Content uploaded by Mathieu Debure
Author content
All content in this area was uploaded by Mathieu Debure on Sep 03, 2018
Content may be subject to copyright.
This article has been accepted for publication and undergone full peer review but has not
been through the copyediting, typesetting, pagination and proofreading process, which may
lead to differences between this version and the Version of Record. Please cite this article as
doi: 10.1111/sed.12531
This article is protected by copyright. All rights reserved.
DR. VINCENT MILESI (Orcid ID : 0000-0003-0873-2049)
Article type : Original Article
Formation of magnesium-smectite during lacustrine carbonates early
diagenesis: Study case of the volcanic crater lake Dziani Dzaha
(Mayotte – Indian Ocean)
VINCENT P. MILESI*1, DIDIER JÉZÉQUEL*, MATHIEU DEBURE†, PIERRE CADEAU*,
FRANÇOIS GUYOT‡, GÉRARD SARAZIN*, FRANCIS CLARET†, EMMANUELLE
VENNIN§, CARINE CHADUTEAU*, AURÉLIEN VIRGONE¶, ERIC C. GAUCHER¶ AND
MAGALI ADER*
*Institut de Physique du Globe de Paris, Sorbonne Paris Cité, Univ Paris Diderot, UMR 7154
CNRS, F-75005 Paris, France (E-mail: vmilesi@asu.edu)
†BRGM, French Geological Survey, Orléans, France
‡IMPMC, Sorbonne Université, MNHN, Paris, France
§Bourgogne University, Dijon, France
¶Total, EP CSTJF, Pau, France
1Present Address - ASU Tempe Campus, School of Earth and Space Exploration, ISTB4 –
BLDG75, 781 E Terrace Mall, Tempe, AZ, USA, 85287-6004
Associate Editor – Jim Hendry
Short Title – Mg-smectitie formation during diagenesis
ABSTRACT
The volcanic crater lake of Dziani Dzaha in Mayotte is studied to constrain the geochemical
settings and the diagenetic processes at the origin of Mg-phyllosilicates associated with
carbonate rocks. The Dziani Dzaha is characterized by intense primary productivity, volcanic
This article is protected by copyright. All rights reserved.
gases bubbling in three locations and a volcanic catchment of phonolitic/alkaline
composition. The lake water has an alkalinity of ca 0.2 mol∙L-1 and pH values of ca 9.3.
Cores of the lake sediments reaching up to one metre in length were collected and studied by
means of carbon–hydrogen–nitrogen elemental analyzer, X-ray fluorescence spectrometry
and X-ray powder diffraction. In surface sediments, the content of total organic carbon
reaches up to 20 wt.%. The mineral content consists of aragonite and hydromagnesite with
minor amounts of alkaline feldspar and clinopyroxene from the volcanic catchment. Below
30 cm depth, X-ray diffraction analyses of the <2 µm clay fraction indicate the presence of a
saponite-like mineral, a Mg-rich smectite. The saponite-like mineral accumulates at depth to
reach up to ca 30 wt.%, concurrent with a decrease of the contents of hydromagnesite and
organic matter. Thermodynamic considerations and mineral assemblages suggest that the
evolution of the sediment composition resulted from early diagenetic reactions. The
formation of the saponite-like mineral instead of Al-free Mg-silicates resulted from high
aluminum availability, which is favoured in restricted lacustrine environments hosted in
alkaline volcanic terrains commonly emplaced during early stages of continental rifting.
Supersaturation of the lake water relative to saponite is especially due to high pH values,
themselves derived from high primary productivity. This suggests that a genetic link may
exist between saponite and the development of organic-rich carbonate rocks, which may be
fuelled by the input of CO2-rich volcanic gases. This provides novel insights into the
composition and formation of saponite-rich deposits under a specific geodynamic context
such as the Cretaceous South Atlantic carbonate reservoirs.
Keywords Authigenesis, early diagenesis, lacustrine carbonates, magnesian clays, organic
matter decomposition, saponite, volcanic lake.
