+ All Categories
Home > Documents > Thesis new

Thesis new

Date post: 24-Jan-2023
Category:
Upload: khangminh22
View: 0 times
Download: 0 times
Share this document with a friend
27
VU Research Portal Dating individual diamond growth zones Gress, M.U. 2020 document version Publisher's PDF, also known as Version of record Link to publication in VU Research Portal citation for published version (APA) Gress, M. U. (2020). Dating individual diamond growth zones: A first step towards quantifying the temporal evolution of the mantle carbon cycle. General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal ? Take down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. E-mail address: [email protected] Download date: 31. Jan. 2022
Transcript

VU Research Portal

Dating individual diamond growth zones

Gress, M.U.

2020

document versionPublisher's PDF, also known as Version of record

Link to publication in VU Research Portal

citation for published version (APA)Gress, M. U. (2020). Dating individual diamond growth zones: A first step towards quantifying the temporalevolution of the mantle carbon cycle.

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal ?

Take down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

E-mail address:[email protected]

Download date: 31. Jan. 2022

INTRODUCTION

Chapter 1. INTRODUCTION

1

CHAPTER 1

20

Chapter 1

1. Introduction

1.1 Rationale of the thesis

1.2 Diamonds

1.2.1 Origin

1.2.2 Xenocrystic relationship to kimberlites

1.2.3 Nitrogen content & aggregation

1.2.4 Inclusion paragenesis

1.2.4.1 Peridotitic (P-type) suite

1.2.4.3 Eclogitic (E-type) suite

1.2.4.4 Transitional suite

1.2.5 Constraints on carbon sources

1.3 General information

1.3.1 Geological setting

1.3.1.1 Jwaneng

1.3.1.2 Orapa – Letlhakane – Damtshaa

2. Research Objectives

2.1 What is the genetic origin of inclusions in diamond?

2.2 What are the number, timing and spatial scale of diamond-forming events?

2.3 Potential implications for the temporal evolution of the deep carbon cycle?

3. Methods

4. Outline

Previous pages. Orapa diamond mine (AK1) in NE Botswana, 09/2017. The open pit has a surface extent of 1.8 x 1.2 km and is currently ~ 250 m deep. The mine has a grade of ~ 88 carats per 100 tons with a daily production of ~ 22,000 tons.

INTRODUCTION

21

1. Introduction

Diamonds are […] by far more than the famous lyrics of a Broadway composition that made

them the object of desire for many. Their inimitable sparkle along with the physical properties,

as the hardest known natural material on Earth, has fascinated people for thousands of years.

However, these unique properties enable the deployment of diamond beyond the gem-stone

and jewellery industry in a broad range of applications e.g., optical/high-powered laser systems,

quantum information technology, invasive surgical treatment or in highly abrasive and corrosive

environments. In addition, diamond mining operations locally provide a platform for economic

stability that theoretically enables the local people to benefit from diamond revenues in various

sectors (education, health, etc., see Brook (2012)).

Scientifically, diamond’s strength and inert nature enables it to endure severe conditions

prevailing inside Earth’s mantle (100’s of kilometres deep and temperatures > 1000°C). The

inertness with its surrounding environment makes it the ideal carrier for ambient mantle material

that has been encapsulated during diamond growth and subsequently shielded from alteration

until its accidental transport to the surface through kimberlite magmatism. The study of these

time-capsules and their inclusions provides us with an image of the structure and processes

prevailing in the otherwise inaccessible interior of our planet. Given the age of some diamonds

(up to 3,500,000,000 years) they allow us to trace the evolution of Earth since its early existence.

1.1 Rationale of the thesis

In the past decades, research has made major advances in understanding the conditions required

for diamond formation, its crystal properties and for the identification of the source carbon, as

summarized in the following sections. Diamonds also enable scientists to trace major processes

that affected Earth in the past to better understand processes occurring at present, such as plate

tectonics (Shirey & Richardson, 2011) or the scale of element (re-)cycling over time (Shirey et al.,

2013; 2019; Howell et al., 2020). Such studies, however, require precise dating of diamonds

before correct implications can be drawn, and the timing of diamond formation has been the

subject of active research.

Botswanan diamonds are known for their unusual predominance of the eclogitic

paragenesis (see 1.2.4.4) that is ascribed to subduction of oceanic lithosphere into Earth’s mantle

(Ireland et al., 1994; Foley et al., 2002; Stachel et al., 2004b). Thus, in this thesis the study of

diamonds from several Botswanan locations (see 1.4) enables discussion of three controversial

aspects of diamond petrology that are expanded upon in 1.5 at the end of this chapter:

- 1st objective: The genetic origin of diamond inclusions

- 2nd objective: The number, timing and spatial scale of diamond-forming events

- 3rd objective: The implications for the temporal evolution of the deep carbon cycle

CHAPTER 1

22

The following sections provide a general overview of lithospheric diamonds, the geological

background and information on the samples of this study to introduce the reader to the topic.

For further information digging deeper into the topic, the text refers to the extensive literature

reviews available.

1.2 Diamonds

Elemental carbon occurs in two major forms in nature, graphite and diamond. Graphite is one of

the softest minerals, consisting of layers of weakly bound carbon. In contrast, the polymorph

diamond with a cubic crystal structure is the hardest known mineral that widely occurs in extra-

/terrestrial bodies (Allamandola et al., 1993; Alexander, 2001) and allows major implications to

be determined on the processes responsible for their formation (see compilation in Pearson et al.

(2003)).

Depending on carbon saturation in the diamond-forming fluid, diamond occurs as

monocrystalline crystals (like the samples in this study), polycrystalline aggregates, or with fibrous

coats (see overview in Tappert & Tappert (2011)). The morphology of monocrystalline diamonds

(Fig. 1) ranges from octahedral to (rounded) dodecahedral habits, cubes, intergrown aggregates

or as irregular forms with no identifiable crystal shape (see classification scheme in Harris et al.

(1975)).

Experiments with aqueous fluids and hydrous silicate melts have established that diamond

growth rates that mimic natural octahedral or polycrystalline morphology can be as high as 14

mm/year (Palyanov et al., 2015; Bureau et al., 2016). While diamond in its pure carbon form is

colourless, the presence of impurities in the crystal lattice that substitute for carbon e.g., nitrogen

(0 – 3000 ppm; Collins (1980)) or boron (< 10 ppm; Gaillou et al. (2012); Smith et al. (2018)), as

well as other lattice defects, so called optical centres, caused for example by radiation damage,

produce characteristic colours (Fig. 1) due to the selective absorption of light in the visible (or

UV/NIR) spectrum (see compilation in Breeding & Shigley (2009)).

1.2.1 Origin

The vast majority (~ 95%; Shirey et al. (2013)) of natural monocrystalline diamonds on Earth

formed in the lithospheric mantle beneath Archaean cratons (Fig. 2) that represent the oldest

regions of Earth’s crust exposed on the surface (4.0 – 2.5 Ga; Bowring et al. (1990)). Subsequent

accretion of Proterozoic/Phanerozoic crustal material via subduction at destructive plate

boundaries promoted the growth of continents and transferred surficial and crustal material into

the mantle. The Archaean nuclei (Fig. 3) have remained stable for > 2.5 Ga because of extensive

partial melt extraction of magmaphile elements such as Fe from the underlying sub-continental

lithospheric mantle (SCLM) to the crust (Schulze, 1986; Kesson & Ringwood, 1989). This melt

depletion caused lower density of the SCLM and, after cooling, lower heat production and hence

INTRODUCTION

23

Figure 1. Natural octahedral, dodecahedral and cuboid morphology of Orapa diamonds (left) and natural colour variation from yellow-brown (colloquially called ginger), yellow to colourless in Jwaneng diamonds (right).

temperature, resulting in a positive buoyancy relative to younger, more fertile/undepleted

asthenospheric mantle (Jordan, 1975; Boyd & McCallister, 1976; Lee, 2003).

