Você está na página 1de 15

Earth-Science Reviews 136 (2014) 21–35

Contents lists available at ScienceDirect

Earth-Science Reviews
journal homepage: www.elsevier.com/locate/earscirev

New insight into polycrystalline diamond genesis from modern


nanoanalytical techniques
Dorrit E. Jacob a,⁎, Larissa Dobrzhinetskaya b, Richard Wirth c
a
Australian Research Council Centre of Excellence for Core to Crust Fluid Systems and Department of Earth and Planetary Sciences, Macquarie University, North Ryde, NSW 2109, Australia
b
Department of Earth Sciences, University of California at Riverside, Riverside, CA 92521-0412, USA
c
Helmholtz Centre Potsdam, GFZ German Research Center for Geosciences, Telegrafenberg, D-14473 Potsdam, Germany

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

Article history: Technical developments in analytical methods that reach nanometer spatial resolution have enabled the interro-
Received 23 December 2013 gation of smaller, submicron-sized inclusions in diamond that had previously been elusive. This has inspired and
Accepted 11 May 2014 enabled studies of non-classical diamond species from different geological settings, resulting in a strongly faceted
Available online 18 May 2014
and dynamic picture of diamond formation. This article reviews the leap of knowledge achieved by employing
state-of-the-art analytical methods with high spatial resolution to polycrystalline diamonds from different
Keywords:
Diamond
settings, i.e. from kimberlite, from crustal ultra-high pressure metamorphic terranes and alluvial carbonados.
Earth's mantle While crustal metamorphic diamonds are generally formed under oxidizing conditions, polycrystalline diamond
Carbonado from the Earth's mantle and carbonado have inclusion suites reflecting variable, and sometimes extreme, redox
Ultra-high pressure metamorphism conditions. Diamond fluid compositions, however, fall in the same compositional field for worldwide diamond
TEM fluids, regardless of their geodynamic environment.
Subduction On the basis of thermodynamic equilibrium data for C–H–O fluids in the mantle we argue that submicron
inclusions in diamonds are products of local remobilization connected to fluid-fluxed partial melting and
redox freezing. Thus, evidence from these inclusions complements information from classical work on larger
inclusions and allows a unique direct insight into the medium in which diamond formed.
© 2014 Published by Elsevier B.V.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2. Modern micro- and nanoanalytical techniques applied to diamond . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1. X-ray tomography and micro-X-ray computed tomography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.2. Micro-X-ray fluorescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3. Focused ion beam (FIB) milling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.4. Transmission electron microscopy (TEM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.5. NanoSIMS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3. Polycrystalline diamonds from kimberlitic sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.1. Mineral intergrowths: macroinclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.2. Oxidation state and clues to formation mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.3. Insights from nano-inclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4. Carbonado . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5. Polycrystalline diamonds from ultrahigh-pressure terranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.1. Nanoinclusions in the Erzgebirge polycrystalline diamonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.2. Nanoinclusions in the Kokchetav polycrystalline diamonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.2.1. Marbles and calcareous-silicate gneisses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.2.2. Feldspathic gneisses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.3. Diamond-forming fluid media and oxidation state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.4. Source of carbon and nitrogen aggregation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.5. Evidence from fluids for crust–mantle interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

⁎ Corresponding author. Tel.: +61 298508428; fax: +61 298508943.


E-mail addresses: dorrit.jacob@mq.edu.au (D.E. Jacob), larissa.dobrzhinetskaya@ucr.edu (L. Dobrzhinetskaya), wirth@gfz-potsdam.de (R. Wirth).

http://dx.doi.org/10.1016/j.earscirev.2014.05.005
0012-8252/© 2014 Published by Elsevier B.V.
22 D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35

6. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
6.1. Diamond fluids in the Earth's mantle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
6.2. Diamond fluids in polycrystalline diamonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
6.3. Oxidation conditions recorded by polycrystalline diamonds: a matter of scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
6.4. Chemical environments for diamond formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
7. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

1. Introduction 2–3 Ma, six of them with well-documented microdiamond occurrences


(Dobrzhinetskaya, 2012; Schertl and Sobolev, 2013) and two of these
The study of mineral inclusions in diamonds from kimberlites has contain polycrystalline diamonds. Lastly, polycrystalline diamond is
been critical in shaping our current understanding of the formation of found in alluvial deposits, most famous perhaps are the carbonado
diamond in the Earth's mantle (Navon, 1999; Stachel and Harris, diamonds from Brazil and Central Africa, whose sources are as yet
2008). However, instrumental limitations have, until recently, restricted unknown and for which a number of contrasting and mutually exclu-
studies to micrometer-sized inclusions in larger, monocrystalline sive models have been proposed (McCall, 2009).
diamonds. Smaller inclusions and inclusions in small diamonds from
kimberlitic and non-kimberlitic settings remained elusive. 2. Modern micro- and nanoanalytical techniques applied to diamond
Progress in instrumental development now enables accurate chem-
ical and structural analyses at unprecedented spatial resolution, which 2.1. X-ray tomography and micro-X-ray computed tomography
has resulted in a leap in our understanding of small-scale, even atom-
scale processes in minerals. These new microanalytical methods have X-ray tomography provides information about the occurrence and
allowed diamond inclusion studies to be carried to the next level, spatial distribution of fluid and solid-state inclusions in diamond
as submicron inclusions are now accessible, thus building on a solid (Jacob et al., 2011). Amongst the solid inclusions, mineral phases can
foundation from studies of large inclusions. This has inspired a new be identified by applying techniques such as Raman spectroscopy,
generation of innovative studies of smaller and rarer diamond species, FTIR spectroscopy, X-ray diffraction analysis (Kopylova et al., 2010)
whose inclusion content had not been accessible before. Here we review and transmission electron microscopy (TEM), particularly electron
recently developed microanalytical techniques that are not yet routinely diffraction (Wirth et al., 2009). Their chemical composition can be mea-
applied in the Geosciences and illustrate the leap in knowledge in the sured by electron microprobe analysis, μ-X-ray fluorescence analysis,
study of smaller diamonds and their inclusions that these methods TEM energy dispersive analysis and/or electron energy-loss spectroscopy
have enabled and will continue to support. We compare and contrast (EELS), and nanometer resolution secondary ion mass spectrometry
evidence from classical studies of micrometer-sized inclusion suites (Nano-SIMS) analysis. The chemical composition of fluids included in
(here referred to as “macroinclusions”) with recent results for submi- diamond can be determined by infrared spectroscopy (IR, FTIR, Weiss
cron inclusions suites (“microinclusions”) focussing on TEM investiga- et al., 2010, 2013) and synchrotron-based μ-X-ray fluorescence analysis
tions of focused ion beam (FIB) prepared samples and evaluate the (Klein-BenDavid et al., 2004; Sitepu et al., 2005).
contribution of these to our current view of diamond formation. However, if the grain size of the individual phases inside the inclu-
Polycrystalline diamonds are aggregates of diamond crystals with sions is significantly smaller than 100 nm and/or the inclusions are
heterogeneous grain sizes and random orientation. An unusual feature polycrystalline, then most of these techniques render insufficient spatial
is that these diamond aggregates can have a highly porous structure resolution. Nanometer-sized phases can be reliably determined only by
with up to 30% porosity (Heaney et al., 2005), which indicates that applying TEM techniques (electron diffraction, high-resolution imaging,
they formed from a volatile-rich medium strongly oversaturated in analytical electron microscopy (AEM)) on site-specific foils prepared by
carbon (Sunagawa, 1990). Monocrystalline diamonds can contain a focused ion beam (FIB) milling (Wirth, 2004, 2009; Dobrzhinetskaya
long history of growth and dissolution resulting in complex zonation et al., 2010). This is why we emphasize FIB/TEM techniques in the
patterns (Howell et al., 2012). In contrast, polycrystalline diamonds following section.
form rapidly (Orlov, 1977; Sunagawa, 2005), presenting snapshots of The application of various appropriate techniques mentioned above
diamond formation conditions that complement the information from provides answers to important questions, such as: How many inclusions
slowly grown monocrystalline diamond. are present in a particular volume? Where are they located in that
Polycrystalline diamonds are found worldwide in rather contrasting particular stone? What is the chemical composition of the inclusions
settings. They occur in typical crustal rocks in ultra-high pressure meta- and which phases are present? What is the size of the inclusions and
morphic (UHPM) terranes as well as in kimberlites, the main source of that of the constituting phases? How can they be made accessible for
diamonds from the Earth's mantle, and also in alluvial deposits with analysis, and by which analytical method?
unknown sources. Micro-computed X-ray tomography (μCT) is a non-destructive
Kimberlites are volatile-rich exotic magmatic rocks that contain technique that allows identification of inclusions larger than 1 μm per
diamond and mantle xenoliths originating from depths of 150– voxel (volume pixel), depending on the original sample size. CT unveils
N450 km. Polycrystalline diamond is found in Group I kimberlites in the internal structure of a sample (inclusions, porosity) down to a
Africa and Siberia where it can locally amount to ca. 20% of the total micrometer scale based on X-ray absorption, which depends on density
diamond production (K. de Corte, pers. comm. 2012). UHPM terranes variations and differences in chemical composition of the sample. A
occur along continental margins delineating a wide spectrum of series of two-dimensional X-ray absorption images are acquired and
Phanerozoic collisions of continental lithospheric plates (Ernst, 2006; then assembled into a 3D image. X-ray sources are usually synchrotron
Ernst and Liou, 2008; Dobrzhinetskaya, 2012; Schertl and Sobolev, radiation facilities or high-brilliance rotating anodes. The resolution is
2013). They are composed of metasedimentary and meta-igneous rocks basically defined by the original focal spot size and the sample size.
with continental affinities that were subjected to deep subduction up to The larger the sample the more beam broadening due to scattering
N250 km followed by exhumation. Microdiamonds have been reported effects results, thus reducing spatial resolution. Micro-computed X-ray
from almost a dozen of UHPM terranes ranging in age from 640 Ma to tomography works well as long as there is a large contrast in X-ray
D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35 23