INTRODUCTION
Magnesium-rich clay minerals, referred to as ‘Mg-silicates’ in this study, are increasingly reported
associated with ancient and modern carbonates rocks (e.g. Hover et al., 1999; Bristow et al., 2009;
Wright, 2012; Zeyen et al., 2015, 2017; Pace et al., 2016; Gérard et al., 2018). In particular, Mg-
silicates have been reported in lacustrine carbonate reservoirs of the South-Atlantic (e.g. Bertani and
Carozzi, 1985a, 1985b; Rehim et al., 1986; Wright, 2012) and their dissolution during diagenesis was
proposed to explain the high porosity of these carbonate rocks (Wright and Barnett, 2015; Tosca and
This article is protected by copyright. All rights reserved.
Wright, 2015). These phases seem to form either through direct precipitation from water or during
very early diagenesis (e.g. Tosca, 2015) and record specific chemical conditions, especially high pH
values. Thus, they have a strong potential to help in palaeoenvironmental reconstruction of continental
carbonates deposits and their evolution through the diagenesis (Chamley, 1989; Weaver, 1989;
Bristow et al., 2009).
The variety of geological environments allowing Mg-rich clays to form is unclear. In
Phanerozoic rocks, the presence of Mg-silicates has been attributed to different depositional
environments including weathering profiles developed on Mg-rich rocks such as basic or ultrabasic
igneous rocks (Huang et al., 2013), hydrothermal settings where fluids interact with Mg-rich igneous
material (Setti et al., 2004), alkaline lakes and evaporative basins (Van Denburgh, 1975; Yuretich and
Cerling, 1983; Darragi and Tardy, 1987; Banfield et al., 1991; Hay et al., 1991; Hover et al., 1999;
Martini et al., 2002; Bristow et al., 2009; Zeyen et al., 2015, 2017; Pace et al., 2016; Gérard et al.,
2018). Spring inflows into lakes hosted in basic volcanic terrains could also contribute to create the
right set of conditions, i.e. high carbonate alkalinity, high Si, Mg and Ca concentrations, for the
development of lacustrine carbonates associated with Mg-silicates (Wright, 2012). The Doushantuo
Formation, which hosts fossil records of the early Ediacaran period (635 to 551 Ma), illustrates the
controversy about the formation of Mg-rich clays. In this carbonate formation, the presence of
saponite, a Mg-rich and Al-rich smectite, was interpreted as either an early diagenetic phase formed in
situ in an alkaline and non-marine environment (Bristow et al., 2009), or as a detrital phase resulting
from local to regional weathering of adjacent continental areas dominated by mafic to ultramafic
volcanic rocks (Huang et al., 2013). In this respect, the aluminum content of Mg-silicates may be of
key interest for such reconstructions because it has been proposed to provide insights into the nature
and the amounts of detrital silicate inputs in lacustrine environments (e.g. Millot, 1970).
Present-day sedimentary features can provide information about the palaeoenvironments
where Mg-rich clays formed. Sediments developed in lakes characterized by waters with high pH
(>9), and either relatively high silica concentrations (Van Denburgh, 1975; Yuretich and Cerling,
1983; Darragi and Tardy, 1987; Banfield et al., 1991; Hay et al., 1991; Zeyen et al., 2015, 2017;
Gérard et al., 2018) or high magnesium concentrations (Hover et al., 1999; Martini et al., 2002), were
found to host Mg-silicates. To date, however, the specific geochemical conditions controlling their
formation remain unclear. In particular, the role of volcanic catchment and spring inflow is uncertain,
limiting palaeoenvironmental and early diagenesis interpretations. To tackle this issue, the tropical
volcanic crater lake Dziani Dzaha in Mayotte (Indian Ocean) was chosen as a possible
This article is protected by copyright. All rights reserved.
palaeoenvironmental analogue for the formation model of Mg-rich silicates associated with
continental carbonates. The lake waters are known to be alkaline and the catchment is composed of
volcanic rocks of phonolitic/alkaline composition. Magnesium-silicate minerals were recently
identified in the stromatolites of the lake (Gérard et al., 2018), suggesting that they may also be
present in the soft carbonated sediments. The mineralogy of the first metre of the sediment column
was studied to identify possible occurrences of Mg-silicates and provide insights on the processes
leading to the formation of carbonates associated with Mg-silicates during early diagenesis.