Today, the asthenospheric mantle generates oceanic lithosphere through partial melting

(~ 15% – 20%) at mid-ocean ridges. During the Archaean (> 2.5 Ga), the prevailing mantle was

100°C – 300°C warmer resulting in higher degrees (up to 40%) of melt extraction from the

mantle leaving a residual harzburgite (Ringwood, 1991; Pearson & Wittig, 2008; Aulbach, 2012;

Herzberg & Rudnick, 2012). Trace element patterns of garnets from peridotitic mantle xenoliths

record lateral and vertical compositional heterogeneities in the SCLM and indicate that the

originally depleted harzburgites and lherzolites experienced several episodes of metasomatic re-

fertilisation (Carlson et al., 1999; Griffin et al., 2003a; Simon et al., 2007).

Eclogite is the other notable component of SCLM. Consideration of eclogitic xenoliths

suggests that they either formed entirely within the mantle (Hills & Haggerty, 1989; Smyth et al.,

1989; Griffin & O’Reilly, 2007) or that the SCLM was infiltrated by residual products

(metabasalts/metagabbros that experienced prior melt extraction) that were originally derived

from subducting oceanic lithosphere retaining strong geochemical records of its heterogeneity;

i.e., hydrothermally altered lavas-dykes – gabbros – peridotite (Helmstaedt & Doig, 1975;

MacGregor & Manton 1986; Helmstaedt & Schulze, 1989; Jacob et al., 1994; 2003; Aulbach et

al., 2007; Aulbach & Jacob, 2016).

The base of the SCLM keel (Shirey et al., 2013) is at depths of 250 – 300 km (Jordan, 1988;

Griffin et al., 2003a). Thus, the SCLM extends into the diamond stability field (Fig. 3) at pressures

of 4.5 GPa – 6 GPa (140 – 200 km) and temperatures between 900°C – 1400°C (Boyd & Gurney,

1986; Haggerty, 1999; Stachel & Harris, 2008). Under these conditions, diamonds are formed

metasomatically in the SCLM via redox reactions involving silicate or carbonate rich liquids which

CHAPTER 1

24

Figure 2. Distribution of Archaean cratons worldwide after Pearson & Wittig (2008).

INTRODUCTION

25

Figure 3. Structure of continental crust (top) subdivided into an Archaean nucleus (rose) surrounded by Proterozoic (purple) and younger (light yellow) continental crustal units together with the schematic structure of the sub-continental lithospheric (SCLM) and convecting asthenospheric mantle, after Shirey et al. (2013). Physical conditions for diamond growth in the Earth (bottom), after Cartigny et al. (2014).

CHAPTER 1

26

are commonly referred as high-density C-H-O fluids or low degree melts at supra-solidus

conditions (Deines, 1980; Schrauder & Navon, 1993; Sobolev et al. 1998; 2009; Stachel et al.,

2005; Stachel & Luth, 2015). Henceforth, the term fluid is used in this research in context with

diamond-forming processes, because distinction between the role of high-density fluids and

melts is difficult to resolve.

Several studies recognised a small fraction of sublithospheric or so-called super-deep

diamonds that formed in the lower mantle at depths exceeding 660 km but that are not further

discussed in this thesis (Harte et al., 1999; McCammon, 2001; Stachel & Harris, 2009; Harte,

2010; Smith et al, 2016; 2018). Other lithospheric occurrences include ultrahigh-pressure

diamonds formed in the context of collisional orogenic belts (Sobolev & Shatsky, 1990;

Ogasawara, 2005) or impact craters from meteorites (Masaitis, 1995).

1.2.2 Xenocrystic relationship to kimberlites

As outlined in section 1.3.1, lithospheric diamonds form in the SCLM via metasomatic processes

and reside in the stable parts beneath Archaean cratons. At the surface, these diamonds occur

as xenocrysts (pre-existing crystals in an igneous host) in kimberlite (or lamproite) volcanic pipes,

or due to their erosion in placer deposits. Accordingly, they were transported to the surface

during kimberlite magmatism that can imprint surface textures through oxidation (e.g., trigons,

pits, hillocks; see overview in Tappert & Tappert (2011)) on monocrystalline diamonds

(Fedortchouk et al., 2019).

Kimberlites are ultrabasic, Si-undersaturated magmas, low in Al and Na, rich in CO2 and

H2O that are generated at the base of the lithosphere (cf, lamproites) or beneath (Kjarsgaard et

al., 2009; Woodhead et al., 2019; Pearson et al., 2019). During their rapid ascent to the surface,

within a few hours (O’hara et al., 1975; Edwards & Russell, 1998), the melts sample prevailing

crustal and mantle xenoliths and xenocrysts (such as diamond) that record the composition of

the local SCLM (Pearson et al., 2019). Kimberlites are distributed globally and some argue they

originate from a largely isolated mantle reservoir that evolved over ~ 2.5 Ga and partly

experienced massive contamination by subducted material in the Mesozoic (Woodhead et al.,

2019).

1.2.3 Nitrogen content & aggregation

The nitrogen content and its aggregation state within the diamond plays an important role in

determining diamond colour. Importantly, this crystal lattice defect facilitates a classification

scheme (Evans, 1992; Mainwood, 1994). Nitrogen is thought to be compatible in diamond during

precipitation under all known redox conditions (Thomassot et al., 2007; Stachel & Harris, 2009;

Petts et al., 2015; Hogberg et al., 2016) and diffusion takes place on a scale of 0.1 µm at SCLM

conditions, over > 4.5 Ga (Koga et al., 2003). Through the application of non-destructive Fourier

Transform Infrared Spectroscopy (FTIR), the nitrogen content and its aggregation state can be

quantitatively determined.

INTRODUCTION

27

Diamonds with no detectable nitrogen content by FTIR (usually ≤ 10 at.ppm) are classified

as Type II, whereas diamonds with higher content are classified as Type I (see overview in Tappert

& Tappert (2011)). Type I diamonds can be subdivided based on the manner of the nitrogen

coordination in the diamond crystal lattice. During diamond precipitation from a fluid, nitrogen

is included into the crystal lattice as single (or isolated) substitutional atoms, so called C-centres.

A diamond containing the majority of nitrogen as C-centres is known as Type Ib. At prevailing

mantle conditions (900°C – 1400°C and 4.5 GPa – 6 GPa) diffusion on a relatively short length

scale (~ 30 – 100 nm; Koga et al. (2003)) of nitrogen atoms within the diamond lattice leads to

progressive aggregation as pairs of nitrogen atoms, so called A-centres, and sub-sequently to

four nitrogen atoms surrounding a vacancy, so call B-centres. A diamond containing the majority

of nitrogen as A-centres is known as Type IaA (Boyd et al., 1994) and as B-centres, it is known

as Type IaB (Boyd et al., 1995). The majority of natural diamonds are Type IaAB i.e., a combination

of A- and B-centres. The conversion rate between these aggregation states is a function of

nitrogen content, temperature and time and can be used to calculate a relative mantle storage

temperature (i.e., depth) or mantle residence time for the diamond when three of these four

parameters are known (Taylor et al., 1990; Mendelssohn & Milledge, 1995; Leahy & Taylor,

1997). Importantly, these are average temperatures experienced over the diamond’s entire

residence in the mantle lithosphere. Since N content and aggregation can be quantified by FTIR,

calculations on relative diamond age can be coupled with the thermobarometry determined on

co-genetic inclusions (Evans & Harris, 1986; Harris, 1992; Navon, 1999; Stachel & Harris, 2008),

while insight into mantle storage temperature can be constrained by radiometric inclusion dating

(Pearson et al., 1998; Aulbach et al., 2009b; Smit et al., 2016a).