attenuation between the host and the inclusions. To date, a number of low because of the lower atomic number of Ar ions compared with Ga
applications of X-ray computed tomography in Earth Sciences have ions used in FIB milling, and the lower acceleration voltage in argon
been published (Mees et al., 2003; Jacob et al., 2011; Cnudde et al., ion milling devices. For argon ion milling of diamond, this translates
2012; Ketcham and Koeberl, 2013 and references therein). into very long milling times, because bond strength in diamond exceeds
that of all of its inclusions and consequently will be sputtered first.
2.2. Micro-X-ray fluorescence In contrast, the FIB technique allows preparation of electron trans-
parent foils or membranes from diamond from the location of interest
Micro-X-ray fluorescence analysis is a method that excites X-ray (site-specific) leaving the inclusions untouched. However, because of
photons in a specific volume defined by the diameter of the primary the extreme hardness of diamond the sputtering yield is significantly
synchrotron beam and records the X-ray photons exited from atoms smaller than with silicates. To overcome this disadvantage a gas injec-
present in that particular volume. The penetration depth of the synchro- tion system (GIS) is usually employed that injects water vapor close to
tron beam is limited by the absorption of the incident beam by the the target surface to be sputtered thus facilitating the sputtering
target material. The absorption of the excited X-ray photons by the process.
target material defines the detection limit for individual elements. In a FIB device Ga ions are extracted from a liquid gallium source by
The use of X-ray-fluorescence techniques on micrometer-sized inclu- applying an extraction voltage (6–14 keV) to it. An appropriate suppres-
sions in minerals requires a focused X-ray beam with very high intensity. sor voltage prevents extraction of electrons from the liquid metal ion
Synchrotron radiation provides the intense source and has been used source (LMIS). Variation of the suppressor voltage adjusts the beam
for μ-X-ray fluorescence analysis of inclusions in minerals since the current to a preset value (e.g. 2.2 μA) maintaining the extractor voltage
nineteen eighties (Frantz et al., 1988). It is a non-destructive method at constant value. A constant ion current is required for reproducible
that leaves the diamond sample unaffected for further investigations. sputtering conditions and automation of the sputtering process. The
This method allows the collection of spectra with excellent detection extracted Ga ions are accelerated to 30 keV and focused onto the target,
limits for individual elements (e.g. 10 ppm for Ti and b1 ppm for Zn; which they hit with sufficient momentum to sputter atoms from the
Sitepu et al., 2005). The spectra can also be used for creating elemental target surface. Inserting appropriate apertures into the path of the ion
distribution maps. Further details of the technique are presented in one beam controls the beam diameter. The sputtering rate depends on the
of the first studies of that kind on diamond inclusions by Sitepu et al. acceleration voltage (momentum of the Ga ions), the target material
(2005), and references therein. (bond strength) and the angle of incidence (penetration depth). At
With μ-X-ray fluorescence analysis the chemical composition of an high angles such as 90° – the ion beam is normal to the target surface –
inclusion in diamond can be determined as well as its location inside the sputtering rate is low because the Ga ions penetrate deep into the
the stone. However, to identify the phase and the crystal structure of sample and interact with the sample atoms deep in the sample, but
inclusions another method such as μ-X-ray diffraction analysis can be sputtering of target atoms occurs predominantly at the surface. At
applied. Identification of micrometer-sized inclusions in diamond very low incident angles (1–2°), with the Ga ion beam approximately
using this method is described in detail in recent articles (Kopylova parallel to the foil surface, sputtering rate is high because the Ga ions
et al., 2010; Smith et al., 2011). The major advantage of this method is do not penetrate deep into the target and sputtering of surface atoms
that many solid-state inclusions can be identified simultaneously, thus occurs much more frequently. Further details of the FIB technique and
providing an indication of the frequency of individual phases in a single the FIB sample preparation of TEM foils with application of a gas
stone. Together with μ-X-ray fluorescence analysis described above it is injection system can be found elsewhere (Wirth, 2004; Giannuzzi and
possible to obtain information on the number of inclusions, their phase Stevie, 2005; Desbois et al., 2008; Wirth, 2009, 2010).
compositions and chemical compositions in a comparatively large A typical FIB prepared TEM foil has the dimensions 15 μm × 10 μm ×
volume of several hundred μm3. Micro-X-ray diffraction analysis does 0.15 μm and takes approximately 4 hours to produce in a fully automated
not require the use of synchrotron radiation. Modern rotating anode process. Significantly thinner membranes with around 35 nm thickness
X-ray generators can be considered as high-brilliance X-ray sources can be produced in a SEM–FIB combined DualBeam equipment with
suitable for this kind of diffraction analysis. in-situ lift out technique (Wirth, 2009). However, for the study of
inclusions in diamond, it is sometimes more favorable to work with
2.3. Focused ion beam (FIB) milling thicker foils to increase the chance of finding nanoinclusions that are
still intact. If the diameter of the inclusion is well below 100 nm and
While the X-ray based techniques described above are state of the total foil thickness is 200–300 nm, intact inclusions can be observed,
the art methods, they are not appropriate to study nanometer-sized while the foil is thin enough to allow electron transparency in diamond.
inclusions because of their limited spatial resolution. The leading tool In these cases, the inclusion's chemical composition can be measured
to study the “nano-world” inside inclusions is the focused ion beam with EDX and/or electron energy-loss spectroscopy (EELS) analysis.
(FIB) technique for site-specific TEM sample preparation combined Subsequently, the electron beam is focused at the inclusion directly,
with TEM. Thanks to the advent of this method in the Earth Sciences, thereby opening it by intensive electron sputtering and releasing the
we know today that many inclusions with a grain size well below liquid content into the system's vacuum. A second EDX analysis after
100 nm exist in diamond. Before FIB sample preparation was available opening the inclusion can demonstrate which elements were present
to prepare electron transparent foils from diamond (starting from in a solid crystalline or quench phase, which were dissolved in the
approximately 2000), there were only two options to produce speci- fluid or gas phase and, by mass balance, which escaped into the vacuum.
mens applicable for TEM studies. FIB-prepared TEM foils have several advantages over other specimen
One of these is crushing of the sample and placing the crushed preparation techniques. (1) FIB prepared foils can be prepared at
material onto a carbon covered TEM grid. This method has many disad- uniform thickness. This allows the acquisition of elemental maps of
vantages, particularly loss of spatial information and inhomogeneous the inclusions because the recorded X-ray intensities depend only on
thickness, the latter excluding application of line scans and element the concentration of the individual elements and not on sample thick-
mapping. The second method is the argon-ion milling technique that ness. Line scan analyses are also possible but are restricted to inclusions
was usually applied in the past to prepare electron transparent foils. that are not sensitive to electron irradiation damage and electron
However, this method has the tremendous drawback that due to the sputtering (sulfides, garnet, olivine). (2) The location of the foil in the
incident angle of the ions (a few degrees) preferential thinning occurs sample can be easily reconstructed because the sputtering site is visible
with the result that the weakest material is always sputtered first. on the sample surface under an optical microscope at moderate magni-
Momentum transfer from Ar ions onto the target atoms is comparatively fication. (3) No preferential sputtering of the inclusions occurs. (4) No
24 D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35

further carbon coating of the foil is required, which is essential when method are given elsewhere (Stadermann et al., 2005; Floss et al.,
investigating carbon-based materials. 2006; Hoppe et al., 2013 and references therein).
Because of the limited sputtering depth (10–15 μm) inclusions Many recent applications of these methods focused on polycrystal-
must be located close to the surface to be accessible by FIB otherwise line diamonds from such contrasting settings as UHP metamorphic
mechanical polishing is required. terranes, alluvial sources and kimberlites. Metamorphic diamonds
from crustal UHP rock are typically in the micron range, thus an ideal
2.4. Transmission electron microscopy (TEM) target for these methods. Carbonados and polycrystalline diamonds
from kimberlites are porous and contain minerals intergrown with the
FIB-prepared foils or membranes are studied with TEM methods. diamond grains in the pores. These minerals have been studied exten-
TEM is an ideal tool to investigate nanometer-sized inclusions in sively to enlighten the origin of their hosts, but because of the intercon-
diamond or any other minerals (Lee, 2010), because it allows the acqui- nectivity of the pore space, the intergrown minerals are not shielded
sition of chemical, microstructural and structure information from the against alteration and the evidence they bring has to be critically
same object even at nanometer-scale. The chemical information is interpreted. Studying submicron minerals and fluids actually included
derived from energy dispersive spectra (EDS) of collected X-ray in the diamonds delivers unambiguous evidence and has proved very
photons escaping from the sample. Another more qualitative method beneficial, as reviewed in detail below.
is the acquisition of Z-contrast sensitive images with a high-angle
annular dark field (HAADF) detector in the scanning transmission
3. Polycrystalline diamonds from kimberlitic sources
(STEM) mode. The HAADF images display differences in chemical com-
position in an inclusion composed of different phases. In some special
Polycrystalline diamond (PCD) is classified by grain size as framesite
cases EELS can be used for qualitative chemical analysis. For example,
(finer; Gurney et al., 1984) or bort (coarser; Orlov, 1977) and typically
if iron is present in an inclusion there is an overlap of the Fe-L peak in
has a porous structure. Gray to silver in color, these diamond aggregates
the EDX spectrum with the F-K peak. If we want to prove the presence
are often intimately intergrown with silicates, oxides and/or sulfides.
of fluorine in a particular phase, the fluorine F-L edge in the EEL spec-
Diamond crystals are randomly oriented and range over at least five
trum can be used because there is no interference with the iron Fe-L
orders of magnitude in size (Sobolev et al., 2009). PCD is also referred
edges. The chemical composition of Fe-carbides can be measured with
to as “diamondite”, which emphasizes the perspective that these
EEL spectroscopy by calculating the Fe/C ratio (Kaminsky and Wirth,
aggregates are rocks, however small, with diamond as the rock-forming
2011).
mineral while silicates and other phases are minor and accessory
In addition to the chemical composition, grain size, grain shape
components.
and defect structure (dislocations, stacking faults, nitrogen platelets)
Polycrystalline diamond is a regular member of the diamond suite in
of the inclusion are of interest. This information can be obtained by
a number of kimberlite pipes worldwide, where it is believed to amount
bright-field and dark-field imaging. Usually, the crystal structure, which
to about 20% of the total production (K. de Corte, pers. comm., 2012).
unambiguously identifies a solid-state inclusion, is derived from
Kimberlite pipes in Africa (Venetia, Premier, Jwaneng, Orapa) yield
electron diffraction patterns or convergent beam electron diffraction
polycrystalline diamonds, as do some diamond mines in Siberia (Mir,
(CBED) patterns. However, for selected area electron diffraction the
Aikhal; Sobolev, 1977).
nanometer-sized crystals are too small and CBED usually cannot be
This variety of diamond is only lately beginning to receive increased
applied because the focused electron beam immediately destroys
attention and sheds light on another facet of diamond formation in the
the crystal structure rendering the inclusion amorphous. The best
Earth's mantle, namely formation by rapid crystallization from a fluid
applicable approach in this case to derive structural information is
supersaturated in carbon (Orlov, 1977; Gurney and Boyd, 1982; Jacob
high-resolution lattice fringe imaging. Due to very short exposure
et al., 2000, 2011). Much of what we know about them comes from
times necessary to acquire images with lattice fringes with a CCD
study of phases intergrown with, rather than actually being included
camera (b 1 s) the crystal will not be destroyed before the image is
in, the diamond crystals. These phases range from several tens to
acquired. After acquisition the lattice fringe image is Fourier trans-
hundreds of micrometers in size and are termed here “macroinclusions”
formed (FFT), calculating a diffraction pattern. From the diffraction
to set them apart from the submicron-sized phases included in the
pattern in reciprocal space it is possible to measure the length of several
diamond crystals that have only been accessible in recent times.
diffraction vectors and angles between them. Transformation of the
length of the vectors into real space (d-spacing) allows a comparison
of the measured d-spacing with calculated d-spacing of the phase we 3.1. Mineral intergrowths: macroinclusions
expect to be present from chemical data. A comparison of the measured
d-spacing with the calculated d-spacing and a comparison of the angles High-pressure minerals intergrown with the diamond aggregates
between adjacent vectors allow an identification of the phase. Thus, are commonly interpreted as syngenetic to the diamonds, based on
TEM methods enable the identification of nanometer-sized phases the fact that they include small diamond crystals (Kurat and Dobosi,
inside inclusions in diamond and minerals in general. 2000) or display chemical correlations with the diamonds (Jacob et al.,
2000, 2004). However, it is noteworthy that unlike inclusions in
2.5. NanoSIMS diamond these intergrown phases are not shielded from later metaso-
matic overprint.
In addition, FIB-prepared TEM foils can also be used for NanoSIMS The aggregates contain mostly silicates and oxides of eclogitic,
studies, which can deliver isotopic information at high spatial resolu- websteritic and peridotitic affinity, broadly similar to the macro-
tion. This has been demonstrated with carbonate inclusions in diamond inclusion suite found in monocrystalline diamonds, although character-
(Wirth et al., 2007), nitride inclusions in coesite (Dobrzhinetskaya et al., istic differences exist. Chromium-poor (non-peridotitic) garnets in
2010) and determinations of the isotopic composition of oxygen and polycrystalline diamonds have a more restricted compositional range
carbon in lower mantle carbonate and silicate inclusions in diamond than inclusions in monocrystalline diamonds, containing typically less
(Kaminsky et al., 2012). NanoSIMS uses an ion beam (Cs+ or O− than 6 wt.% CaO (Fig. 1). Thus, most classify as websteritic (Aulbach
depending on the target material to be sputtered) usually b100 nm in et al., 2002), and don't reflect the large variation in grossular component
diameter that is scanned over a preselected area e.g. 4 × 4 μm. The seen in eclogitic garnets from mantle xenoliths. Nevertheless, some
secondary ions sputtered from the target material are collected and phases can clearly be classified as eclogitic, for example omphacitic
guided into a mass spectrometer for isotopic analysis. Details of this clinopyroxene and rutile in an Orapa sample (Jacob et al., 2011).
D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35 25

Fig. 1. Major element compositions of garnets intergrown with diamond in polycrystalline


diamonds from kimberlites compared to garnets from eclogitic xenoliths (Roberts Victor
Mine, RSA, large gray field; Jacob, 2004). Gray field labeled A = peridotitic–websteritic
garnets; dark gray field B = websteritic–eclogitic compositions. Note the more restricted
grossular component compared to typical eclogitic xenoliths.
Data from Dobosi and Kurat (2002, 2010), Jacob et al. (2011) and Kurat and Dobosi (2000).

While restricted in their grossular contents, the garnets show a large


variation in almandine contents (13–46%), covering nearly the whole
variation reported for eclogite xenoliths (Fig. 1). In this respect, their
major element composition is similar to low Cr garnet megacrysts
found in kimberlites (Gurney et al., 1979). Magnetite, while rarely
reported from monocrystalline diamonds (Harris, 1968; Sobolev et al.,
1989) coexists with almandine-rich garnet in a sample from Orapa
(Jacob et al., 2011) and many other polycrystalline diamond aggregates
are magnetic suggesting the presence of magnetite or even native iron
(Jacob et al., 2004). Thus, compared to macroinclusions in diamonds,
magnetite is overrepresented in polycrystalline diamond aggregates.