GEOLOGICAL SETTING AND LIMNOLOGY OF THE DZIANI DZAHA
The Dziani Dzaha (which means volcanic crater lake in Mahorian) is a tropical volcanic crater lake on
the island of Petite Terre (Pamanzi) located within the Mayotte island complex in the Northern
Mozambique Channel. Mayotte is the most southerly and oldest island complex of the Comoros
Archipelago. It is made of two main volcanic islands; Grande Terre and Petite Terre (Fig. 1A and B).
The initial volcanism at the origin of the Mayotte complex island is estimated at about 10 to
15 Ma (Nougier et al., 1986); however, a late period of volcanic activity occurred during the last 2.5
Myr (Pelleter et al., 2014). The alkaline volcanism at the origin of the Comoros Archipelago was
suggested as deriving from the partial melt of lithospheric mantle metasomatized by CO2-rich fluids
beneath an ocean–continent transitional crust (Nougier et al., 1986; Mougenot et al., 1989; Bertil and
Reynoult, 1998; Coltorti et al., 1999; Pelleter et al., 2014). In the Mayotte island complex, the island
of Petite Terre was emplaced during the late stage of the volcanic activity. It is mostly made of
pyroclastic rocks of phonolite/alkaline composition resulting from the crystal fractionation of magmas
derived from the partial melting of an oceanic lithosphere enriched in CO2 (Pelleter et al., 2014).
The Dziani Dzaha is located in a volcanic crater formed by a phreatomagmatic eruption. The
pyroclastic rocks rest over a coral reef dated from 9 ka, which surrounds Grande Terre, the largest
island in the Mayotte island complex. As the most recent volcanic event is dated at 4 ka (Zinke et al.,
2003), the age of the Dziani Dzaha may range from 9 ka to 4 ka.
This article is protected by copyright. All rights reserved.
The lake (Dziani Dzaha) is a roughly elliptical shape with a length of 640 m and a width of
470 m. The lake watershed is very restricted. The volcanic crater is ca 1.2 km long and 50 m high on
average. The flanks consist of indurated layers of volcanic ashes and pumices. The lake bottom is
approximately at sea-level. The lake has an average depth of 3 m, with a maximum depth of 5 m,
except for a singular depression which reaches a depth of 18 m (Figs 1 and 2). This depression is
probably related to the phreatomagmatic eruption at the origin of the Dziani Dzaha. Magmatic
degassing is still active in the island of Petite Terre, as evidenced by the four bubbling areas of
volcanic CO2 identified at the lake surface. Microbialites, i.e. organo-sedimentary structures formed in
close association with micro-organisms, thrive in the shallow waters of the lake and are mainly
composed of aragonite and calcite (Gérard et al., 2018). The distribution of the microbialites is patchy
and mostly limited to the lake shore (Fig. 1) such that below a few tens of centimetres of water depth,
most of the sediments are composed of gelatinous and non-indurated material.
The physical, chemical and biological features of the lake are very unique (Leboulanger et al.,
2017). The salinity of the lake water reaches up to 52 psu (practical salinity unit), i.e. 1.5 times higher
than sea water. The alkalinity is ca 0.2 mol∙L-1, with pH values ranging from 9.1 to 9.4. The mean
daily surface temperature varies between 28°C and 36°C. The lake ecosystem is dominated by
prokaryotes with a dense and perennial bloom of filamentous cyanobacteria accounting for more than
90% of the primary producer biomass. Biological activity is significant all year round, with a primary
productivity of ca 8 gC/m2/day, close to the maximum for tropical and subtropical lakes proposed by
Lewis (2011).