Combined cathodoluminescence (CL) and FTIR (later coupled with C-N isotope) studies on

central growth plates of diamonds (Fig. 4; Appendices B – F) reveal the growth history of

diamonds and have established that many individual diamonds record evidence of resorption and

marked internal zonation (Bulanova, 1995; Howell et al., 2012a). These observations indicated

marked variations in the composition of diamond-forming fluids that formed the individual

growth zones (Wiggers de Vries et al., 2013; Hogberg et al., 2016; Smit et al., 2016b) and they

demonstrate a potentially complex growth history within the diamond that will affect the N

aggregation systematics (Kohn et al., 2016).

However, the theoretical understanding of N aggregation during mantle residence still

provides a good first approximation for relative diamond ages unless they were stored for

extended periods (100’s Myr) in unusually cold (< 900°C) or warm (> 1300°C) lithosphere.

(Richardson & Harris, 1997; Pearson et al., 1998; 1999; Navon, 1999; Aulbach et al., 2009b; Smit

et al., 2016b; Timmerman et al., 2017). Importantly, however, the high thermal stability of the

lithosphere means that regional cooling rates are low (0.4°C – 1°C per Myr; Bedini et al., 2004;

Michaut & Jaupart, 2004) so that temperature slowly decrease, unless there is a major thermal

perturbation due to a tectono-magmatic event.

CHAPTER 1

28

Figure 4. Cathodoluminescence (CL) image (bottom) and light microscope image (top) of a central diamond

plate. Blue CL colours correspond to different levels of N impurities in the diamond crystal lattice. Note the

concentric growth structure of individual growth zones and the location of the inclusions within the diamond.

For further examples, see Appendices B-F.

INTRODUCTION

29

1.2.4 Inclusion paragenesis

Inclusions studied in this thesis are generally between 50 – 500 µm in (spherical) size and 1 – 150

µg in weight. Diamond inclusions (silicates, sulphides, oxides) reflect the prevailing composition

of the growth environment in the SCLM which allows for classification into different genetic

groups. Understanding the genetic origin of the inclusions and their time of formation are crucial

aspects to interpret processes within the SCLM and to constrain its structure (see objectives in

1.5.1 & 1.5.2).

Initial distinction between silicate mineral assemblages is based on differences in colour (Fig.

5) between, for example, eclogitic pyrope-almandine garnet (orange) and peridotitic pyrope

garnet (purple) inclusions (see overview in Tappert & Tappert (2011)). Sulphide inclusions are

almost ubiquitously associated with a characteristic rosette fracture system around the inclusion,

however, their paragenesis cannot be discriminated by optical examination. The paragenesis of

sulphide inclusions is discriminated by Ni and Os contents (Yefimova et al., 1983; Bulanova et al.,

1996; Pearson et al., 1998; Pearson & Shirey, 1999). At high temperatures in the mantle,

sulphides are present as mono-sulphide solid solution (mss). During cooling after kimberlite

eruption and transport to the near surface, low temperature phases such as pentlandite

((Fe,Ni)9S8), pyrrhotite (Fe1-xS), chalcopyrite (CuFeS2) and pyrite (FeS) are exsolved.

In general, diamond-forming fluids are considered to perturbate the most incompatible

elements (e.g., LREE or large ion lithophile elements) in the protolith mineralogy and allow to

make inferences on the fluid composition, while the original protolith signature is reflected in

more compatible elements (e.g., major or high field-strength elements; see Stachel et al.

(2004b)). Geochemical analysis of the inclusions allows a more precise classification of the

diamonds (see review in Stachel & Harris (2008)). In this respect, two main inclusion parageneses

have been recognised, peridotitic (P-type) and eclogitic (E-type), that relate to the two dominant

types of mantle rocks, peridotite and eclogite (Sobolev, 1977; Meyer, 1987; Bureau et al., 2016).

Mantle peridotite and associated minerals can be subdivided into a predomi-nantly

harzburgitic, lherzolitic and minor wehrlitic suites (Dawson & Stephens, 1975; Stachel & Harris,

2008). A third, less abundant transitional paragenesis, websteritic (W-type) is also recognised

(Gurney, 1984; Sobolev, 1984). The relative abundance of the parageneses varies from mine to

mine (Stachel & Harris, 2008; 2009), hence, the mineral abundance studies of this research (see Chapters 3, 5, 6) will add valuable insight for the distribution in Botswanan diamonds.

1.2.4.1 Peridotitic (P-type) suite

Peridotitic inclusions include greenish to colourless olivine, pyrope garnet, ortho- and

clinopyroxene (e.g., emerald green Cr-diopside), brown to black chromite, as well as Ni- (> 12

wt.%) and Os- (2 – 30 ppm) rich sulphides. With respect to garnet the classification is based on

diagnostic Ca and Cr contents (see classification in Grütter et al. (2004)).

Harzburgitic purple coloured (G10) pyrope garnets (Fig. 5) are characterised by low CaO (0

– 3 wt.%) and high Cr2O3 (1 – 22 wt.%) contents (Grütter et al., 2004). It was previously assumed

CHAPTER 1

30

Figure 5. Purple peridotitic garnet inclusion with imposed cubic symmetry (top left) and inside host diamond next to metallic sulphide inclusion that can be recognised by a black rosette around the inclusion (bottom right); orange eclogitic garnet inclusions (top right) inside host diamond; metallic sulphide inclusion and colourless silicates inside an octahedral diamond (bottom left).

that the formation of diamonds in residual harzburgitic portions of the SCLM (~ 90% garnet)

were restricted to the Archaean with ages between 3.5 – 3.1 Ga (Richardson et al., 1984; Gurney

et al., 2010), however, Koornneef et al. (2017) show that different styles of large-scale

magmatism resulted in peridotitic diamond formation in the Archaean and in the Proterozoic.

Lherzolitic inclusions include purplish-red pyrope garnet (G9; Grütter et al. (2004)), emerald

green Cr-diopside, colourless to pale yellow olivine and orthopyroxene, as well as brownish

chromite. Lherzolitic SCLM is thought to have been metasomatically re-fertilised by addition of

basaltic components to former residual harzburgite (Simon et al., 2007; Tang et al., 2013).

Diamonds formed in this substrate appear younger (~ 2.03 – 0.72 Ga) than harzburgitic

diamonds (Richardson & Shirey, 2008; Richardson et al., 2009; Aulbach et al., 2018). It has been

shown by Koornneef et al. (2017), however, that some of these ages represent a mix of two

discrete populations. Older lherzolitic ages up to 2.7 Ga were also found in lherzolitic diamonds

from the ‘Group II’ Voorspoed kimberlite near Kroonstad, South Africa, by researchers from the

VU (manuscript in preparation) and young peridotitic sulphide inclusion ages (69 ± 30 Ma;

Pearson et al. (1998)) have also been reported. The origin and timing of lherzolitic diamond

growth therefore remains an open field for further research.