3.2. Oxidation state and clues to formation mechanisms


Fig. 2. Carbon isotopic compositions of polycrystalline diamonds from kimberlites (a),
from UHPM terranes ((b): white = Kokchetav, gray = Erzgebirge) and of carbonado
Phases intergrown with polycrystalline diamonds show a similarly
diamonds (c) compared to the typical range of monocrystalline diamonds with peridotitic
large range of redox conditions to macroinclusions in monocrystalline inclusions (light gray band) and fibrous diamonds (dark gray band). Data for polycrystalline
diamonds of over ca. 10 log units fO2. Inclusions consisting of native diamonds from Jacob et al. (2000), Maruoka et al. (2004), and Mikhail et al. (2013), for
iron and iron carbide, for example, provide evidence for reducing condi- UHPM diamonds from Cartigny et al. (2001), Imamura et al. (2013), and Dobrzhinetskaya
tions well below the iron–wüstite oxygen buffer (Jacob et al., 2004). The et al. (2010), for carbonado from Cartigny (2010) and for fibrous diamonds Cartigny
(2005) and references therein.
majority of polycrystalline diamonds, however, are compatible with
more oxidizing conditions involving carbonates.
The origin of the carbon-bearing fluid is debated. Some authors have The only study to date on radiogenic isotope systems of minerals
argued for an entirely mantle-derived carbon fluid, which subsequently included in polycrystalline diamonds reported unradiogenic Nd isotopic
undergoes Rayleigh fractionation (Deines, 1980; Maruoka et al., 2004), ratios of −15.9 to − 21.7 εNd, typical for ancient lithospheric material
while others infer a subducted carbon source (Mikhail et al., 2013). and argued for recent remobilization of this material to form polycrys-
The δ13C values (n = 115) overlap with those for eclogitic and meta- talline diamonds (Jacob et al., 2000). The authors based their arguments
morphic diamonds and show a broad bimodal distribution from −1.3 on coupled chemical characteristics of diamonds and intergrown
to − 29‰ with major frequencies of occurrence at − 5‰ and at garnets as well as on trace element zonation in the garnets, which
− 19‰ (Fig. 2; Table 1). The first peak at the typical mantle value of would have been erased by diffusion during expanded storage at mantle
− 5‰ (Cartigny, 2005) argues for at least partial involvement of temperatures and pressures. Contrasting with this apparently recent
mantle-derived carbon. The second peak at −19‰, however, is distinct formation of polycrystalline diamonds at the Venetia Mine are findings
and typical only for polycrystalline diamond aggregates (Cartigny, that some samples from an unknown South African locality show high-
2005). Whether this mean value is indicative of the formation processes temperature deformation and annealing structures in Electron
or source of carbon or if it is connected to biased sampling of the small Backscattered Diffraction (EBSD) that apparently require long mantle
database remains to be seen. residence times (Rubanova et al., 2012).
Nitrogen concentrations and nitrogen isotope ratios in these In summary, polycrystalline diamonds form episodically by small-
diamonds cover a large range of 8 to 3635 ppm and δ15N − 6.1 to scale, rapidly occurring fluid-dominated redox processes within the
+ 22.6‰ (Gautheron et al., 2005; Mikhail et al., 2013), but neither subcratonic lithosphere. Intergrown micron-sized phases are most
nitrogen concentrations nor δ15N values are coupled with δ13C values, often hybrids, carrying evidence for a mixed derivation involving fluids
apparently discounting coupled fractionation processes operating and peridotite or eclogite solids and preserving inhomogeneity as a
during diamond growth (Deines, 1980; Maruoka et al., 2004; Cartigny, result of their rapid formation. Their trace element patterns often
2005). Nitrogen aggregation states are across the whole spectrum show strong enrichment in elements indicative for carbonatite metaso-
from pure IaA to pure IaB (Mikhail et al., in press). matism, thus arguing for a role for mainly oxidized fluids. However, very
26
Table 1
Summary of some major characteristics of diamonds discussed in this study compared to fibrous and monocrystalline diamonds.

Geological setting Macro-inclusions Micro-inclusions Carbon isotopes Nitrogen characteristics P and T fO2 range of
(5–200 μm) (b1 μm) (δ13C values) inclusions

Polycrystalline Subcratonic lithosphere Intergrowths mostly websteritic, Silicates, oxides, sulfides, −2 to −28‰, bimodal with means 0–4000 ppm, δ15N = −5.7 to +22.5‰ Diamond stability Mostly FMQ to IW
aggregates (Group I kimberlites) but also eclogitic and peridotitic metals, fluids. Carbonate at −5‰ and −18‰ (see Fig. 2 for (Mikhail et al., 2013) IaA–IaB (Mikhail et al., field buffers, some be-
suites present but rare references) in press) yond
UHPM terranes Continental collisions None (due to the small size of the Silicates, carbonates, halides, −10 to −27‰ 0–11, 150 ppm, δ15N = −1.8 to +14‰ 6–9 GPa, 900– Around FMQ/CCO
diamonds) metals, fluids (Cartigny et al., 2001) Ib–IaA (De Corte et al., 1100°C buffers
1999)
Carbonado Alluvial, inferred to be Rare, mostly intergrown phases Silicates, carbonates, halides, −5 to −30‰, most −22 to −30‰ 100–300 ppm (Cartigny, 2010) δ15N = −17 Unknown Mostly FMQ to IW
from the Earth's mantle metals, fluids (Cartigny, 2010) to +8‰ (Heaney et al., 2005), Ib–IaA
(Cartigny, 2010)
Fibrous Subcratonic lithosphere None Fluids, silicates, carbonates, −4 to −8‰ (Boyd et al., 1994) 0–3000 ppm, δ15N = −2 to −9‰, IaA, some Diamond stability Around FMQ
halides. All K-bearing Ib–IaA (Cartigny, 2010) field

D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35


Monocrystalline Mostly subcratonic Peridotitic, eclogitic, websteritic Unstudied +1 to −40‰, mean for peridotitic 0–3000 ppm, δ15N = +5 to −10‰ 4.5–7.5 GPa, 950– FMQ to IW but
lithosphere, some suites at −5‰ and wide distribution for (Cartigny, 2005), all aggregation states, but 1350°°C (and mostly around
asthenospheric eclogitic (Stachel et al., 2009) mostly IaA–IaB (Shirey et al., 2013) above) FMQ
version). Scale bars 200 nm.
inclusion in a polycrystalline diamond (see Jacob et al., 2011 for an online animated
magnetite included in polycrystalline diamond with porous structure, (b) intact fluid
Fig. 3. TEM HAADF images of inclusions in polycrystalline diamond grains. (a) shows a

were not observed in the magnetites, this may be due to the fact that
the diamond-forming fluid. Although cavities still containing fluid
which was interpreted as evidence for formation in the presence of
displayed a distinctive polycrystalline and porous texture (Fig. 3a),
mond. Magnetite, volumetrically the most prominent inclusion phase,
inclusions, thus confirming that this paragenesis is syngenetic to dia-
pyrrhotite, omphacite and rutile were found as both nano- and micro-
crystals, the latter were considerably altered at a late stage. Magnetite,
identical to the micro-sized phases intergrown with the diamond
although major nano-phases included in the diamond were largely

tion of the diamonds.


overprinted or even formed later and are thus unrelated to the forma-
shielded by diamond to evaluate whether the former were chemically
micrometer-sized intergrowth phases could be compared to those
inclusions and, most importantly, of coexisting fluids. Furthermore,
TEM enabled a more complete and direct characterization of nano-

3.3. Insights from nano-inclusions

possible scenario.
pronounced heterogeneity in their source regions upon formation as a
reducing inclusions can also be found in these rocks and underline
A sample from the Orapa kimberlite (Jacob et al., 2011) showed that

Taking these earlier studies to the nano-scale with the aid of FIB–
D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35 27

a typical TEM foil represents only a very small volume of the total crystal that at typical upper mantle pressures and temperatures the C–H–O
or an approximate 2D section. medium from which these diamonds formed is subcritical, being either
Similarly, several open cavities were encountered in the diamonds, a carbonatitic melt under oxidizing conditions (Dasgupta et al., 2004) or
which were interpreted to have lost their fluid content upon FIB sec- a CH4-rich fluid at low oxygen fugacity (Stalder et al., 2001). A halide
tioning. These cavities contained rutile and/or pyrrhotite and omphacite component that represents one of the end-members identified in
and often non-stoichiometric phases that were interpreted as quench diamond fluids worldwide (Klein-BenDavid et al., 2004) has not yet
phases. One occurrence of an intact inclusion enabled identification of been encountered.
a fluid by its characteristic continuous changes in diffraction contrast
due to density fluctuations caused by the electron beam (Fig. 3b). This 4. Carbonado
fluid was associated with fine-grained pyrrhotite, a silicate phase rich
in Fe, P, Mg, Al, Ca and K and a quench phase rich in non-stoichiometric Carbonado is a smooth-surfaced fine-crystalline (1–10 μm) and
amounts of Fe, P and Si. Amongst all inclusions only one ca. 10 nm highly porous diamond aggregate, mostly black in color, which lacks
large inclusion was found whose chemical composition suggested it to most of the typical macroinclusions found in monocrystalline diamonds
be a carbonate, although phase identification was not possible. The (Heaney et al., 2005; McCall, 2009). Their origin is still under debate
fluid found in the polycrystalline diamonds from Orapa is thus charac- with models ranging between crustal derivation, genesis in the
terized by silicate rich in Fe, Mg, Al, Ca, K and P, while carbonate is upper Earth's mantle and an extraterrestrial origin (see Heaney et al.,
rare and halide phases are absent (Fig. 4). 2005; McCall, 2009; Cartigny, 2010; Haggerty, 2014 for detailed
The majority of polycrystalline diamonds bear evidence for having summaries).
precipitated from an oxidized C–H–O medium, while only a minor pro- Characteristic features of carbonado are inclusions of crustal
portion contain very reduced inclusions. Experimental studies suggest minerals, such as graphite, quartz, rutile, florencite (a water-bearing
aluminium phosphate), and clay minerals that occur along partially
open grain boundaries and in pores. Carbonado (sensu stricto) is found
in alluvial deposits in Brazil and in Central Africa, and its primary source
is unknown; there is a high chance that some of the inclusions in carbo-
nado are in fact epigenetic, and are thus unable to provide any insights
into the origin of this enigmatic diamond (Heaney et al., 2005). Modern
submicron methods now provide the means to test this by comparing
the nanoinclusion suite in the carbonado with the chemistry of the
macrocinclusions.
Carbonado is similar to polycrystalline diamond aggregates from
kimberlites and the latter are sometimes mistakenly classified as carbo-
nados. Both have a porous structure and random crystallographic orien-
tation of the individual diamond grains within their polycrystalline
aggregates (Ishibashi et al., 2012; Rubanova et al., 2012; Kaminsky
et al., 2013). Furthermore, some polycrystalline diamond aggregates
can show equally small grain sizes and the microporphyritic textures
that are typical of carbonado (see Heaney et al., 2005; Haggerty, 2014
for reviews), and carbonado in turn can display coarse-grained
microtexture (e.g. Trueb and Butterman, 1969). However, carbonado
has much higher porosity (Vicenzi et al., 2003) and displays a very dis-
tinct and unusual nanostructure recognized with TEM, displaying
column-like diamond grains that form linear and zigzag patterns with
low-angle and even partially open grain boundaries. A characteristic
feature of individual diamond grains in carbonado is the presence
of uneven and rough grain boundary planes as displayed in Fig. 4
(Rondeau et al., 2008), very unlike the microstructure of polycrystalline
diamonds from kimberlites. In addition, carbonado occurs in remark-
ably large sizes, much bigger than those reported for polycrystalline
diamond aggregates from kimberlites. Svisero (1995) pointed out that
ten of the eleven largest diamonds reported from Brazil are carbonados,
the largest weighing 3167 carats.
Isotopic dating using Pb isotopes of bulk diamonds (Ozima and
Tatsumoto, 1997) and Pb–Pb SHRIMP dating of quartz and rutile inclu-
sions (Sano et al., 2002) of both Brazilian and Central African carbona-
dos yielded Archean ages, which served as a basis to hypothesize that
both populations are derived from the same eroded source (McCall,
2009). Carbon isotopes define a distinct range of δ13C = − 23.5 to
−33‰ (Table 1; but note two outliers at higher, typical mantle values,
Fig. 2) which led a number of authors to argue for a crustal origin involv-
ing organic carbon (Smith and Dawson, 1985) or irradiated carburanium
Fig. 4. Demonstrates the characteristic microstructure of carbonado inside a diamond (Kaminsky, 1987, 1991; Ozima et al., 1991). Carbonados are rich in
grain from the Central African Republic (CAR2#1414, Kaminsky et al., 2013). (a) TEM hydrogen (Garai et al., 2006), contain single nitrogen impurities (type
bright field image of several diamond grains with irregular grain boundaries. Some grains Ib–IaA; Garai, 2012) and polycyclic aromatic hydrocarbons (Garai
show straight dislocation lines. (b) and (c) TEM dark field images taken at the 111 reflex
and displaying three individual diamond grains (copy of a film negative, therefore, 111
et al., 2006), which are otherwise rarely encountered in terrestrial
crystal faces are dark in the image). Note the irregular grain boundaries and the locally diamond, but do occur in presolar diamond grains from meteorites.
changing inclination of the grain boundary planes. These characteristics and the apparent lack of primary terrestrial silicate
28 D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35

inclusions led to speculations of an extraterrestrial origin of these


enigmatic diamonds (Haggerty, 1999, 2014).
FIB–TEM analyses provided the means to search for smaller silicate
or oxide inclusions inside individual diamond crystals in order to
exclude any potentially epigenetic material (De Souza Martins, 2006;
Sautter et al., 2011; Kaminsky et al., 2013). The inclusions found with
this method were generally relatively small (in the hundred nanometer
range). Phases comprised almandine–pyrope garnet, augite, apatite
(including fluorapatite), phlogopite, SiO2, Ca–Mg–Sr–and Ca–Ba-
carbonates, halides (sylvite and bismocolite BiOCl), native Ni and
metal alloys (Fe–Ni, Cr–Fe–Mn, and Pb–As–Mo), oxides (FeO, Fe–Sn–
O, TiO2, SnO2, FeTiO3 and PbO2), and Fe–sulfides, as well as fluid inclu-
sions. These results are similar to an earlier TEM study of syngenetic
inclusions in carbonado by De et al. (1998) who reported native metals
and alloys in addition to SiC, sylvite, calcite, smithsonite (ZnCO3) and
Mg–Al silicates. The fluid inclusions are predominantly silicate–carbonate
compositions; halide-rich fluids were also observed, but more rarely.
Molecular H2O, detected by FTIR, was found in a carbonado from Central
Africa (Ishibashi et al., 2012).
The nanoinclusion suites of carbonados thus show much more
similarity to findings in mantle-derived diamonds (Klein-BenDavid
et al., 2007; Cartigny, 2010; Jacob et al., 2011) and metamorphic
diamonds from ultrahigh pressure terranes (e.g. Dobrzhinetskaya,
2012; Dobrzhinetskaya et al., 2013) than the larger inclusions. In con-
trast, low-pressure crustal phases identified in earlier studies, such as
florencite, were not encountered, pointing to an epigenetic origin of
these minerals. Therefore, the solid phase inclusions reported from
carbonados so far (while maybe not yet being representative) do not
support the extraterrestrial origin suggested by Haggerty (1999, 2014)
and Garai et al. (2006).