MATERIALS AND METHODS
Nomenclature
An example of the scheme used for naming samples follows. For sample DZ14-10 C4, ‘DZ’ indicates
Dziani, ‘14-10’ indicates the year (2014) and the month (October) of the survey and C4 refers to the
sediment core number, which was 4 (see location in Fig. 1). As the sediment cores were numbered
according to their sampling order, they are named in Fig. 1 and in the text only as CX, with X the
number of the core from 1 to 13. Sediment cores C3, C5, C7, C8 and C12 were not dedicated to the
study of the sediment composition and are not used in this study, except for the water content of
sediment core C12. The samples are listed in Table S1 in Supporting Information.
This article is protected by copyright. All rights reserved.
Sampling, sample treatment and water content of sediment
Sediment cores were collected at different sites within the lake during the surveys performed in April
2012, April 2014, October 2014 and August 2016 (Fig. 1; Table S1). Data from nine sediment cores
are presented in this study: C1, C2, C4, C6, C9, C10, C11, C12 and C13. A Uwitec core sampler (90
mm diameter PVC tubes) (Uwitec, Mondsee, Austria) was used to collect sediment from the water–
sediment interface to a maximum depth of 1 m. The sediment cores were subsampled, as much as
possible, on a colour basis in the form of 3 to 5 cm thick slices, which were preserved under anoxic
conditions in a cold room. The samples were freeze-dried, before being rinsed with deionized water,
centrifuged three to four times, and crushed down to <80 µm. In the sediment core C12, the water
content was estimated for 29 samples by weighing fresh samples before and after drying at 60°C. The
precipitation of halite (NaCl) was corrected assuming salinity similar to that of the water column, i.e.
of 52 g∙kg-1 (Leboulanger et al., 2017). Due to the large number of samples, the analytical procedures
presented below could not be performed on all samples.
Total carbon content and carbon isotope composition
The total carbon (TC) content and carbon isotope composition were analysed on 102 samples of seven
sediment cores. The TC content was quantified with a carbon-hydrogen-nitrogen (CHN) elemental
analyzer (Shimadzu TOC V CSH; Shimadzu Corp, Kyoto, Japan) on bulk samples with a precision of
±10%. The isotopic composition of total carbon was analysed with Flash-EA1112 elemental analyzer
coupled to a Thermo Finnigan Deltaplus XP mass spectrometer via a Conflo IV interface (Thermo
Fisher Scientific, Waltham, MA, USA). Four organic laboratory standards were analysed for
concentration and isotopic calibration; VH1, CAP, MX33 and LC. Carbon isotopic compositions are
reported relative to the Vienna Pee Dee Belemnite (VPDB) standard with a precision of ±0.5‰. In
some samples, carbonates or organic matter were removed to measure the content and the isotopic
composition of the carbon end-members. Decarbonatation was performed on 45 samples with HCl
6N. The content of total organic carbon (TOC) and the carbon isotope composition of organic matter
were measured with the same instruments as total carbon. For most samples, significant loss occurred
during the centrifugation step of the decarbonatation process, which precluded accurate quantification
of the TOC content. The measured TOC content is only reported for sediment cores C6 and C9, for
which samples were freeze-dried before centrifugation, which allowed for the limitation of mass loss.
For 19 samples, organic matter was removed using low-temperature oxygen-plasma ashing in a
POLARON PT7160 RF system (Polaron Equipment Limited, Watford, UK) and the carbon isotope
composition of carbonates was measured with a GasBench coupled to a Thermo Finnigan Deltaplus XP
mass spectrometer (Thermo Fisher Scientific). Three laboratory carbonate standards were used for
This article is protected by copyright. All rights reserved.
both carbonate quantification and for isotopic calibration; Across, Merck and Rennes II (details in
Assayag et al, 2006; Lebeau et al., 2014). Isotopic compositions are reported relative to the VPDB
standard for carbon with a precision of ±0.1‰. All carbonate and organic laboratory standards for C-
isotope composition were calibrated independently against NBS19 and IAEA CO 1 international
standards.