INTRODUCTION

31

1.2.4.3 Eclogitic (E-type) suite

Eclogitic inclusions, that constitute the majority of samples in this research, include orange

pyrope-almandine garnet (Fig. 5), pale green omphacite, blue kyanite, colourless coesite, black

elongated rutile and corundum. The garnets (G3) are low in Cr2O3 (< 1 wt.%) with a broad range

in CaO (6 – 32 wt.%; Grütter et al. (2004)). Sulphides have low Ni contents (< 12 wt.%) and

high Re/Os ratios due to their low (< 800 ppb) Os contents (see Pearson et al. (1998)).

Eclogitic inclusions have been shown to have geochemical signatures inherited from Earth’s

near surface such as Eu anomalies (Jacob, 2004; Weiss et al. 2015; Aulbach & Jacob 2016), highly

variable oxygen isotope ratios (Schulze et al., 2003; Sobolev et al., 2011; Ickert et al., 2013) and

mass independently fractionated (MIF) sulphur signatures from the Archaean (Chaussidon et al.,

1987; Farquhar et al., 2002; Thomassot et al., 2009; Smit et al., 2019a) that established a general

link to subduction (see review in Shirey et al. (2013)). Inclusion ages range from ~ 2.9 – 0.1 Ga

(Richardson et al., 2004; Smit et al., 2016a; Timmerman et al., 2017) and suggest the occurrence

of multiple distinct generations (see review in Gurney et al. (2010)) that potentially allow tracing

of subduction and major plate tectonic processes (Shirey & Richardson, 2011).

1.2.4.4 Transitional suite

This suite includes websteritic and pyroxenitic inclusions, predominantly garnet (orange – brown),

ortho- (colourless) and clinopyroxene (light green) and sulphides. Garnets fall in the G4 and G5

categories after Grütter et al. (2004), sulphides overlap with compositions intermediate between

E- and P-type (Pearson et al., 1998). Websteritic mantle appears related to re-fertilisation

reactions of slab-derived melts with peridotitic SCLM (Aulbach et al., 2002; Taylor et al., 2003;

Bodinier et al., 2008; Smit et al., 2014) or the occurrence of reaction envelopes around an

ascending mantle plume (Sleep, 2006; Viljoen et al., 2018). The Orapa cluster, part of this study,

is one of the few locations on the Kalahari Craton where a websteritic diamond suite is present

(Gurney et al., 1984; Deines et al., 1993; Viljoen et al., 1998, 2018; Aulbach et al., 2002; Stachel

& Harris, 2008; Helmstaedt et al., 2010) enabling this study to provide new age and petrogenetic

constraints on this paragenesis.

1.2.5 Constraints on carbon sources

Carbon isotope ratios determined on distinct diamond growth zones are used to characterise the

diamond-forming fluid and to identify potential sources for carbon (Deines, 1980). In addition to

the energetically unfavourable direct conversion from graphite to diamond (Boyd et al., 1988;

1994), diamonds form through redox reactions i.e., by reduction of carbonate (CO2 and CO%&')

or by oxidation of methane (CH4) in variously buffered mantle environments (Shirey et al., 2013;

2019; Cartigny et al., 2014).

The carbon isotope composition is expressed in delta notation relative to Vienna PeeDee

Belemnite (VPDB), δ13CVPDB [‰] = (13C/12Csample / 13C/12CVPDB - 1) x 103. Similar notation is used for

nitrogen isotopes, δ15NAIR [‰] = (15N/14Nsample / 15N/14NAIR - 1) x 103, that are expressed relative to the

CHAPTER 1

32

composition of air (Cartigny et al., 2014; Hogberg et al., 2016). In terrestrial diamonds, δ13C

ranges from -41‰ to +5‰ with a mode at -5 ± 1‰ (Cartigny et al., 2014). Peridotitic diamonds

are relatively restricted to between -10‰ to -1‰, whereas eclogitic diamonds cover the full range

(Cartigny et al., 2014). Several models have been proposed to account for this isotopic variability

(see reviews in Cartigny et al. (2014); Shirey et al. (2019); Howell et al. (2020)) and include a

heterogeneous carbon isotope composition of the mantle (Deines, 1980; Deines et al., 1993),

different carbon sources in the mantle and the involvement of recycled (e.g., marine carbonates,

organic material) crustal material (Galimov, 1991; Kirkley et al., 1991; Li et al., 2019), high

temperature Rayleigh fractionation processes in the mantle (Javoy et al., 1986; Kirkley et al.,

1991; Cartigny et al., 1999; Thomassot et al., 2007) and to a lesser extent, changing P-T- fO2 and

pH conditions during diamond growth (Deines et al., 1989; Boyd et al., 1992; Sverjensky &

Huang, 2015).

In support of Jablon & Navon (2016), recent findings of Krebs et al. (2019) suggest a similar

origin between fibrous and monocrystalline diamonds involving common fluids. Their various

end-members (saline, high-/low-Mg carbonatitic and silicic; Schrauder & Navon, 1994; Weiss et

al., 2009) were previously only attributed with the genesis of fibrous diamonds (Navon et al.,

1988).

1.3 General information

The diamonds used in this research are from the 93.1 Ma Cretaceous Orapa kimberlite cluster in

NE Botswana (Allsopp et al., 1989; Davis, 1977), including Orapa (OR), Letlhakane (LK) and

Damtshaa (DM) mines (see Chapter 5) that are 20 to 30 km apart and extend over an area of ~

450 km2. Additional diamonds are from the 235 ± 4 Ma kimberlites (Kinny et al., 1989) at

Jwaneng (JW) mine (see Chapters 2 to 4), Earth’s richest diamond mine by value (Field et al.,

2008). Jwaneng is located 360 km to the southwest of the Orapa cluster in SE Botswana. The

distance between these locations allows a detailed comparison of diamond-forming processes

occurring in the SCLM on a regional scale.

In order to obtain a representative set of inclusion-bearing diamonds from these locations,

several sampling campaigns were undertaken at the Diamond Trading Company Botswana

(DTCB) in Gaborone in the past 5 years. This provided the opportunity to perform an inclusion

abundance study on run-of -mine production (see Chapters 3, 5 and 6). Such studies are rare

due to the restricted access to the samples. Sampling concentrated on diamonds containing

multiple large (> 50 μm) garnet (gnt), clinopyroxene (cpx) and sulphide inclusions. In total, over

370,000 gem to near-gem quality diamonds (~ 194,000 carats) ranging in size from 0.1 to 0.8

carat (DTC sieve classes +07, +09, +11 and 3 grainers) were characterised visually for their

inclusion (> 10 µm in size) content using binocular microscopes by a total of 12 researchers.

INTRODUCTION

33

1.3.1 Geological setting

Geologically, Jwaneng is situated near the NW margin of the Archaean Kaapvaal Craton, whereas

the Orapa cluster is located on the boundary between the Proterozoic Kheis-Okwa-Magondi

Mobile Belt and the Archaean Zimbabwe Craton (Fig. 6). Together with the Limpopo Mobile Belt,

all these geological units are the major constituents of the well-studied Kalahari Craton (Jacobs

et al., 2008) and its underlying SCLM (see review in Griffin et al. (2003b)) that has an extended

history as a part of several super-continents e.g., Rodinia, Gondwana or Pangea (Oriolo & Becker,

2018).