5. Polycrystalline diamonds from ultrahigh-pressure terranes

Polycrystalline diamonds (Fig. 5) were identified recently within


felsic quartz–feldspathic gneisses of the Erzgebirge massif, Germany
(Dobrzhinetskaya et al., 2013) where they occur in addition to abundant
monocrystalline microdiamonds. Massonne (2003, 2006) showed that
these rocks were subjected to UHP metamorphism during continent–
continent collision at T = 1100 °C and P = 7–8 GPa, which corresponds Fig. 5. Typical diamonds from ultrahigh pressure metamorphic terranes: (A) diamond
to depths of ~200–210 km. inclusion in zircon from the Erzgebirge massif, Germany (modified from Dobrzhinetskaya
et al., 2007), (B) diamond from the garnet–biotite gneisses of Kokchetav massif,
While these represent one of the most recent finds, polycrystalline
Kazakhstan (collection Dobrzhinetskaya).
diamonds have been known for more than a decade from rocks of the
Kokchetav massif as so-called “S- and R-type” diamonds (Ishida et al.,
2003; Ogasawara and Aoki, 2005) or “overgrown” diamonds (Sitnikova KAlSi3O8 that are associated with perforated cavities. The amorphous
and Shatsky, 2009). These diamond-bearing metasedimentary rocks of matter is assumed to be the residue of evaporated remnants of former
the Kokchetav UHP massif are thought to have formed under similar fluid inclusions that were perforated by either the FIB sputtering during
conditions as those in the Erzgebirge, namely in a deep subduction foil preparation or the TEM electron-beam sputtering upon focussing
zone at T = 980–1200 °C and P = 6–9 GPa, corresponding to depths the beam to form a very small probe of only several nanometers in
of ca. 160–250 km (Dobrzhinetskaya et al., 2005; Ogasawara and Aoki, diameter. The process of fluid bubble penetration by electron sputtering
2005). has been observed several times during numerous TEM study sessions
and detailed descriptions of this effect can be found elsewhere
5.1. Nanoinclusions in the Erzgebirge polycrystalline diamonds (e.g. Dobrzhinetskaya et al., 2005). EDX analysis combined with TEM
observations showed that the suite of amorphous inclusions contains
The diamonds included in felsic gneisses contain abundant traces of Cl, S, P, Al, K, Fe, As, Hg, Mo, Co and Ti mixed in different
nanoinclusions of 10–200 nm in size. These can be differentiated into proportions and combinations (Dobrzhinetskaya et al., 2013).
two suites, based on their occurrence within the diamonds: along the The histogram in Fig. 6a comprises the combined compositional
grain boundaries of microcrystalline diamond aggregates (suite 1), or data for all suites of fluid inclusions in the Erzgebirge gneisses (for
on the dislocations that are widely developed inside diamond crystals details on the data reduction see Dobrzhinetskaya et al., 2013). Data
(suite 2). Suite 1 includes crystalline and amorphous nanoinclusions of analysis suggests that the integral composition of the Erzgebirge
SiO2, rare CaCO3, BaCO3 and KAlSi3O8 as well as abundant amorphous diamond-forming fluids correspond to a high-density C–O–H fluid, a
quench phases of an approximate composition of NaSO4 and KCl. All hydrous-silicic fluid rich in Al and K and a hydrous-saline fluid rich in
phases contain traces of Pb, K, Ca, Cr, Fe, Al, P, Cl, S, Ti, and Zr in different Cl, K and Na. The overall composition of these fluids is similar to that
combinations. Structures of crystalline phases of SiO2, CaCO3 and KAlSiO8 of fluids in fibrous diamonds plotted in Fig. 9. Traces of carbonate,
could not be determined due to their very small size (~10 nm). Suite 2 is sulfate, phosphate and metallic cations such as Fe, Mg, Ti, Cr, Co, Mo,
composed of rare crystalline nanoinclusions of ZrSiO4, NaSO4, BaSO4 Zn, Zr, Pb, As and Hg reflect the local geochemical diversity of the
and countless b 1–10 nm amorphous inclusions, mostly SiO2 and metasedimentary rocks which host these diamonds.
D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35 29

The nanoinclusions are situated in cavities along diamond–diamond


interfaces or even create films of amorphous matter between individual
grains of the polycrystalline aggregates, but they are also observed inside
individual diamond crystals. Most of these crystalline nanoinclusions
are oxides (SiO2, TiO2, FexOy, Cr2O3, ThxOy), or silicates, e.g. KAlSi3O8,
and rare inclusions of ZrSiO4, BaSO4, and CaCO3–aragonite occur as
well (Fig. 6b).
The inclusions contain variable amounts of K, Na, P, S, Pb, Zn, Nb, Al,
Co, Mo, P, and Cl (e.g. de Corte et al., 1999; Dobrzhinetskaya et al., 2005).
Amorphous nanoinclusions associated with cavities have similarly
diverse compositions and exhibit large variations of trace elements.
They are considered to be non-stoichiometric “quench” products of
fluid inclusions that were opened during preparation or analysis.
Diamonds hosted by felsic rocks have abundant Si- and Al-bearing
inclusions and contain rare carbonate inclusions, whereas diamonds
from marble and calcareous-silicate gneisses are mostly rich in carbon-
ate inclusions and only rarely contain Si-, or Al-bearing minerals. This
compositional diversity of the crystalline and amorphous nanoinclusions
shown by the Kokchetav polycrystalline diamonds in all rock types
clearly indicates that in general the fluid from which diamonds crystal-
lized has a local origin in UHP terranes.

5.3. Diamond-forming fluid media and oxidation state

The concept of UHPM diamond crystallization from a supercritical


C–O–H fluid in the diamond stability field is agreed upon by a majority
of researchers (see review papers: Dobrzhinetskaya, 2012; Schertl and
Sobolev, 2013). This has received considerable support from detailed
analytical studies of nanoinclusions in natural UHPM diamonds and
from experiments, which successfully synthesized diamond from differ-
ent carbon species in the presence of H2O at pressures and temperatures
corresponding to the diamond stability field (e.g. Dobrzhinetskaya et al.,
2009).
The diamond-hosting sedimentary rocks are rich in H2O and
diamonds themselves contain nanoinclusions that are enriched with
almost all elements available in the host rocks. These observations
Fig. 6. Histograms showing average (n) elemental composition of fluids in diamonds
show that the origin of the diamond-forming fluid is local.
measured as raw counts per second TEM-EDX data: (a) garnet–quartz–feldspathic
gneisses of the Erzgebirge massif, Germany (n = 40 inclusions), (b) garnet–biotite
The presence of carbonate inclusions in polycrystalline diamonds
gneisses of the Kokchetav massif, Kazakhstan (n = 50 inclusions), (c) marbles of the and in single diamond crystals from both the Erzgebirge and the
Kokchetav massif, Kazakhstan (n = 45 inclusions). Kokchetav UHPM terranes showed that the oxygen fugacity must have
been close to the CCO and the FMQ oxygen buffers (logfO2 = − 6 in
5.2. Nanoinclusions in the Kokchetav polycrystalline diamonds Fig. 7; e.g. Dobrzhinetskaya, 2012; Dobrzhinetskaya et al., 2013).

In the UHP Kokchetav massif, polycrystalline diamonds are found 5.4. Source of carbon and nitrogen aggregation
as inclusions in diopside and phlogopite from metacarbonate rocks
(marbles and calcareous rocks) and in feldspathic gneisses. Polycrystalline and single crystal diamonds from UHPM terranes
originate from the same fluid source (e.g., Dobrzhinetskaya, 2012;
Dobrzhinetskaya et al., 2013). The diamonds from the Kokchetav massif,
5.2.1. Marbles and calcareous-silicate gneisses
Kazakhstan have δ13C = −10.57‰ and N contents of up to 11,385 ppm
Diamonds from these carbonaceous rocks contain abundant crystal-
(Cartigny et al., 2001), whereas diamonds from dolomitic marble are
line and amorphous nanoinclusions, and their chemical composition
characterized by δ13C = − 10.19‰ and N = 2,762 ppm (Table 1,
clearly indicates that the source for the diamonds is locally derived.
Fig. 8). These δ13C values are interpreted to be a mixture between crustal
They have distinct morphologies and are known as “S- and R”-type
and mantle carbon reservoirs (Cartigny et al., 2001; Ogasawara and
diamonds (Ishida et al., 2003). The nanoinclusions are aragonite, mag-
Aoki, 2005), which is supported by He and Ne isotope studies in these
nesite and rarely also TiO2. No SiO2 inclusions nor any Al–Si-bearing
diamonds (Sumino et al., 2011).
minerals nor their quench products have been observed. The crystalline
The Erzgebirge microdiamonds (Fig. 8) are characterized by
nanophases and the amorphous matter lining the walls of small cavities
significantly lighter carbon in comparison with the Kokchetav terrane,
(remnants of decrepitated fluid inclusions) contain a diverse spectrum
ranging in δ13C between − 17 to − 27‰ and N = 100–4,647 ppm
of lithophile and siderophile trace elements such as Co, Cu, Zn, Mn, Pb,
(Dobrzhinetskaya et al., 2010). This carbon isotopic variation suggests
K, Ba, Fe, Sr, Th, Zn, Cl, S and P (Fig. 6c).
that the Erzgebirge diamonds, in contrast to the Kokchetav suite, crys-
tallized from an entirely crustal carbon reservoir, although whether
5.2.2. Feldspathic gneisses that carbon was organic or inorganic may be debated. The wide varia-
Monocrystalline and polycrystalline diamonds from feldspathic tion of nitrogen abundances was explained by an inhomogeneous
gneisses of the Kokchetav massif are similar to those from felsic gneisses distribution of N-bearing fluid inclusions, or by the presence of N-
of the Erzgebirge massif in that they also contain two suites of bearing mica such as phengite, which is observed in association
nanoinclusions of which one is crystalline while the other is amorphous. with microdiamonds or as inclusions (e.g. Stöckhert et al., 2001;
30 D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35

residence times in the mantle (e.g. Jones et al., 1992; Cartigny et al.,
2001).