X-ray powder diffraction and energy dispersive X-ray fluorescence analyses
X-ray powder diffraction (XRPD) analyses were performed on 128 samples of seven sediment cores
and 12 samples of the volcanic crater with an Empyrean (PANalytical, Almelo, The Netherlands)
diffractometer equipped with a multichannel PIXcel 3D detector (Malvern PANalytical, Malvern,
UK) and a Cu Kα X-ray source (λ = 1.541874 Å). Each pattern was recorded in the θ-θ Bragg-
Brentano geometry, in the 5° to 90° 2θ range (0.0131° for 70 sec). Analysis of X-ray patterns and
mineralogical identification was performed using the High Score Plus software (Malvern
PANalytical), the Crystallography Open Database (COD) and the Inorganic Crystal Structures
Database (ISCD). The reference intensity ratio (RIR) method was used for semi-quantifications of
mineral phases (Visser and de Wolff, 1964); however, reference intensity ratios for clay minerals
were frequently missing from the databases.
Elemental analyses were further conducted by energy dispersive X-ray fluorescence
(EDXRF) on 99 samples of six sediment cores and four samples of the volcanic catchment with an
Epsilon 3XL (PANalytical) spectrometer equipped with a Au X-ray tube. A 10 min measurement
repeated six times for each sample yielded accurate quantitative analysis.
Scanning electron microscopy and transmission electron microscopy analyses
Scanning and transmission electron microscopy were performed on crushed (<80 µm) samples after
freeze-drying and rinsing. Crushed samples were used for microscopic analyses because the
preservation of textural relationships among different components was rendered difficult due to the
gelatinous texture of the sediment (Description of lake sediment textures section). Scanning electron
microscopy (SEM) was performed on 19 samples with a Zeiss Ultra 55 FEG microscope (Carl Zeiss
AG, Oberkochen, Germany) operated at 10 kV and a working distance of 7.5 mm using a
backscattered electron detector for imaging. Transmission electron microscopy (TEM) was performed
with a JEOL FEG 2100F (JEOL Limited, Tokyo, Japan) operated at 200 kV on 1 clay-rich sample
This article is protected by copyright. All rights reserved.
selected on the basis of the XRPD analyses. Both SEM and TEM were equipped with an energy
dispersive X-ray spectroscopy (EDXS) system. The SEM was equipped with a Quantax EDS system
XFlash 4010 (Bruker, Billerica, MA, USA) with a silicon drift detector equipped with an ultrathin
polymer window. The TEM was equipped with a built-in JEOL EDS system with a Si(Li) detector
equipped with an ultrathin polymer window.
Clay fraction
Clay fraction analysis was performed on two clay-rich samples selected on the basis of the XRPD
analyses of bulk samples. Carbonates were removed using the acetic acid-acetate buffer method
(Ostrom, 1961). Organic matter was removed at 50°C by adding small aliquots of H2O2 to the
suspension until gaseous emission had ceased. The <2 µm fraction was separated using gravity
sedimentation of particles in water (Stoke’s law), and then saturated with a 1 M NaCl solution for 24
h at room temperature. Oriented preparations of the Na-saturated <2 µm fraction were carried out on
glass slides and dried at room temperature to obtain an air-dried (AD) preparation. Ethylene glycol
(EG) solvation was achieved by exposing the oriented slides to ethylene glycol vapour for 12 h.
Glycerol (GL) solvation was carried out by saturation of samples with a 2 M MgCl2 solution for 24 h
at room temperature and then exposure to glycerol for three days at 80°C. The air-dried oriented slides
were also heated at 500°C for 90 min and then solvated in EG for 16 h as described in detail by
Christidis and Koutsopoulou (2013). X-ray diffraction patterns were acquired on the oriented slides
over the 2° to 50° angular region, with 0.03° 2θ angular steps, using a Bruker D8 diffractometer
(Bruker, Billerica, MA, USA) with Cu Kα radiation (λ = 1.5418 Å). Counting time was adapted to the
low amount of sample available and was 1.65 sec per step; the two Soller slits were 2.5°. Analysis of
XRD patterns and mineralogical identification was performed using Diffracplus EVA software
(Bruker) and the ICDD database (International Center for Diffraction Data).