The Kaapvaal Craton (1.2 x 106 km2) formed between 3.7 – 2.7 Ga and consists of two

shields, the ~ 3.7 Ga Witwatersrand block in the east and the ~ 3.2 Ga Kimberley block in the

west that are in juxtaposition along the N-S trending Colesberg magnetic lineament (see review

in Thomas et al. (1993)). The ~ 3.64 Ga Ancient Gneiss Complex and the 3.5 – 3.2 Ga Barberton

Greenstone Belt in the Witwatersrand block represent the oldest shield in southern Africa and

carry remnants of the first continental crust that formed between ~ 3.2 to 3.08 Ga (Armstrong

et al., 1990; de Wit et al., 1992; Thomas et al., 1993; Moser et al., 2001). The oldest harzburgitic

diamond inclusions formed around this time (3.2 – 3.3 Ga) and indicate that the lithospheric

mantle in this area was already > 150 km thick (Richardson et al., 1984; de Wit et al., 1992). The

oldest exposed crystalline basement rocks in the Kimberley block are the ~ 3.18 – 2.88 Ga

Kraaipan and Amalia greenstone belts (Schmitz et al., 2004). The Witwatersrand block

amalgamated via west-dipping subduction/collision with the Kimberley block and stabilised the

Craton between 2.9 – 2.7 Ga (de Wit et al., 1992; Schmitz et al., 2004). Late Archaean (3.1 –

2.6 Ga) craton development is further characterised by extensional processes in the central and

southern parts resulting in the Dominion, Witwatersrand, Pongola and Ventersdorp sedimentary

basins (de Wit et al., 1992; Thomas et al., 1993).

The Zimbabwe Craton, to the northeast of the Kaapvaal, also consists of two different

blocks. The Tokwe Terrain is made up of ~ 3.5 – 3.3 Ga tonalitic gneisses with relicts of older

greenstone successions that are unconformably overlain by ~ 2.9 – 2.8 Ga conglomerates,

volcanics and sedimentary sequences; the Northwestern accretionary Terrain is separated from

the Tokwe block by 2.75 – 2.65 Ga greenstone belts (Griffin et al., 2003b). Both terrains

experienced potassic granite intrusions associated with the Great Dyke and its satellites at 2.69 –

2.57 Ga (Wilson, 1996). The Kaapvaal and Zimbabwe cratons subsequently amalgamated along

the Limpopo Province at ~ 2.7 – 2.6 Ga during the Limpopo Orogeny (Barton & van Reenen,

1992; de Wit et al., 1992; Roering et al., 1992; Goodwin, 1996; Frei et al., 1999).

The Limpopo Mobile Belt, marking their suture zone, is subdivided into three zones of which

the Central Zone experienced crustal stabilisation at ~ 3.4 – 3.27 Ga (Zeh et al., 2007; 2010) and

was proposed to represent a separate micro-continent carrying the 3.8 Ga granulite-

/amphibolite-facies Sand River Gneiss (Griffin et al., 2003b). The Northern and Southern Zones

are considered as high-grade metamorphic equivalents of the reworked Kaapvaal and Zimbabwe

craton margins (van der Meer et al., 2013). The SCLM beneath the Limpopo Mobile Belt is ~ 235

CHAPTER 1

34

Figure 6. Simplified geological map (after Pearson et al. (1998); Aulbach et al. (2009a)) with political borders, major cities, important structural units of the Kalahari Craton in Southern Africa and localities of the studied mines Jwaneng (JW), Orapa (OR), Letlhakane (LK) and Damtshaa (DM) and other diamond mines (see legend) referenced in the thesis.

20°

24°

28°

32°

20°

24°

28°

32°

34°30°

26°22°

18°14°

34°30°

26°22°

18°14°

Bushveld

Molopo

Farms

Vredefort C

rater

Pretoria

Cape Tow

n

Windhoek

Gaborone

Maseru

Mbabane

Maputo

NA

MIB

IA

SO

UTH

AFR

ICA

BO

TSW

AN

A

LESOTH

O

SW

AZI-

LAN

D

MO

ZAM

BIQ

UE

ZIMB

AB

WE

Namaqua NatalKheis Belt O

kwa

InlierMobile Belt

Limpopo M

obile Belt

Magondi Belt

Rehoboth

Subprovince

Colesberg

Great Dyke

Kaapvaal C

raton

Zimbabw

eC

raton

Lineament

Diam

ond Mines

OR

O

rapaD

M

Dam

tshaaLK

Letlhakane

JW

Jwaneng

V

VenetiaV

P Voorspoed

P P

remier/ C

ullinanD

P D

eBeers P

oolF

FinschK

K

offiefonteinJ

Jagersfontein

NS

EW

0400

km

JW OR

DM

LK

FD

PKJ

VP

P

V

Kalahari

Craton

INTRODUCTION

35

km deep and records melt depletion (up to 50%) between 50 – 130 km with Si-enriched

metasomatism characteristic of the Kaapvaal Craton, whereas the origin of deeper, more

enriched parts between 130 – 235 km is unclear (van der Meer et al., 2013). The upper mantle

part beneath the Limpopo Mobile Belt is also believed to represent the most depleted and least

modified SCLM that was sampled by kimberlites (Griffin et al., 2003b). The SCLM beneath the

Kaapvaal and Zimbabwe cratons extends to 250 – 300 km depth and thickened via melt additions

from the asthenosphere or underplating (James et al., 2001). The formerly depleted

harzburgitic/lherzolitic SCLM has been progressively re-fertilised by melt metasomatism during

episodes of massive intraplate magmatism (Griffin et al., 2003b) including the 2.06 Ga Bushveld

Large Igneous Province (LIP) and its satellites (e.g., at Molopo Farms), that carry world’s largest

reserves of platinum-group elements (Cawthorn & Walraven, 1998; Zeh et al., 2015), the 1.11

Ga Umkondo LIP (Hanson et al., 2004a) or the 0.18 Ga Karoo-Ferrar LIP (Svensen et al., 2012).

Post-Archaean and early Proterozoic evolution in the area was dominated by intra-cratonic

extension accompanied by sedimentation of the Transvaal Supergroup into which the Bushveld

LIP intruded. The Vredefort crater (~ 300 km in diameter), the largest verified impact crater on

Earth, formed around ~ 2.02 Ga (Roering et al., 1990). Potentially during the ~ 2.0 – 1.7 Ga

Eburian orogenesis, the Kheis-Okwa-Magondi Belts containing the ~ 2.0 Ga Mahalapy granite-

migmatite complex as basement began to amalgamate along the north western edge of the

Kaapvaal Craton (Thomas et al., 1993). The Kheis Belt consists of siliciclastic sediments and

volcanic rocks deposited on the passive margin of the Kaapvaal Craton, the Okwa Inlier basement

is overlain by low-grade clastic sediments and limestones, the Magondi Belt contains thick

sequences of meta-sediments and -volcanics, interpreted as deposits on the passive margin of

the Zimbabwe Craton (Griffin et al., 2003b). West of the Kheis-Okwa-Magondi Belt, the

Rehoboth Subprovince that reflects the transition from a volcanic arc into a back-arc setting

accreted between 1.8 – 1.75 Ga and marks the completion of the assembly of the 7 x 106 km2

Proto-Kalahari Craton (Jacobs et al., 2008). These mobile belts contain relatively fertile reworked

Archaean SCLM that mixed with Paleo-/Mesoproterozoic material (Griffin et al., 2003b).

Subsequent crustal growth e.g., during the 1.2 – 1.0 Ga Kibaran orogeny or formation of

the Namaqua-Natal Orogenic Belt, accompanied by widespread anorogenic magmatism between

1.4 – 1.1 Ga, occurred through accretion of Paleo-/Mesoproterozoic microcontinents or arcs and

resulted in the 13 x 106 km2 Kalahari Craton (Jacobs et al., 1993; 2008; Thomas et al., 1993). At

~ 0.95 – 0.9 Ga, the end of the Namaqua-Natal orogenesis, extensive continental rifting initiated

the breakup of the Rodinia supercontinent and culminated in the assembly of the Gondwana

supercontinent (Oriolo & Becker, 2018). Around this time, the Neoproterozoic Pan-African

Orogeny formed e.g., the Damara Belt, which is part of world’s largest sediment-hosted Cu (Pb-

Zn-Co-Ag-W-Ta) deposits in the Kalahari, West Congo and Central African Copperbelts (Thomas

et al., 1993; Oriolo & Becker, 2018; Borg & Gauert, 2018).