5.5. Evidence from fluids for crust–mantle interaction

Though most observations have pointed to a crustal origin there are


several lines of evidence that the diamond-bearing fluid in at least some
UHPM terranes is a product of crust–mantle interaction. One of them is
the presence of eskolaite (Cr2O3) inclusions that contain traces of Al, Si, P,
Ni, and Fe in Kokchetav diamonds from felsic gneisses (Dobrzhinetskaya
et al., 2003). Dobrzhinetskaya and Wirth (2009) postulated that Ni- and
Fe-contents as well as the Cr-content of eskolaite were probably derived
from a mantle wedge reservoir. Noble gas studies (Sumino et al., 2011)
supported this proposition. These authors found that the Kokchetav
diamonds contain primordial He and Ne, suggesting that a deep mantle
plume affected the subducted continental slab. Though no evidence
of crust–mantle interaction has yet been found in the Erzgebirge
diamonds (potentially mostly because of the lack of He–Ne isotopes
Fig. 7. Composition and oxidation state of C–O–H fluids in equilibrium with diamond for studies), the mechanism of diamond formation from both UHPM
P = 5 GPa, T = 1400 K (modified after Taylor, 1990; Odling, 1989). The heavy curve
labeled “diamond saturation line” delineates the stability field of diamond and coexisting
terranes is characterized by active “crustal” fluid migration through
fluid at this P and T. The oxidized fluid-only region lies above, and the reduced fluid-only the subduction zones. In Kokchetav, the process was complicated by
region is to the lower left of the saturation curve. They are separated by a diamond + fluid the interaction of this local, crust-derived fluid with mantle fluids.
field that extends to very water-rich fluid compositions labeled “water maximum”.
Positions of the magnetite–wüstite (MW) and the iron–wüstite (IW) oxygen buffers are
6. Discussion
indicated by dashed lines. Reducing fluids coexisting with diamond at fO2 = IW and
below consist of mixtures of CH4 N H2O N H2 and C2H6 progressing to mixtures of CH4 N
C2H6 N H2 at fO2 well below IW. 6.1. Diamond fluids in the Earth's mantle

While mantle and crustal environments for the formation of poly-


Dobrzhinetskaya et al., 2010). Kinetic effects upon diamond crystalliza- crystalline diamonds appear strikingly different, all of these diamond
tion may also have played a role. crystals contain fluid inclusions, thus emphasizing the prominent role
Nitrogen aggregation characteristics show that all known diamonds of fluids in diamond formation in general (Navon et al., 1988).
from UHPM terranes belong to the type 1b–1aA, suggesting a short Much of what we know about diamond-forming fluids comes
residence time (ca. 5 Ma) at high temperature conditions (~ 900– from intensive studies of fibrous and cloudy diamonds from the Earth's
1100 °C) (e.g. Cartigny et al., 2001; Dobrzhinetskaya et al., 2006). This mantle (Schrauder and Navon, 1994; Izraeli et al., 2001; Klein-BenDavid
distinguishes them from many kimberlitic monocrystalline diamonds, et al., 2004; Zegdenizov et al., 2004; Rege et al., 2005; Tomlinson et al.,
which belong to type 1aAB and are interpreted to have had longer 2006, 2009; Logvinova et al., 2008; Weiss et al., 2009). Fibrous
diamonds form in the same environments in the Earth's mantle as
“normal” monocrystalline diamonds, but represent an extreme end of
this process, providing a different perspective. Although they are only
a minor component of the diamond population in Group I kimberlites
(Boyd et al., 1994), fibrous stones are ideal study objects for nano-
methods, because they can contain thousands of fluid inclusions.
For the first time, direct observation of diamond fluids was possible
and these studies significantly extended our knowledge, which had
previously been based solely upon indirect evidence from carbon and
nitrogen and their isotopes within the diamond itself.
Many studies worldwide revealed that the fluids trapped in these
stones range compositionally between three end-members, namely a
carbonatitic end-member rich in Ca, Mg, Fe, K and CO2, a hydrous-
silicic end-member rich in Si, Al, K and H2O, and a hydrous-saline fluid
rich in Cl, K, Na and H2O (Fig. 9; Klein-BenDavid et al., 2004). Most
recently, high Mg/Ca carbonatitic fluid inclusions were reported
(Klein-BenDavid et al., 2008; Kopylova et al., 2010; Weiss et al., 2011),
which perhaps represent an additional end-member for these fluids,
situated deeper in the lithosphere.
A characteristic feature of these fluids is a pronounced miscibility
gap between the silicate-dominated and the halide-rich end-members
at pressures and temperatures relevant to diamond formation in the
Earth's mantle, described from case studies (Izraeli et al., 2001; Klein-
Fig. 8. Comparative diagram of nitrogen content versus δ13C in diamonds from ultrahigh
pressure metamorphic terranes, polycrystalline diamonds from kimberlite (black spots, BenDavid et al., 2004, 2006, 2007) and substantiated experimentally
Jacob et al. unpublished SIMS data, open circles, Mikhail et al., 2013) and worldwide (Safonov et al., 2007). Most fibrous diamond fluid compositions fall
monocrystalline diamonds from kimberlitic and lamproitic sources (Field 3). Fields 1–3 along mixing lines between either the carbonatitic-hydrous saline or
represent Kokchetav massif, Kazakhstan and are adopted from Cartigny et al. (2001, the carbonatitic-hydrous silicic end-members for which contradictory
2003). Field 1 (black squares): diamonds from grt pyroxenites; Field 2 (gray squares):
alluvial microdiamonds, and diamonds from marbles (open squares). Field 4 is from
explanations exist. Silicates are the liquidus phases in carbonatite
Dobrzhinetskaya et al. (2010) and represents Erzgebirge microdiamonds included in melts (Safonov et al., 2007; Litasov and Ohtani, 2009); therefore
garnets from quartz–feldspathic gneisses. diamond crystallization upon successive cooling should produce a
D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35 31

6.3. Oxidation conditions recorded by polycrystalline diamonds: a matter


of scale

Macro-inclusions indicate that diamond has a large stability field


stretching over more than 10 log units in oxygen fugacity (fO2), from
conditions of about six units below the iron–wüstite oxygen buffer,
where moissanite (SiC) is stable (Moore and Gurney, 1989; Ulmer
et al., 1998) to distinctly oxidized conditions where carbonate inclu-
sions are present, at an oxygen fugacity of about 4 log units above the
iron–wüstite equilibrium (Fig. 7). Depending on the exact bulk compo-
sition, even higher fO2 is suggested by rare pure CO2 inclusions in
diamond (Schrauder and Navon, 1994; Chinn, 1995). An important
insight from the application of FIB–TEM methods to diamond was that
in some cases the nano- and micro-inclusion suite within a single
diamond can trace fO2 conditions covering the entire 10 log units fO2
of the diamond stability field (Klein-BenDavid et al., 2007).
Most microinclusions (fluid and solid alike) record oxidized condi-
tions; however, there is also a pronounced record of heterogeneous of
Fig. 9. Fluid end-member compositions derived from studies on fibrous diamonds (bold fO2 conditions from microinclusions. Similarly, macro-inclusion suites
dashed line: Izraeli et al., 2001; Klein-BenDavid et al., 2004) and experimental results of also suggest that most of their host diamonds formed by reduction of
Safonov et al. (2007, solid line — 1600 °C, dashed line — 1500 °C, stippled line — an oxidized medium (Stachel and Harris, 2008), but inclusion evidence
1400 °C) showing the compositional range of fluid inclusions in diamonds. Note the for very reducing conditions can be found, too. However, this skewed
large miscibility gap between the saline and the siliceous end-member in this system.
Overlain are the general range of inclusion compositions for UHP metamorphic diamond
distribution towards oxidized carbon sources is not reflected by
inclusions from metacarbonate rocks (Kokchetav Massif — light gray field) and feldspathic the inclusions in polycrystalline diamonds from the Earth's mantle.
gneisses (Kokchetav Massif and Erzgebirge — dark gray fields) and the approximate Here, inclusions that reflect very reducing conditions appear to be
composition for the polycrystalline diamond aggregate from Orapa (silicate inclusions as just as common as those formed under oxidizing conditions. Native
dark gray field connected with a tie line to the single carbonatic inclusion (star)).
metals and alloys (e.g. Fe-Ni alloy, Sn) are found in carbonados but
so are carbonates (Kaminsky et al., 2013). Carbonates and magnetite
progression from silicate-rich to carbonate-rich inclusions. However, or cohenite and native iron occur in polycrystalline diamonds from
this contradicts other models (Schrauder and Navon, 1994) that argued kimberlites (Jacob et al., 2004, 2011). The large range of oxygen fugacity
for fractional crystallization of a carbonatitic liquid, which would result recorded by polycrystalline diamonds suggests a prominent role
in an opposite trend in inclusion chemistry. of redox gradients and transient, small-scale equilibria in their
formation.
6.2. Diamond fluids in polycrystalline diamonds Compositions of mantle volatiles in equilibrium with diamond
vary at the relevant pressures and temperatures depending on oxygen
The fluid compositions from polycrystalline diamonds from Orapa fugacity. While they consist of pure CO2 at oxidizing conditions, they
plot towards the hydrous-silicic end-member of these compositions comprise mixtures of CH4, H2O and H2 close to the iron–wüstite buffer
and lack the halide component (Fig. 4). Potassium is present in many and below (Fig. 7; Saxena, 1989; Taylor, 1990; Sokol et al., 2009;
of the nano-inclusions in this sample, but also Ca, Mg and particularly Zhang and Duan, 2009). Under oxidizing conditions the CO2-rich
Fe, which are more typical for the carbonate-dominated fluids in fibrous volatiles form very mobile carbonatitic melts that provide efficient
and cloudy diamonds. Although it is too early to draw general conclu- transport media for carbon in the mantle (Green and Wallace, 1988).
sions for the polycrystalline diamonds from kimberlites, these observa- In a reducing environment, the main carriers of carbon are methane-
tions show that this particular fluid bears general similarities with those rich fluids (Jakobsson and Holloway, 1986; Sokol et al., 2009; Foley,
described from the fibrous and cloudy diamonds in representing a 2011), because reducing silicic melts dissolve b100 to 2000 ppm carbon
composition close to the silicate-dominated end-member, although (Taylor and Green, 1987; Ardia et al., 2013). Reduction of carbonatitic
different in divalent cation composition. melts or oxidation of methane-bearing fluids results in the formation
In contrast, fluid inclusions in the carbonado suite studied here (not of carbon-saturated aqueous fluids (Fig. 7) from which diamond can
shown) cover the broad spectrum of fluid compositions depicted in precipitate across a wide pressure and temperature range.
Fig. 9, including the apparent lack of compositions plotting into the For reducing conditions more than 3 orders of magnitude below the
miscibility gap. water maximum in Fig. 7, fO2 changes upon diamond precipitation
The hypothesis that worldwide diamond fluids comprise silicates, occur from a fluid with near-constant C/H ratio, which facilitates the
water, carbonate and halogens also holds for polycrystalline diamonds precipitation of suites of inclusions with variable oxidation states.
from crustal rocks. Diamonds from metacarbonates of the Kokchetav Here, rapid diamond crystallization may occur with essentially
UHP terrane contain hydrous-saline and carbonatic inclusions, while unchanging fluid composition if the fluid throughput is high. The
silicate-bearing fluid inclusions are rare. Felsic gneisses from the rapid change in fluid composition 2 orders of magnitude below the
Kokchetav and Erzgebirge massifs, on the other hand, have inclusions water maximum would be associated with rapid diamond crystalliza-
in diamond that are of hydrous-silicic or hydrous-saline compositions. tion and entrapment of inclusions slightly above iron–wüstite. The
The exact composition for each individual nano-inclusion in diamond situation is different for oxidizing conditions above the water maxi-
from the felsic gneisses was difficult to assess, and the exact composi- mum, where a fluid in equilibrium with diamond has quite constant
tional range distribution between the individual end-members remains fO2, but evolves from CO2- to H2O-rich compositions. Thus, inclusion
unclear as of yet. The large miscibility gap at 5 GPa in Fig. 9 persists suites in diamonds formed on the oxidized side of this diagram will
towards lower temperatures and pressures (Suk, 2001); however, the show a more uniform oxidation state.
relevant experimental studies did not include water. Safonov et al. Particularly these latter conditions apply to UHPM polycrystalline
(2007) and Litasov and Ohtani (2009) suggested that the presence of diamonds. Neither native metals nor alloys and carbides have been
water would significantly increase the miscibility of Si- and Cl-bearing found in these diamonds; instead, they are dominated by inclusions
liquids. that indicate high oxygen fugacity (e.g. carbonates), close to the FMQ
32 D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35

