnets
58, 761–771. from the kimberlites of Yakutia and their
SHILOBREEVA, S., MARTINEZ, I., BUSIGNY, V., paragenesis. Contrib. Mineral. Petrol. 40, 39–52.
AGRINIER, P., LAVERNE, C. (2011): Insights into SOBOLEV, N.V., GALIMOV, E.M., IVANOVSKAIA,
C and H storage in the altered oceanic crust: I.N. & EFIMOVA, E.S. (1979): Isotopic
Results from ODP/IODP Hole 1256D. Geochim. composition of carbon from diamonds containing
Cosmochim. Acta 75, 2237–2255 crystalline inclusions. Doklady Akademii Nauk
SHIREY, S.B. & RICHARDSON, S.H. (2011) Start of SSSR 249, 1217–1220.
the Wilson cycle at 3 Ga shown by diamonds SOBOLEV, N.V., SNYDER, G.A., TAYLOR, L.A.,
from subcontinental mantle. Science 333, 434– KELLER, R.A., YEFIMOVA, E.S., SOBOLEV, V.N.
436. & SHIMIZU, N. (1998): Extreme chemical
SHIREY, S.B., CARLSON, R.W., RICHARDSON, S.H., diversity in the mantle during eclogitic diamond
MENZIES, A., GURNEY, J.J., PEARSON, D.G., formation: Evidence from 35 garnet and 5
HARRIS, J.W. & WIECHERT, U. (2001): Archean pyroxene inclusions in a single diamond. Int.
emplacement of eclogitic components into the Geol. Rev. 40, 567–578.
lithospheric mantle during formation of the
Kaapvaal Craton. Geophys. Res. Lett. 28, 2509–
2512.

26
T. STACHEL – DIAMOND

SOBOLEV, V.S. & SOBOLEV, N.V. (1980): New TAPPERT, R., STACHEL, T., HARRIS, J.W.,
proof on very deep subsidence of eclogitized MUEHLENBACHS, K., LUDWIG, T. & BREY, G.P.
crustal rocks. Doklady Akademii Nauk SSSR 250, (2005b): Subducting oceanic crust: the source of
683–685. deep diamonds. Geology 33, 565–568.
SPETSIUS, Z.V., BELOUSOVA, E.A., GRIFFIN, W.L., TAPPERT, R., STACHEL, T., HARRIS, J.W.,
O'REILLY, S.Y. & PEARSON, N.J. (2002): Archean MUEHLENBACHS, K., LUDWIG, T. & BREY, G.P.
sulfide inclusions in Paleozoic zircon megacrysts (2005c): Diamonds from Jagersfontein (South
from the Mir kimberlite, Yakutia: implications for Africa): messengers from the sublithospheric
the dating of diamonds. Earth. Planet. Sci. Lett. mantle. Contrib. Mineral. Petrol. 150, 505–522.
199, 111–126. TAYLOR, L.A., KELLER, R.A., SNYDER, G.A.,
STACHEL, T. (2001): Diamonds from the WANG, W.Y., CARLSON, W.D., HAURI, E.H.,
asthenosphere and the transition zone. Eur. J. MCCANDLESS, T., KIM, K.R., SOBOLEV, N.V. &
Mineral. 13, 883–892. BEZBORODOV, S.M. (2000): Diamonds and their
STACHEL, T. & HARRIS, J.W. (2008): The origin of mineral inclusions, and what they tell us: A
cratonic diamonds – constraints from mineral detailed "pull-apart" of a diamondiferous eclogite.
inclusions. Ore Geol. Rev. 34, 5–32. Int. Geol. Rev. 42, 959–983.
STACHEL, T., HARRIS, J.W., BREY, G.P. & JOSWIG, TAYLOR, W.R. (2004): Impurity measurements in
W. (2000): Kankan diamonds (Guinea) II: lower diamonds using IR. Rough Diamond Review 4,
mantle inclusion parageneses. Contrib. Mineral. 40–42.
Petrol. 140, 16–27. TAYLOR, W.R. & GREEN, D.H. (1989): The role of
STACHEL, T., HARRIS, J.W., TAPPERT, R. & BREY, reduced C–O–H fluids in mantle partial melting.
G.P. (2003): Peridotitic diamonds from the Slave In Kimberlites and related rocks. (J. Ross et al.,
and the Kaapvaal cratons – similarities and eds), Blackwell, Carlton, 592–602.
differences based on a preliminary data set. Lithos TAYLOR, W.R., JAQUES, A.L. & RIDD, M. (1990):
71, 489–503. Nitrogen-defect aggregation characteristics of
STACHEL, T., VILJOEN, K.S., MCDADE, P. & some Australasian diamonds: Time-temperature
HARRIS, J.W. (2004): Diamondiferous constraints on the source regions of pipe and
lithospheric roots along the western margin of the alluvial diamonds. Amer. Mineral. 75, 1290–
Kalahari Craton – the peridotitic inclusion suite in 1310.
diamonds from Orapa and Jwaneng. Contrib. TAYLOR, W.R., CANIL, D. & MILLEDGE, H.J.
Mineral. Petrol. 147, 32–47. (1996): Kinetics of Ib to IaA nitrogen aggregation
STACHEL, T., BREY, G.P. & HARRIS, J.W. (2005): in diamond. Geochim. Cosmochim. Acta 60,
Inclusions in sublithospheric diamonds: glimpses 4725–4733.
of deep earth. Elements 1, 73–78. TOMLINSON, E., DE SCHRIJVER, I., DE CORTE, K.,
STACHEL, T., HARRIS, J.W. & MUEHLENBACHS, K. JONES, A.P., MOENS, L. & VANHAECKE, F. (2005):
(2009): Sources of carbon in inclusion bearing Trace element compositions of submicroscopic
diamonds. Lithos 112 (S2), 625–637. inclusions in coated diamond: A tool for
understanding diamond petrogenesis. Geochim.
SUNAGAWA, I. (1984): Morphology of natural and
Cosmochim. Acta 69, 4719–4732
synthetic diamond crystals. In Materials science
of the Earth's interior, (I. Sunagawa, ed.), Terra TRULL, T., NADEAU, S., PINEAU, F., POLVÉ, M. &
Scientific, Tokyo, 303–330. JAVOY, M. (1993): C–He systematics in hotspot
xenoliths: implications for mantle carbon contents
TAPPERT, R. & TAPPERT, M.C. (2011): Diamonds in
and carbon recycling. Earth. Planet. Sci. Lett.
Nature: A Guide to Rough Diamonds. Springer,
118, 43–64.
New York, pp. 142.
WALTER, M.J., KOHN, S.C., ARAUJO, D.,
TAPPERT, R., STACHEL, T., HARRIS, J.W., SHIMIZU,
BULANOVA, G.P., SMITH, C.B., GAILLOU, E.,
N. & BREY, G.P. (2005a): Mineral inclusions in
WANG, J., STEELE, A. & SHIREY, S.B. (2011):
diamonds from the Panda kimberlite, Slave
Deep mantle cycling of oceanic crust: evidence
Province, Canada. Eur. J. Mineral. 17, 423–440.
from diamonds and their mineral inclusions.
Science 333, 54–57.