Following the Pan-African tectono-thermal event, the Cape Supergroup was deposited

along the southern passive continental margin of the Kalahari Craton in the Early Phanerozoic

CHAPTER 1

36

and formed the Cape Fold Belt in the Carboniferous to Early Permian; terrestrial sediments of the

Karoo Supergroup were subsequently deposited in an intracontinental basin in which the

extensive Karoo flood basalts mark the initiation of Gondwana breakup in the Early Jurassic

(Thomas et al., 1993). In this tectonic setting, kimberlitic magmatism occurred over a period of

520 – 80 Ma and transported diamonds from the SCLM to the surface at several locations (see

review in Field et al. (2008)). Further details on the geological background and current state of

research are provided in Chapters 3 to 6.

1.3.1.1 Jwaneng

Jwaneng mine, 120 km west of Gaborone, Botswana, is part of a field encompassing 28

kimberlites (Brown et al, 2008). The open-pit mine (Fig. 7) itself consists of three steep-sided

interconnected pipes (North, Centre and South Lobes) with a surface area of 54 ha and

dominated by volcaniclastic and dark pyroclastic kimberlite (2424D/K2).

At 235 ± 4 Ma, the kimberlites intruded into quarzitic shales and dolomites of the

Proterozoic Transvaal Supergroup and the pipe contains inclusions of unconsolidated Karoo

sediments that have been eroded in the meantime (Kinny et al., 1989; Brown et al, 2008; Field

et al., 2008). Notably, they are the only kimberlites in southern Africa with Permian emplacement

ages (Field et al., 2008). The Transvaal Supergroup in this area is also cross-cut by the products

of voluminous intraplate volcanism associated with the adjacent 1.1 Ga Umkondo Large Igneous

Province (LIP; Hanson et al., 2004a). Discovered underneath a 40 m thick layer of Kalahari

sediments in 1973, Jwaneng has been in operation since 1982 and is considered to be the world’s

richest diamond mine by value and produced ~ 11.9 million carats in 2018. According to

Debswana, the open-pit is currently mined in ‘Cut-8’ until 2024, increasing the depth of mine

from 400 to 650 m. The mine’s lifespan will be extended until at least 2035 by developing Cut-

9 that is expected to mine an additional 53 million carats.

Figure 7. Open pit of Jwaneng diamond mine that has a surface extent of 2.5 x 1.5 km; © Debswana. Jwaneng means ‘a place of small stones’ in Setswana.

INTRODUCTION

37

Geologically, Jwaneng is located on the south-eastern margin of the Central Kalahari sub-

basin in the north-western part of the Kaapvaal Craton that is underlain by ~ 200 km lithospheric

mantle (Simon et al., 2004; Brown et al., 2008). Xenolith studies are rare due to the high grade

of alteration but a recent geothermobarometric study on garnet lherzolites confirms previous

findings of a thick lithospheric mantle with estimated pressures of ~ 5 – 6.5 GPa and

temperatures between 1240°C – 1320°C that plot between a 42 – 45 mW/m2 geothermal

gradient (Branchetti, 2020). The data correlate with geothermobarometry from peridotitic

diamond inclusions reported in Stachel et al. (2004a). Disruption of the xenolith Sm-Nd isotope

system is indicated at the time of kimberlite eruption and at ~ 0.75 Ga; in the Lu-Hf system

garnets show disruption at ~ 1.57 ± 0.21 Ga (Branchetti, personal communication). This agrees

with considerations that major parts of the depleted SCLM were re-fertilised subsequent to

peridotitic diamond crystallisation (Schulze, 1995; Griffin et al., 1999). Overall, Jwaneng

diamonds and their inclusions have been the subject of extensive research and have contributed

to a better understanding of diamond-forming processes and the involved components (see

introduction of Chapters 2 to 4).

1.3.1.2 Orapa – Letlhakane – Damtshaa

The Orapa kimberlite cluster in the Central province of north-eastern Botswana hosts Debswana’s

Orapa, Letlhakane and Damtshaa mines, and (more recently) Lucara’s Karowe and Firestone

Diamond’s B/K11 mines, in addition to ~ 110 other non-mined kimberlites. At 93.1 Ma, the

kimberlites intruded into the early Proterozoic Magondi Mobile Belt that is underlain by Archaean

SCLM of the Zimbabwe Craton (Davis, 1977; Stiefenhofer et al., 1997). The 118 ha Orapa

kimberlite (2125A/K1) discovered in 1967, began production in 1971 and as of today, it is world’s

largest open-pit diamond mine. Mining in ‘Cut-4’ is expected to finish in 2030 with

considerations for a subsequent underground operation. According to Debswana, Orapa is

currently mined at a depth of 250 m aiming for 450 m in 2026.

Figure 8. Open pit of Letlhakane diamond mine (~ 1 km in diameter) partly filled with groundwater.

CHAPTER 1

38

The nearby 15.2 ha Letlhakane mine (2125D/K1 & D/K2) was in operation from 1977 –

2017 (Fig. 8) and its tailings are now processed extending the mine’s life-time for 20 years.

Damtshaa, a 24.5 ha satellite operation of Orapa, consists of four kimberlites that are planned

to be successively mined from 2003 until 2017 (2125B/K12), 2026 (2125B/K9) and thereafter

(2125B/K1 & B/K15). The Orapa, Letlhakane and Damtshaa mines yielded a combined production

of 12.2 million carats in 2018. In addition to Jwaneng, Letlhakane is the second major source of

samples for this thesis and hosts a varied suite of mantle xenoliths, mainly peridotites

(Stiefenhofer, 1993; Stiefenhofer et al., 1997). At Orapa, peridotite xenoliths are absent

potentially due to the highly weathered nature of the kimberlite and eclogite xenoliths dominate

at the surface (Shee & Gurney, 1977; McDonald & Viljoen, 2006). Previous studies of peridotitic

xenoliths from Letlhakane record pressure-temperature conditions comparable to xenoliths from

kimberlites from the Kaapvaal Craton and define a geotherm corresponding to 40 mW/m2

surface heat flow (Borst, 2012; Stiefenhofer, 1993; Stiefenhofer et al., 1997). The source region

of Letlhakane diamonds experienced a higher degree of partial melting relative to Orapa (Deines

& Harris, 2004).

Many of the first diamond and inclusion studies were initially undertaken using diamonds

from the Orapa kimberlite cluster. As for Jwaneng, the samples contributed to a better

understanding of diamond-forming processes, the abundance of inclusion paragenesis and

interpretation of carbon sources (see introduction of Chapter 5).

2. Research Objectives

The following sections describe the research objectives outlined in section 1.1 and summarise the

state of current knowledge on the corresponding topics. They also highlight major advances in

diamond research in the past decades.

2.1 What is the genetic origin of inclusions in diamond?

In order to correctly interpret diamond inclusion ages, it has to be established at what time

mineral inclusions formed with respect to the diamond host i.e., before (proto-) or during (syn-

chronous) diamond crystallisation. Moreover, it is important to establish if chemical equilibrium

existed between the inclusion and the surrounding mantle/diamond-forming media at the time

of encapsulation (syn-genetic).