oxygen buffer. As predicted by the C–H–O speciation diagram in Fig. 7, from the carbon-bearing fluids, in both the marbles and the felsic
fO2 variability is minimal in inclusion suites from these environments. gneisses corresponds closely to those of their respective hosts. Major
The recent discovery by Bali et al. (2013) that water and hydrogen fluid species identified by FTIR spectroscopy are carbonate and water in
are likely to be immiscible under the conditions of the mantle geotherm inclusions from marbles and felsic gneisses alike (de Corte et al., 1999;
may help to provide an explanation for the extreme small-scale hetero- Kikuchi et al., 2005; Dobrzhinetskaya et al., 2006; Dobrzhinetskaya and
geneities in fO2 conditions represented by the inclusions in polycrystal- Wirth, 2009; Sitnikova and Shatsky, 2009). In order to explain the
line diamonds. During precipitation of diamond along the saturation differences in diamond grade between directly neighboring lithologies,
curve (Fig. 7), fluids become enriched in H2O, so that hydrogen is likely however, cm-scale heterogeneities in fluid species, e.g. in XCO2, must
to be released: the differential mobility of water and hydrogen (Bali have existed during peak metamorphism (Ogasawara et al., 2000;
et al., 2013) will lead to the production of very reducing environments Ogasawara and Aoki, 2005).
on the scale of nanometers to centimeters, promoting the crystallization Nevertheless, despite the majority of evidence pointing to local
of nanophases such as metals, carbides and nitrides. sources for the carbon-bearing fluids in these UHP terranes, noble gas
isotopic studies (Sumino et al., 2011) suggested the involvement of an
6.4. Chemical environments for diamond formation exotic, mantle-derived component at least for the Kokchetav massif,
which is the only UHPM locality for which a noble gas study was carried
Decades of studies of carbon and nitrogen in diamond have out to date. A major challenge for the future will be to carefully decipher
underlined the complications associated with deciphering the origins these sources of mixed information.
of its ingredients (for recent reviews see Stachel and Harris, 2008;
Shirey et al., 2013 and references therein). As carbon is subjected to
efficient deep recycling on Earth (Javoy et al., 1982), there is no question 7. Conclusions
that sources of carbon in diamond are multiple, comprising both mantle
and subducted crustal components. Source characteristics for carbon State-of-the-art analytical techniques with high spatial resolution
and nitrogen, however, are often encrypted by mixing and fractionation have deepened our understanding of natural diamond and its inclu-
processes upon subduction and at mantle temperatures (Deines, 1980). sions, while at the same time they highlight the complexity of the pro-
It is likely that the silicic, carbonatitic and saline compositions of fluids cesses at work. Taking microanalysis to the submicron scale in diamond
included in diamonds are subjected to similarly complex processes, research up to now benefited in particular from TEM analysis using FIB
which complicate interpretations. Silicate and carbonatitic liquids prepared electron-transparent samples. Its application to carbonado
have long been recognized as major metasomatic agents in the mantle. and to polycrystalline diamond from kimberlite allowed syngenetic
In comparison, the roles of halides and halogens in the Earth's mantle inclusions to be differentiated from epigenetic material, which proved
are much more enigmatic and as of yet data on the deep halogen particularly critical to test models for the origin of carbonado, finally
cycle is virtually non-existent (Izraeli et al., 2001; Kamenetsky et al., resolving the Earth's mantle as its origin.
2004; Litasov and Ohtani, 2009). In situ sampling of diamond fluids revealed a large heterogeneity in
Volatiles, even when present in only small amounts, have large redox-conditions and fluid chemical compositions at small scale, which
effects on the melting points of peridotite and ultramafic rocks in the is not reflected in the macro-inclusion suite.
mantle (Green, 1973; Eggler, 1976; Wyllie, 1978; Taylor and Green, However, despite originating from such chemically different
1988; Dasgupta and Hirschmann, 2006; Foley, 2011). Precipitation of environments as the mantle and the crust, fluids from polycrystalline
diamond from a C–H–O fluid via redox reaction rapidly increases the diamonds have compositions that conform with the fluid end-
water activity of the fluid (Fig. 7), which may promote melting of the members established by studies on fibrous diamonds, thus suggesting
surrounding mantle material. This process couples the precipitation of a universally important role of a limited number of basic ingredients,
diamond with melting caused by the increase in water activity of the namely carbonates, silicates, halides and water.
fluid. It occurs in both reduced and oxidized conditions, but will be Strong redox gradients reflected by the micro-inclusions indicate
particularly effective in oxidized conditions where the depression of diamond precipitation via small-scale, ephemeral redox processes driven
melting temperatures in the presence of both H2O and CO2 is stronger by the contrasting oxidation states of fluids and their depositional
(Foley et al., 2009). This process results in the generation of silicate environment. The susceptibility of the melting point of Earth's mantle
partial melts simultaneous with diamond precipitation that mix with rocks to the presence of even small amounts of volatiles promotes
the diamond fluid and, thus, are likely to be included into the growing melting simultaneous to diamond precipitation, further enhanced by
diamonds. It is therefore hypothesized here that syngenetic solid the changing of fluid composition towards higher water activity. This
phase inclusions in diamonds generally represent mixtures of compo- creates a chemically heterogeneous environment, in which diamonds
nents carried in the fluid and material produced by partial melting of and their inclusions are precipitated via redox-freezing processes.
surrounding rocks at the site of diamond precipitation on a very local Recent experimental work implies that the generation of ephemeral
scale. small-scale redox gradients may be even further enhanced by the
The composition of a given inclusion suite would then depend on immiscibility of water and hydrogen in the diamond fluids at these
both the chemical load of the carbon-bearing fluid which may vary pressures and temperatures (Bali et al., 2013).
depending on its oxidation state and on the degree of partial melting The submicron inclusion suites are therefore products of redox reac-
of the local rock. Higher degrees of melting lead to inclusion suites tions and mixing of components from the C–H–O fluid, which can be
dominated by silicate material typical for the local environment, thus foreign, and those derived locally from the surrounding mantle rocks.
with compositions representing the Earth's mantle (eclogitic–peridotitic). Simple mass balance considerations predict that in instances where
Lower degrees of melting of local rock generate inclusions that are more partial melting of the host rock dominates, the fluid component is
representative for the chemical composition of the original C–H–O fluid occluded in the resulting inclusion chemistry. In these cases the partic-
and its solutes. However, in either case, the redox conditions recorded ular inclusion only bears information on the depositional environment
by the syngenetic inclusion suite are a result of the diamond precipita- rather than on the carbon-bearing medium.
tion process and bear little information on the original oxygen fugacity In contrast to polycrystalline diamonds from the mantle, those from
of the diamond's environment. UHPM terranes carry evidence for homogeneously oxidizing formation
For the UHPM polycrystalline diamonds, the marble host rocks conditions and carry a strong local signature in their submicron
present potential near-infinite reservoirs of carbon. Indeed, the major inclusion suite. Still, rare gas isotopes suggest the seemingly occluded
and trace element chemistry of the solid nano-inclusions precipitated involvement of a mantle fluid in their formation, suggesting that a
D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35 33

foreign fluid component may merely be masked by dilution in the major Dobosi, G., Kurat, G., 2002. Trace element abundances in garnets and clinopyroxenes from
diamondites — a signature of carbonatitic fluids. Mineral. Petrol. 76, 21–38.
and trace element compositions. Dobosi, G., Kurat, G., 2010. On the origin of silicate-bearing diamondites. Mineral. Petrol.
Development in modern microanalytical techniques allows in situ 99, 29–42.
analysis of diamonds and their inclusions spanning roughly five orders Dobrzhinetskaya, L.F., 2012. Microdiamonds — frontiers of ultrahigh-pressure metamor-
phism: a review. Gondwana Res. 21, 207–223.
of magnitude in spatial resolution. Information derived from inclusion Dobrzhinetskaya, L.F., Wirth, R., 2009. Ultradeep rocks and diamonds in the light of
suites at either end of the size range complements each other. Macro- advanced scientific technologies. In: Cloetingh, S., Negendank, J. (Eds.), New Frontiers
inclusions, for once, give critical insight into the chemical environment in Integrated Solid Earth Sciences. Springer, pp. 373–395.
Dobrzhinetskaya, L.F., Green II, H.W., Bozhilov, K.N., Mitchell, T.E., Dickerson, R.M., 2003.
of diamond formation as well as the pressure and temperature range Crystallization environment of Kazakhstan microdiamond: evidence from
of diamond growth. On the other hand, micro-inclusions at the lower nanometric inclusions and mineral associations. J. Metamorph. Geol. 21, 425–437.
end of the size distribution deliver direct information on the diamond Dobrzhinetskaya, L.F., Wirth, R., Green II, H.W., 2005. Direct observation and analysis of
trapped COH fluid growth medium in metamorphic diamond. Terra Nova 17, 472–477.
growth medium represented by included minerals, quench phases
Dobrzhinetskaya, L.F., Liu, Z., Cartigny, P., Zhang, J., Tchkhetia, N.N., Green II, H.W., Hemley,
and fluid, which cannot be derived from macroinclusions alone. R.J., 2006. Synchrotron infrared and Raman spectroscopy of microdiamonds from
Erzgebirge, Germany. Earth Planet. Sci. Lett. 248, 340–349.
Dobrzhinetskaya, L.F., Wirth, R., Green, H.W., 2007. A look inside of diamond-forming
Acknowledgments media in deep subduction zones. Proc. Natl. Acad. Sci. U. S. A. 104, 9128–9132.
Dobrzhinetskaya, L.F., Wirth, R., Yang, J., Hutcheon, I.D., Weber, P.K., Green II, H.W., 2009.
DJ gratefully acknowledges financial support from DFG grant JA781/ High-pressure highly reduced nitrides and oxides from chromitite of a Tibetan
ophiolite. Proc. Natl. Acad. Sci. U. S. A. 106 (46), 19233–19238.
9-1 (Heisenberg Professur), the Earth System Science Research Centre
Dobrzhinetskaya, L.F., Green, H.W., Takahata, N., Sano, Y., Shirai, K., 2010. Crustal signature
at Johannes-Gutenberg University, Mainz and Macquarie University, of delta C-13 and nitrogen content in microdiamonds from Erzgebirge, Germany: ion
Sydney. De Beers is thanked for the accessibility of suitable sample microprobe studies. J. Earth Sci. 21, 623–634.
material. LD acknowledges support from US National Science Foundation Dobrzhinetskaya, L.F., Wirth, R., Green II, H.W., Schreiber, A., Obannon, E., 2013. First find
of polycrystalline diamond in ultrahigh-pressure metamorphic terrane of Erzgebirge,
(grants EAR-0408505, EAR-1118796) and the Lab Fee Research Germany. J. Metamorph. Geol. 31 (1), 5–18.
Award (No. 09-LR-05-116946-DOBL). Anja Schreiber (GFZ Potsdam) is Eggler, D.H., 1976. Does CO2 cause partial melting in the low-velocity layer of the mantle?
thanked for skilful FIB sample preparation. We are grateful to Oded Geology 4, 69–72.
Ernst, W.G., 2006. Preservation/exhumation of ultrahigh-pressure subduction complexes.
Navon and Andy Moore who reviewed an earlier version of this ms. Lithos 92, 321–335.
This is contribution 460 from the ARC Centre of Excellence for Core to Ernst, W.G., Liou, J.G., 2008. High- and ultrahigh-pressure metamorphism — past results,
Crust Fluid Systems (http://www.ccfs.mq.edu.au). future prospects. Am. Mineral. 93, 1771–1786.
Floss, C., Stadermann, F.J., Bradley, J.P., Zu, R.D., Bajt, S., Graham, G., Lea, A.S., 2006.
Identification of isotopically primitive interplanetary dust particles: a NanoSIMS
References isotopic imaging study. Geochim. Cosmochim. Acta 70, 2371–2399.
Foley, S.F., 2011. A reappraisal of redox melting in the Earth's mantle as a function of
Ardia, P., Hirschmann, M.M., Withers, A.C., Stanley, B.D., 2013. Solubility of CH4 in a tectonic setting and time. J. Petrol. 52, 1363–1391.
synthetic basaltic melt, with applications to atmosphere-magma ocean-core Foley, S.F., Yaxley, G.M., Rosenthal, A., Buhre, S., Kiseeva, E.S., Rapp, R.P., Jacob, D.E., 2009.
partitioning of volatiles and to the evolution of the Martian atmosphere. Geochim. The composition of near-solidus melts of peridotite in the presence of CO2 and H2O
Cosmochim. Acta 114, 52–71. between 40 and 60 kbar. Lithos 112S1 (Special issue 9IKC), 247–283.
Aulbach, S., Stachel, T., Viljoen, K.S., Brey, G.P., Harris, J.W., 2002. Eclogitic and websteritic Frantz, J.D., Mao, H.K., Zhang, Y., Wu, Y., Thompson, A.C., Underwood, J.H., Giauque, R.D.,
diamond sources beneath the Limpopo Belt — is slab-melting the link? Contrib. Jones, K.W., Rivers, M.-L., 1988. Analysis of fluid inclusions by X-ray fluorescence
Mineral. Petrol. 114, 331–348. using synchrotron radiation. Chem. Geol. 69, 235–244.
Bali, E., Audetat, A., Keppler, H., 2013. Water and hydrogen are immiscible in Earth's Garai, J., 2012. Ionized nitrogen mono-hydride bands are identified in the presolar and
mantle. Nature 495, 220–223. carbonado diamond spectra. Meteorit. Planet. Sci. 47, 1–7.
Boyd, S.R., Pineau, F., Javoy, M., 1994. Modelling the growth of natural diamonds. Chem. Garai, J., Haggerty, S.E., Rekhi, S., Chance, M., 2006. Infrared absorption investigations
Geol. 116, 29–42. confirm the extraterrestrial origin of carbonado diamonds. Astrophys. J. 653, L153–L156.
Cartigny, P., 2005. Stable isotopes and the origin of diamond. Elements 1, 79–84. Gautheron, C., Cartigny, P., Moreira, M., Harris, J.W., Allegre, C.J., 2005. Evidence for a
Cartigny, P., 2010. Mantle-derived carbonados? Geochemical insights from diamonds mantle component shown by rare gases, C and N isotopes in polycrystalline
from the Dachine komatiite (French Guiana). Earth Planet. Sci. Lett. 296, 329–339. diamonds from Orapa (Botswana). Earth Planet. Sci. Lett. 240, 559–572.
Cartigny, P., De Corte, K., Shatsky, V.S., Ader, M., De Paepe, P., Sobolev, N.V., Javoy, M., Giannuzzi, L.A., Stevie, F.A., 2005. Introduction to Focused Ion Beams: Instrumentation,
2001. The origin and formation of metamorphic microdiamonds from the Kokchetav Theory, Techniques and Practice. Springer.
massif, Kazakhstan: a nitrogen and carbon isotopic study. Chem. Geol. 176, 265–281. Green, D.H., 1973. Experimental melting studies on a model upper mantle composition at
Cartigny, P., Harris, J.W., Taylor, A., Davies, R., Javoy, P., 2003. On the possibility of a kinetic high pressures under water-saturated and water-undersaturated conditions. Earth
fractionation of nitrogen stable isotopes during natural diamond growth. Geochim. Planet. Sci. Lett. 19, 37–53.
Cosmochim. Acta 67, 1571–1576. Green, D.H., Wallace, M.E., 1988. Mantle metasomatism by ephemeral carbonatite melts.
Chinn, I.L., 1995. A Study of Unusual Diamonds from the George Creek K1 Kimberlite Nature 336, 459–462.
Dyke, Colorado. (Ph.D. Thesis) University of Cape Town, p. 94. Gurney, J.J., Boyd, F.R., 1982. Mineral Intergrowths with Polycrystalline Diamonds from
Cnudde, V., Dewanckele, J., De Boever, W., Brabant, L., De Kock, T., 2012. 3D characterization the Orapa Mine, Botswana. Carnegie Institution Year Book, pp. 267–273.
of grain size distributions in sandstone by means of X-ray computed tomography. In: Gurney, J.J., Jacob, W.R.O., Dawson, J.B., 1979. Megacrysts from the Monastery kimberlite
Sylvester, P. (Ed.), Quantitative Mineralogy and Microanalysis of Sediments and pipe, South Africa. In: Boyd, F.R., Meyer, H.O.A. (Eds.), Proceedings 2nd International
Sedimentary Rocks. short course series, vol. 42, pp. 99–113. Kimberlite Conference, 2. AGU, pp. 227–243.
Dasgupta, R., Hirschmann, M.M., 2006. Melting in the Earth's deep upper mantle caused Gurney, J.J., Harris, J.W., Rickard, R.S., 1984. Silicate and oxide inclusions in diamonds from
by carbon dioxide. Nature 440, 659–662. the Orapa Mine, Botswana. In: Kornprobst, J. (Ed.), Kimberlites II: The Mantle and
Dasgupta, R., Hirschmann, M.M., Withers, A.C., 2004. Deep global cycling of carbon Crust–Mantle Relationships. Elsevier, Amsterdam.
constrained by the solidus of anhydrous, carbonated eclogite under upper mantle Haggerty, S.E., 1999. A diamond trilogy: superplumes, supercontinents and supernova.
conditions. Earth Planet. Sci. Lett. 227, 73–85. Science 289, 852–860.
De Corte, K., Cartigny, P., Shatsky, V.S., De Paepe, P., Sobolev, N.V., Javoy, M., 1999. Charac- Haggerty, S.E., 2014. Carbonado: physical and chemical properties, a critical evaluation of
teristics of microdiamonds from UHPM rocks of the Kokchetav massif (Kazakhstan). proposed origins, and a revised genetic model. Earth-Sci. Rev. 130, 49–72.
In: Gurney, J.J., Gurney, J.L., Pascoe, M.D., Richardson, S.H. (Eds.), Proceedings of the Harris, J.W., 1968. The recognition of diamond inclusions. Industrial Diamond
VIIth International Kimberlitic Conference. Red Roof Design cc, Cape Town, South ReviewDeBeers Industrial Diamond Division.
Africa, pp. 174–182. Heaney, P.J., Vicenzi, E.P., De, S., 2005. Strange diamonds: the mysterious origins of
De Souza Martins, M., 2006. Geologia dos Diamantes e Carbonados Aluvionares da Bacia carbonado and framesite. Elements 1 (2), 85–89.
do Rio Macaubas (MG). Tese de DoutoramentoUniversidade Federal De Minas Gerais, Hoppe, P., Cohen, S., Meibom, A., 2013. NanoSIMS: technical aspects and applications in
Instituto de Geociencias, Belo Horizonte, Minas Gerais, Brazil, (Ph.D. thesis). cosmochemistry and biological geochemistry. Geostand. Geoanal. Res. 37, 111–154.
De, S., Heaney, P.J., Hargraves, R.B., Vicenzi, E.P., Taylor, P.T., 1998. Microstructural obser- Howell, D., O'Neill, C.J., Grant, K.J., Griffin, W.L., Pearson, N.J., O'Reilly, S.Y., 2012. μ-FTIR
vations of polycrystalline diamond: a contribution to the carbonado conundrum. mapping: distribution of impurities in different types of diamond growth. Diam.
Earth Planet. Sci. Lett. 164, 421–433. Relat. Mater. 29, 29–36.
Deines, P., 1980. The carbon isotopic composition of diamonds — relationship to diamond Imamura, K., Ogasawara, Y., Yurimoto, H., Kusakabe, M., 2013. Carbon isotope heteroge-
shape, color, occurrence and vapor composition. Geochim. Cosmochim. Acta 44, neity in metamorphic diamond from the Kokchetav UHP dolomite marble, northern
943–961. Kazakhstan. Int. Geol. Rev. 55, 453–467.
Desbois, G., Urai, J.L., Burkhardt, C., Drury, M.R., Hayles, M., Humbel, B.M., 2008. Cryogenic Ishibashi, H., Kagi, H., Sakuai, H., Ohfuji, H., Sumino, H., 2012. Hydrous fluid as the growth
vitrification and 3D serial sectioning using high cryo FIB SEM technology for brine- media of natural polycrystalline diamond, carbonado: implication from IR spectra
filled grain boundaries in Halite: first results. Geofluids 8, 60–72. and microtextural observations. Am. Mineral. 97, 1366–1372.
34 D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35