27
T. STACHEL – DIAMOND

WEISS, Y., KIFLAWI, I. & NAVON, O., 2010. IR London Ser. A – Math. Phys. Eng. Sci. 407, 219–
spectroscopy: Quantitative determination of the 238.
mineralogy and bulk composition of fluid WYLLIE, P.J. & HUANG, W.-L. (1976): Carbonation
microinclusions in diamonds. Chem. Geol. 275, and melting reactions in the system CaO-MgO-
26–34. SiO2-CO2 at mantle pressures with geophysical
WEISS, Y., GRIFFIN, W.L., BELL, D.R. & NAVON, O. and petrological applications. Contrib. Mineral.
(2011) High-Mg carbonatitic melts in diamonds, Petrol. 54, 79–107.
kimberlites and the sub-continental lithosphere. YAMAOKA, S., KOMATSU, H., KANDA, H. & S
Earth. Planet. Sci. Lett. 309, 337–347. ETAKA, N. (1977): Growth of diamond with
WELBOURN, C.M., ROONEY, M.L.T. & EVANS, rhombic dodecahedral faces. Journal of Crystal
D.J.F. (1989): A Study of diamonds of cube and Growth 37, 349–352.
cube-related shape from the Jwaneng Mine. ZEDGENIZOV, D.A., REGE, S., GRIFFIN, W.L., KAGI,
Journal of Crystal Growth 94, 229–252. H. & SHATSKY, V.S. (2007): Composition of
WILKS, J. & WILKS, E. (1991): Properties and trapped fluids in cuboid fibrous diamonds from
Applications of diamond. Butterworth-Heinemann the Udachnaya kimberlite: LAM–ICPMS
Ltd, Oxford, pp. 525. analysis. Chem. Geol. 240, 151–162.
WOODLAND, A.B. & KOCH, M. (2003): Variation in ZHANG, R.Y., LIOU, J.G., ERNST, W.G., COLEMAN,
oxygen fugacity with depth in the upper mantle R.G., SOBOLEV, N.V. & SHATSKY, V.S. (1997):
beneath the Kaapvaal craton, Southern Africa. Metamorphic evolution of diamond-bearing and
Earth. Planet. Sci. Lett. 214, 295–310. associated rocks from the Kokchetav Massif,
WOODS, G.S. (1986): Platelets and the infrared- northern Kazakhstan. J. Metam. Geol. 15, 479–
absorption of Type-Ia diamonds. Proc. R. Soc. 496.

28

View publication stats

You might also like