The cubo-octahedral morphology imposed by diamond on incorporated inclusions (Fig. 5)

that naturally crystallise in different crystal systems such as the cubic (for garnet), monoclinic (for

cpx) or hexagonal (for monosulphide solid solution) has long been considered as proof for

inclusions precipitating synchronously (simultaneously) with the enclosing diamond growth zone

from the diamond-forming fluid (Futergendler & Frank-Kamenetsky, 1961; Orlov, 1977; Sobolev,

1977, Harris & Gurney, 1979; Meyer, 1987; Bulanova, 1995). However, the presence of proto-

genetic (pre-existing) inclusions, that are considered to have formed prior to diamond

INTRODUCTION

39

crystallisation lacking such cubo-octahedral morphology, raise concerns with this general

assumption (Bulanova, 1995; Taylor et al., 2003; Nestola et al., 2014; 2017; Agrosi et al., 2016;

Nimis et al., 2016).

Possible evidence for proto-genetic inclusions/de-coupling of diamond and inclusion-

forming sources is provided by sulphur isotope anomalies that indicate the presence of Archaean

surficial material in Proterozoic sulphide inclusions (Chaussidon et al., 1987; Farquhar et al., 2002,

Thomassot et al., 2009; Smit et al., 2016a; 2019a). These data undoubtedly indicate that older

sulphur was incorporated into sulphides that were subsequently trapped in diamonds and raise

the question of whether siderophile elements are recycled within the SCLM at the time of

diamond formation or whether the sulphides are protogenetic. However, modelling of chemical

diffusive equilibration between a fluid and a protogenetic inclusion under diamond-forming

conditions (1100°C and 5 GPa) previously demonstrated that isotopic resetting would generally

occur over geologically short time scales (Nestola et al., 2019). Conservative estimates of these

authors yielded diffusion rates for REE (e.g., Sm) of ~ 0.52 Myr to equilibrate a garnet grain of

100 µm radius. For a 500 µm spherical pyrrhotite (the most common sulphide inclusion in

diamond) Os diffusion (> 500°C) would have occurred on a similar time-scale ≤ 0.5 Ma (Brenan

et al., 2000).

Thus, the term ‘synchronous’ has been proposed to describe the relationship between

diamond and its inclusions (Nestola et al., 2017) because uncertainties in inclusion dating studies

commonly are in the order of 10’s of millions of years, far above the equilibration rates. Such

equilibrium state is also indicated by isochronous relationships between different inclusions from

the same diamond that have the same initial ratio (Westerlund et al., 2004; Aulbach et al., 2018).

In addition, syngenetic textures in protogenetic inclusions may also provide strong evidence that

inclusions co-crystallised with the diamonds.

In order to place inclusion ages in a geological context, the growth histories of the dated

individual diamonds must be determined. The complex growth histories between host diamond

and inclusion have been studied by an integrated approach (Wiggers de Vries et al., 2011)

combining CL imaging and electron backscatter diffraction (EBSD) to quantitatively identify the

crystallographic orientations of the diamond inclusions and to study the diamond-inclusion

interface in eclogitic samples from Jwaneng (see Chapter 2).

2.2 What are the number, timing and spatial scale of diamond-forming events?

Diamonds have a broad range of size, colour and morphology that indicates that they represent

multiple suites (see Chapter 3) that formed in different parts of the SCLM (Stachel et al., 2004b;

Stachel & Luth, 2015). Although only 1 – 10% of all diamonds are inclusion-bearing (Gurney,

1986; Deines & Harris, 2004; Stachel & Harris, 2008), their study allows general implications to

be reached for inclusion-free diamonds (Richardson et al., 1984). Inclusion dating studies (see

review in Smit & Shirey (2019)) initially focussed on the U-Pb (Welke et al., 1974; Kramers, 1979,

Eldrige et al., 1991) and Ar-Ar systems (Burgess et al., 1989; Phillips et al., 1989). Subsequently,

CHAPTER 1

40

the Sm-Nd (Richardson et al., 1984) and Re-Os (Pearson et al., 1998) isochron techniques,

coupled with spectroscopic (CL, FTIR) and compositional (C-N isotopes) properties of the diamond

host (Boyd et al., 1987; Taylor et al., 1990; Deines et al., 1993; Bulanova, 1995; Cartigny et al.,

1998) appeared to provide strong evidence of episodic diamond formation, recurring throughout

the Archaean and Proterozoic Eras (see reviews in Stachel & Harris (2008); Gurney et al. (2010);

Helmstaedt et al. (2010)).

Due to the low Os content in single sulphide inclusions and thus, low Os ion yield, the

analysis of single sulphide inclusions for Re-Os isotope systematics can be achieved using a

secondary electron multiplier (SEM) in peak-jumping mode as the detection unit in mass

spectrometry. This approach is not viable for the Rb-Sr or Sm-Nd isotope analysis of silicate

inclusions as precisions better than 0.05% are required for good age determinations. Hence,

either large individual (> mg) silicates (Smith et al., 1991) were analysed using Faraday Cups (FCs)

to establish model and isochron ages or garnet/cpx inclusions with similar characteristics were

pooled (10 – 100 µg per inclusion) in order to obtain a quantifiable signal for mass spectrometry

(Richardson et al., 1984; 1990).

Technical advances in isotope mass spectrometry through the deployment of latest

generation highly sensitive detectors (FCs attached to amplifiers with 1013 Ohm resistors;

Koornneef et al. (2013); (2014)) now allow dating of individual diamond silicate inclusions (see Chapters 4, 5) using the Sm-Nd system (Timmerman et al., 2017; Koornneef et al., 2017).

Together with the well-established Re-Os system for sulphide inclusions (Pearson et al., 1998),

the application of these different isotopic systems is capable of validating whether mineral

inclusions recovered from the same growth zone of a diamond are in equilibrium with the

surrounding mantle and diamond-forming fluid and thus, record the time of diamond growth

(see Chapter 6).

Yet, this combined approach of Sm-Nd and Re-Os dating on co-genetic silicate and sulphide

inclusions from the same diamond growth zone has not been undertaken before. Inclusions from

adjacent growth zones within a diamond will establish the time between individual diamond-

forming events. Examination of diamonds from adjacent kimberlites (Orapa, Letlhakane,

Damtshaa) within a cluster will provide a local (10’s of km) snapshot of diamond formation (see Chapter 5). Comparison to other kimberlites (Jwaneng) will establish if regional scale (100’s of

km) processes occurred across the Kalahari Craton (see Chapters 5, 6).

2.3 Potential implications for the temporal evolution of the deep carbon cycle?

Carbon, along with hydrogen, nitrogen and oxygen, is one of the basic constituents for life on

Earth and by tracing its cycling between distinct reservoirs (atmosphere, hydrosphere, lithosphere)

it is possible to draw major conclusions regarding planetary scale differentiation and plate

tectonics (Shirey & Richardson, 2011). However, the large carbon reservoirs inside Earth (~ 1016

to 1017 tons carbon; Dasgupta & Hirschmann, 2010) are particularly difficult to access (Shirey et

al., 2019; Howell et al., 2020).

INTRODUCTION

41

Although diamonds represent only a fraction (83 ppt; Kelemen & Manning, 2015) of Earth’s

mantle, they have the potential to address fundamental questions of how Earth has evolved.