Ishida, H., Ogasawara, Y., Ohsumi, K., Saito, A., 2003. Two stage growth of microdiamond Moore, R.O., Gurney, J.J., 1989. Mineral inclusions in diamond from the Monastery
in UHP dolomite marble from Kokchetav Massif, Kazakhstan. J. Metamorph. Geol. 21, kimberlite, South Africa. Spec. Publ. Geol. Soc. Aust. 14, 1029–1041.
515–522. Navon, O., 1999. Diamond formation in the Earth's mantle. Proc 7th Int Kimberlite Conf, 2,
Izraeli, E.S., Harris, J.W., Navon, O., 2001. Brine inclusions in diamond: a new upper mantle pp. 584–604.
fluid. Earth Planet. Sci. Lett. 187, 323–332. Navon, O., Hutcheon, I.D., Rossman, G.R., Wasserburg, G.J., 1988. Mantle-derived fluids in
Jacob, D.E., 2004. Nature and origin of eclogites from the Earth's mantle. Lithos 77, diamond micro-inclusions. Nature 355, 784–789.
295–316. Odling, N.W.A., 1989. The Role of Sulphur in Upper Mantle Processes. (Ph.D. Thesis)
Jacob, D.E., Viljoen, K.S., Grassineau, N., Jagoutz, E., 2000. Remobilization in the cratonic University of Tasmania.
lithosphere recorded by polycrystalline diamond (framesite). Science 289, Ogasawara, Y., Aoki, K., 2005. The role of fluid for diamond-free UHP dolomite marble
1182–1185. from Kokchetav Massif. Int. Geol. Rev. 47, 1178–1193.
Jacob, D.E., Kronz, A., Viljoen, K.S., 2004. Cohenite, native iron and troilite inclusions in Ogasawara, Y., Ohta, M., Fukusava, K., Katayama, I., Maruyama, S., 2000. Diamond-bearing
garnets from polycrystalline diamond aggregates. Contrib. Mineral. Petrol. 146, and diamond-free metacarbonate rocks from Kumdy-Kol from the Kokchetav massif,
566–576. northern Kazakhstan. Island Arc 9, 400–416.
Jacob, D.E., Wirth, R., Enzmann, F., Kronz, A., Schreiber, A., 2011. Nano-inclusion suite and Orlov, J.L., 1977. The Mineralogy of Diamond. Wiley, New York.
high resolution micro-computed-tomography of polycrystalline diamond (framesite) Ozima, M., Tatsumoto, M., 1997. Radiation-induced diamond crystallization: origin of
from Orapa, Botswana. Earth Planet. Sci. Lett. 308 (3–4), 307–316. carbonados and its implications on meteorite nano-diamonds. Geochim. Cosmochim.
Jakobsson, S., Holloway, J., 1986. Crystal-liquid experiments in the presence of a C–O–H Acta 61, 369–376.
fluid buffered by graphite + iron + wüstite: experimental method and near liquidus Ozima, M., Zashu, S., Tomura, K., Matsuhisa, Y., 1991. Constraints from noble-gas contents
relations in basanite. J. Volcanol. Geotherm. Res. 29, 265–291. on the origin of carbonado diamonds. Nature 351, 472–474.
Javoy, M., Pineau, F., Allegre, C.J., 1982. Carbon geodynamic cycle. Nature 300, 171–173. Rege, S., Jackson, S., Griffin, W.L., Davies, R.M., Pearson, N.J., O'Reilly, S.Y., 2005. Quantita-
Jones, R., Briddon, P.R., Oberg, S., 1992. First-principles theory of nitrogen aggregates in tive trace-element analysis of diamond by laser ablation inductively coupled plasma
diamond. Philos. Mag. Lett. 66, 67–74. mass spectrometry. J. Anal. At. Spectrom. 20, 601–611.
Kamenetsky, M.B., Sobolev, A.V., Kamenetsky, V.S., Maas, R., Danyushevsky, L.V., Thomas, Rondeau, B., Sautter, V., Barjon, J., 2008. New columnar texture of carbonado:
R., Sobolev, N.V., Pokhilenko, N.P., 2004. Kimberlite melts rich in alkali chlorides and cathodoluminescence study. Diamond Relat. Mater. 17, 1897–1901.
carbonates: a potent metasomatic agent in the mantle. Geology 32, 845–848. Rubanova, E.V., Piazolo, S., Griffin, W.L., O'Reilly, S.Y., 2012. Deformation microstructures
Kaminsky, F.V., 1987. Origin of polycrystalline carbonado diamond aggregates. Dokl. Earth reveal a complex mantle history for polycrystalline diamond. Geochem. Geophys.
Sci. 294, 122–123. Geosyst. 13, Q10010. http://dx.doi.org/10.1029/2012GC004250.
Kaminsky, F.V., 1991. Carbonado and yakutite: properties and possible genesis. In: Meyer, Safonov, O.G., Perchuk, L.L., Litvin, Y.A., 2007. Melting relations in the chloride–carbonate–
H.O.A., Leonardo, O.H. (Eds.), Proc. of the 5th International Kimberlite Conference, 2, silicate systems at high-pressure and the model for formation of alkalic diamond-
pp. 136–143. forming liquids in the upper mantle. Earth Planet. Sci. Lett. 253, 112–128.
Kaminsky, F.V., Wirth, R., 2011. Iron carbide inclusions in lower-mantle diamond from Sano, Y., Yokochi, R., Terada, K., Chaves, M.L., Ozima, M., 2002. Ion microprobe Pb–Pb
Juina, Brazil. Can. Mineral. 49 (2), 555–572. dating of carbonado, polycrystalline diamond. Precambrian Res. 113, 155–168.
Kaminsky, F., Matzel, J., Jacobsen, B., Hutcheon, I.D., Wirth, R., 2012. First data on isotopic Sautter, V., Lorand, J.P., Cordier, P., Rondeau, B., Leroux, H., Ferraris, C., Pont, S., 2011.
compositions of oxygen and carbon in lower-mantle carbonates and silicates. Petrogenesis of mineral micro-inclusions in an uncommon carbonado. Eur. J. Mineral.
Abstract, Goldschmidt Conference, Montreal, Canada, June 2012. 23, 721–729.
Kaminsky, F.V., Wirth, R., Morales, L., 2013. Internal texture and syngenetic inclusions in Saxena, S.K., 1989. Oxidation state of the mantle. Geochim. Cosmochim. Acta 53, 89–95.
carbonado. Can. Mineral. 51, 39–55. Schertl, P.P., Sobolev, N.V., 2013. The Kokchetav Massif, Kazakhstan: "Type locality" of
Ketcham, R.A., Koeberl, C., 2013. New textural evidence on the origin of carbonado diamond-bearing UHP metamorphic rocks. Journal of Asian Earth Sciences 63,
diamond: an example of 3D petrography using X-ray computed tomography. 5–38.
Geosphere. http://dx.doi.org/10.1130/GES00908.1. Schrauder, M., Navon, O., 1994. Hydrous and carbonatitic mantle fluids in fibrous
Kikuchi, M., Kagi, H., Ogasawara, Y., Dobrzhinetskaya, L., 2005. FTIR study of diamonds from Jwaneng, Botswana. Geochim. Cosmochim. Acta 58, 761–771.
microdiamonds from garnet–biotite gneisses of the Kokchetav massif, Kazakhstan. Shirey, S.B., Cartigny, P., Frost, D.J., Keshav, S., Nestola, F., Nimis, P., Pearson, D.G., Sobolev,
EOS, AGU Meeting, p. F-1237. N.V., Walter, M.J., 2013. Diamonds and the geology of mantle carbon. Rev. Mineral.
Klein-BenDavid, O., Wirth, R., Izraeli, E.S., Hauri, E., Navon, O., 2004. Brine and carbonatitic Geochem. 75, 355–421.
melts in a diamond from Diavik — implications for mantle fluid evolution. Geochim. Sitepu, H., Kopylova, M.G., Quirt, D.H., Cutler, J.N., Kotzer, T., 2005. Synchrotron micro-X-
Cosmochim. Acta 68, A276. ray fluorescence analysis of natural diamonds: first steps in identification of mineral
Klein-BenDavid, O., Wirth, R., Navon, O., 2006. TEM imaging and analysis of inclusions in situ. Am. Mineral. 90, 1740–1747.
microinclusions in diamonds: a close look at diamond-growing fluids. Am. Mineral. Sitnikova, E.S., Shatsky, V.S., 2009. New data of FTIR spectroscopy on the composition of
91 (2–3), 353–365. diamond crystallization medium in metamorphic rocks of the Kokchetav massif.
Klein-BenDavid, O., Wirth, R., Navon, O., 2007. Micrometer-scale cavities in fibrous and Russ. Geol. Geophys. 50, 842–849.
cloudy diamonds — a glance into diamond dissolution events. Earth Planet. Sci. Lett. Smith, J.V., Dawson, J.B., 1985. Carbonado: diamond aggregates from early impact of
264, 89–103. crustal rocks? Geology 13, 342–343.
Klein-BenDavid, O., Logvinova, A.M., Schrauder, M., Spetius, Z.V., Weiss, Y., Hauri, E.H., Smith, E.M., Kopylova, M.G., Dubrovinsky, L., Navon, O., Ryder, J., Tomlinson, E.L., 2011.
Kaminsky, F.V., Sobolev, N.V., Navon, O., 2008. High-Mg carbonatitic microinclusions in Transmission X-ray diffraction as a new tool for diamond fluid inclusion studies.
some Yakutian diamonds—a new type of diamond-forming fluid. Lithos 112, 648–859. Mineral. Mag. 75, 2657–2675.
Kopylova, M., Navon, O., Dubrovinsky, L., Khachatryan, G., 2010. Carbonatitic mineralogy Sobolev, N.V., 1977. Deep-seated Inclusions in Kimberlites and the Problem of the
of natural diamond-forming fluids. EPSL 291, 126–137. Composition of the Upper Mantle. AGU, New York.
Kurat, G., Dobosi, G., 2000. Garnet and diopside-bearing diamondites (framesites). Sobolev, N.V., Yefimova, E.S., Pospelova, L.N., 1989. Native iron in Yakutian diamonds and
Mineral. Petrol. 69, 143–159. its paragenesis. Sov. Geol. Geophys. 22, 18–21.
Lee, M.R., 2010. Transmission electron microscopy (TEM) of Earth and planetary Sobolev, N.V., Logvinova, A.M., Efimova, E.S., 2009. Syngenetic phlogopite inclusions in
materials: a review. Mineral. Mag. 74, 1–27. kimberlite-hosted diamonds: implications for role of volatiles in diamond formation.
Litasov, K.D., Ohtani, E., 2009. Phase relations in the peridotite–carbonate–chloride Russ. Geol. Geophys. 50, 1234–1248.
system at 7.0–16.5 GPa and the role of chlorides in the origin of kimberlite and Sokol, A.G., Palyanova, G.A., Palyanov, Y.N., Tomilenko, A.A., Melenevsky, V.N., 2009. Fluid
diamond. Chem. Geol. 262, 29–41. regime and diamond formation in the reduced mantle: experimental constraints.
Logvinova, A.M., Wirth, R., Fedorova, E.N., Sobolev, N.V., 2008. Nanometre-sized mineral Geochim. Cosmochim. Acta 73, 5820–5834.
and fluid inclusions in cloudy Siberian diamonds: new insights on diamond formation. Stachel, T., Harris, J.W., 2008. The origin of cratonic diamonds — constraints from mineral
Eur. J. Mineral. 20, 317–331. inclusions. Ore Geol. Rev. 34, 5–32.
Maruoka, T., Kurat, G., Dobosi, G., Koeberl, C., 2004. Isotopic composition of carbon in Stachel, T., Harris, J.W., Muehlenbachs, K., 2009. Sources of carbon ins inclusion bearing
diamonds of diamondites: record of mass fractionation in the upper mantle. diamonds. Lithos 112 (2), 625–637.
Geochim. Cosmochim. Acta 68, 1635–1644. Stadermann, F.J., Croat, T.K., Bernatowicz, T., Amari, S., Messenger, S., Walker, R.M., Zinner,
Massonne, H.J., 2003. A comparison of the evolution of diamondiferous quartz-rich rocks E., 2005. Supernova graphite in the NanoSIMS: carbon, oxygen and titanium isotopic
from the Saxonian Erzgebirge and the Kokchetav Massif: are so-called diamondiferous compositions of a spherule and its TiC sub-components. Geochim. Cosmochim. Acta
gneisses magmatic rocks? Earth Planet. Sci. Lett. 216, 347–364. 69, 177–188.
Massonne, H.J., 2006. Early metamorphic evolution and exhumation of felsic high- Stalder, R., Ulmer, P., Thompson, A.B., Günther, D., 2001. High pressure fluids in the
pressure granulites from the northwestern Bohemian Massif. Mineral. Petrol. 177, system MgO–SiO2–H2O under upper mantle conditions. Contrib. Mineral. Petrol.
177–202. 140, 607–618.
McCall, G.J.H., 2009. The carbonado diamond conundrum. Earth Sci. Rev. 93 (3–4), 85–91. Stöckhert, B., Duyster, J., Trepmann, C., Massonne, H.J., 2001. Microdiamond daughter
Mees, F., Swennen, R., van Geet, M., Jacobs, P., 2003. Applications of X-ray computed crystals precipitated from supercritical COH + silicate fluids included in garnet,
tomography in the geosciences. Geol. Soc. Lond. Spec. Publ. 215, 1–6. Erzgebirge, Germany. Geology 29, 391–394.
Mikhail, S., Dobosi, G., Verchovsky, A.B., Kurat, G., Jones, A.P., 2013. Peridotitic and Suk, N.I., 2001. Experimental study of liquid immiscibility in silicate–carbonate systems.
websteritic diamondites provide new information regarding mantle melting Petrology 9, 477–488.
and metasomatism induced through the subduction of crustal volatiles. Geochim. Sumino, H., Dobrzhinetskaya, L.F., Burgess, R., Kagi, H., 2011. Deep-mantle-derived noble
Cosmochim. Acta 107, 1–11. gases in metamorphic diamonds from the Kokchetav massif, Kazakhstan. Earth
Mikhail, S., Howell, D., McCubbin, F.M., 2014. Evidence for multiple diamondite-forming Planet. Sci. Lett. 307, 439–449.
events in the mantle. Am. Mineral. http://dx.doi.org/10.2138/am.2014.4778 (in Sunagawa, I., 1990. Growth and morphology of diamond crystals under stable and meta-
press). stable conditions. J. Cryst. Growth 99, 1156–1161.
D.E. Jacob et al. / Earth-Science Reviews 136 (2014) 21–35 35

Sunagawa, I., 2005. Crystals — Growth, Morphology and Perfection. Cambridge University Weiss, Y., Griffin, W.L., Bell, D.R., Navon, O., 2011. High-Mg carbonatitic melts in
Press, Cambridge. diamonds, kimberlites and the sub-continental lithosphere. Earth Planet. Sci. Lett.
Svisero, D.P., 1995. Distribution and origin of diamonds in Brazil. J. Geodyn. 20, 493–514. 309, 337–347.
Taylor, W.R., 1990. A reappraisal of the nature of fluids included by diamond — a window Weiss, Y., Kiflawi, I., Navon, O., 2013. The IR absorption spectrum of water in
to deep-seated mantle fluids and redox conditions. In: Herbert, H.K., Ho, S.E. (Eds.), microinclusion-bearing diamonds. In: Pearson, D.G., Grütter, H., Harris, J.W.,
Stable Isotopes and Fluid Processes in Mineralization. University of Western Kjarsgaard, B.A., O'Brien, H., Chalapathi Rao, N.V., Sparks, S. (Eds.), Proceedings of
Australia, Perth. 10th International Kimberlite Conference. Spec. Issue Journal of the Geological Society
Taylor, W.R., Green, D.H., 1987. The petrogenetic role of methane: effect on liquidus phase of India, vol. 1, pp. 271–280.
relations and the solubility mechanism of reduced C–H volatiles. In: Mysen, B.O. (Ed.), Wirth, R., 2004. A novel technology for advanced application of micro- and nanoanalysis
Magmatic Processes: Physicochemical Principles. Geochem. Soc, Washington, in geosciences and applied mineralogy. Eur. J. Mineral. 16 (6), 863–876.
pp. 121–138. Wirth, R., 2009. Focused Ion Beam (FIB) combined with SEM and TEM: advanced analytical
Taylor, W.R., Green, D.H., 1988. Measurement of reduced peridotite — C–O–H solidus and tools for studies of chemical composition, microstructure and crystal structure in
implications for redox melting of the mantle. Nature 332, 349–352. geomaterials on a nanometre scale. Chem. Geol. 261 (3–4), 217–229.
Tomlinson, E.L., Jones, A.P., Harris, J.W., 2006. Co-existing fluid and inclusions in mantle Wirth, R., 2010. Focused Ion Beam (FIB): site-specific sample preparation, nano-analysis,
diamond. Earth Planet. Sci. Lett. 250, 581–595. nano-characterization and nano-machining. In: Brenker, F.E., Jordan, G. (Eds.),
Tomlinson, E.L., Müller, W., EIMF, 2009. A snapshot of mantle metasomatism: trace Nanoscopic Approaches in Earth and Planetary Sciences.
element analysis of coexisting fluid (LA-ICP-MS) and silicate (SIMS) inclusions in Wirth, R., Pinti, D.L., Sano, Y., Takahata, N., Kaminsky, F., Wirth, R., Pinti, D.L., Sano, Y.,
fibrous diamonds. Earth Planet. Sci. Lett. 279, 362–372. Takahata, N., Kaminsky, F., 2007. FIB/TEM and NanoSIMS investigations of micro-
Trueb, L.F., Butterman, W.C., 1969. Carbonado: a microstructural study. Am. Mineral. 54, and nanoinclusions in diamond: new insights into the Earth's mantle and diamond
412–425. formation. Abstract AGU Fall Meeting, San Francisco. EOS, Transactions, American
Ulmer, G.C., Grandstaff, D.E., Woermann, E., Goebbels, M., Schonitz, M., Woodland, A.B., Geophysical Union, Supplement, 88, (52), pp. 10–14 (December, 2007, Abstract
1998. The redox stability of moissanite (SiC) compared with metal–metal oxide MR23D-10).
buffers at 1773 K and at pressures up to 90 kbar. N. Jahrb. Mineral. 172, 279–307. Wirth, R., Kaminsky, F., Matsyuk, S., Schreiber, A., 2009. Unusual micro- and nano-
Vicenzi, E.P., Heaney, P.J., Ketcham, R.A., Rivers, M., 2003. High-Resolution X-ray inclusions in diamonds from the Juina Area, Brazil. Earth Planet. Sci. Lett. 286 (1–2),
Microtomography of Polycrystalline Diamond. Goldschmidt Conference Abstracts, 292–303.
p. A513. Wyllie, P.J., 1978. Mantle fluid compositions buffered in peridotite — CO2–H2O by carbonates,
Weiss, Y., Kessel, R., Griffin, W.L., Kiflawi, I., Klein-BenDavid, O., Bell, D.R., Harris, J.W., amphibole, and phlogopite. J. Geol. 86, 687–713.
Navon, O., 2009. A new model for the evolution of diamond-forming fluids; evidence Zegdenizov, D.A., Kagi, H., Shatsky, V.S., Sobolev, N.V., 2004. Carbonatitic melts in cuboid
from microinclusion-bearing diamonds from Kankan, Guinea. Lithos 112, 660–674. diamonds from Udachnaya kimberlite pipe (Yakutia). Evidence from vibrational
Weiss, Y., Kiflawi, I., Navon, O., 2010. IR spectroscopy: quantitative determination of the spectroscopy. Mineral. Mag. 68, 61–73.
mineralogy and bulk composition of fluid microinclusions in diamonds. Chem. Zhang, C., Duan, Z., 2009. A model for C–O–H fluid in the Earth's mantle. Geochim.
Geol. 275 (1–2), 26–34. Cosmochim. Acta 73, 2089–2102.

Você também pode gostar