According to Shirey et al. (2013), ~ 30% of diamonds worldwide contain crustal/sedimentary

carbon recycled from Earth’s surface and transferred into the mantle via subduction. By relating

eclogitic inclusion ages to C and N isotope characteristics of individual diamond growth zones,

information concerning the deep cycling of carbon and other volatile species may be obtained

(see Chapters 4 to 6). Implications derived from the scale and style of diamond formation may

ultimately provide a first step for the study of the temporal evolution of the deep carbon cycle

(see Chapter 7) in order to contribute to models of the past, present and future carbon cycle.

3. Methods

Combination of multiple spectroscopic and geochemical techniques introduced in the past

decades have proven to effectively extract information from diamonds and their inclusions. The

methods deployed in this study will integrate these applications in combination with state-of-

the-art technology in radiogenic isotope dating. In this respect, a ‘standard procedure’ (Fig.9) has

been established according to the studies of Wiggers de Vries et al. (2011; 2013), Smit et al.

(2016a) and Timmerman et al. (2017). Technical remarks on the deployed methods

supplementing the sketch and reference to the corresponding methods section of Chapters 2 to 6 can be found in Appendix A.1.

4. Outline

This PhD thesis consists of five main Chapters (2 to 6) that have been individually published in (or

submitted to) peer-reviewed journals and they systematically assess the objectives of the PhD

project (outlined in 1.5). The current section gives a short description for each chapter. Chapter

7 will synthesise constraints on diamond formation beneath Botswana based on the conclusions

from all previous chapters.

In addition, I made significant contributions to two publications, in the Journal of Geochemical Exploration (Appendix A.2) and Nature Communications (Appendix A.3). The

papers are not part of the main body of work for this thesis but they include valuable aspects

related to data evaluation of inclusion major element composition and support the presented

findings on inclusion ages observed in other locations.

Chapter 1. This chapter provides a summary of the rationales and objectives of this PhD

research project and a general overview of lithospheric diamond research with a focus on

southern Africa.

Chapter 2. Examination of crystallographic growth relationships between inclusions and

the host diamond and provides insight into the genetic origin of the inclusions (first objective).

CHAPTER 1

42

Findings allow assessment of what information (e.g., age, geochemical environment, P-T)

determined from an inclusion can also be applied to the diamond host. For this study, I co-

supervised a master student, Q. van den Heuvel, on a day-to-day basis and I made major

contributions to data re-evaluation and writing the article.

Chapter 3. Reassessment of Jwaneng’s diamond populations by CL, FTIR and inclusion

composition to establish the number of discrete diamond-forming events (second objective). The

results demonstrate that detailed petrographic and spectrographic examination is necessary to

unveil the complex history recorded in diamonds.

Chapter 4. Distinct Sm-Nd isochron ages of eclogitic Jwaneng inclusions – characterised

in Chapter 3 – are used to address the multi-stage growth history of diamonds (second objective)

and different diamond-forming processes and to draw implications for the deep carbon cycle

(third objective).

Chapter 5. The evolution of the SCLM beneath Orapa, Letlhakane and Damtshaa is

revealed through Sm-Nd isochron ages of eclogitic and websteritic silicate inclusions. The data

allow insight into the local scale of diamond formation in the Orapa cluster (second objective)

and potential carbon sources are discussed (third objective).

Chapter 6. Correlation between Re-Os and Sm-Nd isochron ages of co-genetic, individual

sulphide and silicate inclusions from the Orapa cluster and Jwaneng are used to validate if mineral

inclusions in diamond record the time of diamond growth (first objective), establish the timing

and spatial scale of diamond-forming events (second objective) and provide insight into the deep

cycling of volatiles (third objective).

Chapter 7. This chapter represents a synopsis with the implications and conclusions from

the previous chapters in order to provide constraints on diamond formation beneath Botswana.

Appendices The appendices A – F include all aspects related to data evaluation and

interpretation, deployed methodology and references to the online repositories of electronic

supplemental material that were published together with the corresponding articles in Chapters 1 to 7.

Figure 9: (right) Schematic processing of a diamond and the inclusions with reference to Appendix A.1. Light grey shades delimit discrete complexes of the protocol, dark grey shades highlight the essential analytical techniques carried out in this research.

INTRODUCTION

43

INC

LUSI

ON

Wei

ghin

gM

ettle

rUM

T-7

Spik

ing

88S

r (2

0-10

0)149S

m/1

50N

d (2

0)87R

b (0

.34)

Dis

solu

tion

conc

. HF/

HN

O3

Sr c

olum

ns

TRU

-min

i col

umns

Sr

REE

mat

rix

+ R

b

Nd

Sm

min

i-LN

col

umns

mat

rixR

b

Rb

min

i col

umns

Trito

n(F

Cs

10

11!)

Mat

262

(FC

s 1

01

1!)

Trito

n(F

Cs

10

13!)

Trito

n(F

Cs

10

13!)

ICPM

S

Spik

ing

(sam

ple/

spik

e)190O

s/185R

e (3

)

Dis

tilla

tion

Cr 2

O3

in 1

2N H

2SO

4

Re

colu

mns N

i-Cu-

Co-

Fe

ICPM

S

Trito

n (S

EM

)@

Uni

vers

ity o

f Alb

erta

, E

dmon

ton

Rou

gh

DIA

MO

ND

Mor

phol

ogy

CEN

TRA

L PL

ATE

Lase

r cut

ting

@La

serE

dge

bvba

, Ant

wer

p

Polis

hing

@G

assa

nD

iam

onds

BV

, A

mst

erda

m

Ram

anm

ine

ral i

de

ntif

ica

tion

CL

imag

ing

FTIR

trave

rses/

maps

@U

trech

t Uni

vers

ity@

Uni

vers

ity o

f Bris

tol

@U

nive

rsity

of A

lber

ta,

Edm

onto

n@

GIA

, New

Yor

k

SIM

S (in

situ

)@

CR

PG

-CN

RS

, N

ancy

C-N

isot

opes

Gas

-IRM

S (e

x si

tu)

SIM

S@

CR

PG

-CN

RS

, N

an

cy

S-is

otop

es

HR

-ICPM

S@

Uni

vers

ity o

f A

lber

ta, E

dmon

ton

Nitr

ogen

con

tent

/ ag

greg

atio

n

Gro

wth

zon

atio

n

Os

mat

rix

Re

+ PG

E

C-is

otop

es

EMPA

/ SEM

@U

trech

t Uni

vers

ity

Cle

anin

g

A.1

-4.

Dia

mon

d pr

oper

ties

Sam

ple

sele

ctio

n@

DT

CB

, G

ab

oro

ne

Incl

usio

nab

unda

nce

stud

y3

70

,00

0 d

iam

on

ds

A.1

-6.

A.1

-7.

Re-

Os

age

A.1

-7.

Sm-N

d ag

e

SILI

CA

TE

Impo

rtK

imb

erle

y P

roce

ss

EBSD

(in

situ

)@

Utr

ech

t U

niv

ers

ity

FIB

(in

situ

)In

clu

sio

n c

lea

nin

g@

Utr

ech

t U

niv

ers

ity

Cry

stal

logr

aphi

cor

ient

atio

n Impu

rity

dete

ctio

n

TEM

@U

tre

cht U

niv

ers

ity

A.1

-5.

Incl

usio

n cr

ysta

llogr

aphy

FTIR

bulk

A.1-1.

Samplin

g

A.1

-3. P

roce

ssin

gB

reak

age

& re

cove

ry

A.1

-2.

Cha

ract

eris

atio

n A.1

-6.

Incl

usio

n co

mpo

sitio

n

A.1

-3.

A.1

-8.

Cle

an la

b pr

oced

ures

A.1

-6.

A.1

-6.

A.1

-6.

A.1

-8.

Cle

an la

b pr

oced

ures

SULP

HID

E


Recommended