Electron microprobe analyses (EMPA) were conducted after coating of the <2 µm clay
fraction with carbon. Spot analyses were performed using a CAMECA SXFive electron microprobe
(CAMECA, Paris, France) with an accelerating voltage of 15 kV, a beam current of 12 nA, and a 1
µm beam diameter. More details on the analytical methods are found in Lerouge et al. (2017).
This article is protected by copyright. All rights reserved.
Cation exchange capacity (CEC) was measured using the hexamine cobalt(III) chloride
method developed by Hadi et al. (2013) and adapted from Orsini and Remy (1976). The CEC is
deduced from two cross analyses; the amount of hexamine cobalt(III) that disappears from the
solution (adsorbed on the clay), which is directly measured by spectrophotometry at 475 nm, and
analysis by ionic chromatography [high performance liquid chromatography (HPLC) Dionex
Corporation, Sunnyvale, CA, USA] of the major cations present in the supernatant (Na+, K+, Ca2+ and
Mg2+) which also indicates the relative populations of desorbed exchangeable cations.
Quantification of sediment composition
Based on XRPD and XRF analyses, bulk carbon contents and carbon isotope compositions,
calculations were performed to quantify the contents of the different mineral phases and of the organic
matter. The organic matter content was calculated in two steps. First, the content of total organic
carbon (TOC) was calculated using an isotope mass balance calculation. The total carbon content is a
mixture between organic and inorganic carbon, for which the isotopic compositions were measured in
this study. The contribution of organic carbon x in the total carbon content can be calculated with the
equation:
δ13CTC = x * δ13CTOC + (1-x) * δ13CTIC (1)
with δ13C of TC, TOC and TIC being the isotopic composition of total carbon, total organic carbon
and total inorganic carbon, respectively. Then the TOC content is given by:
TOC = x * TC (2)
Second, the content of organic matter was calculated considering a TOC to organic matter weight
ratio of 1.9 (Broadbent, 1953). The mineral content is deduced by simple mass balance. For the
sediment core C13, as the carbon content and the carbon isotope composition were not measured, the
content of organic matter is considered to be the same as in C6, which is a reasonable assumption
since the two cores were collected in the 18 m deep depression of the lake.
The content of each mineral phase was calculated using the XRF analyses, which requires
fixing the chemical formula of each phase. Empirical formulas were used for most minerals (Table 1).
For saponite, the structural formula has been determined in this study (Thermodynamic model
section). On the basis of SEM-EDS analyses of alkaline feldspars and volcanic pumices in lake
sediment samples, the empirical formula of anorthoclase is considered for these phases. Then, the
content of each mineral is calculated based on a given element.
This article is protected by copyright. All rights reserved.
Thermodynamic calculation
Mineral stability domains in the system Mg-Ca-C-O-H and Mg-Al-Si-O-H were calculated at 30°C
and 1 bar with the Thermoddem database (Blanc et al., 2012) and compared with lake water
compositions (Leboulanger et al., 2017). The speciation of dissolved species was calculated with the
Phreeqc software (Parkhurst and Appelo, 2013).
RESULTS
Description of lake sediment textures
Three main sediment textures have been observed within the lake (Fig. 2A). In the sediment cores
collected at water depths ranging from 1 to 5 m; which represents most of the lake area, the sediment
texture is highly gelatinous and characterized by horizontal laminae (Fig. 2B). The uppermost
sediment is markedly laminated with colours ranging from light to dark brown. The laminae are
mostly <1 mm thick but seem to be organized, on a colour basis, in laminae packages of few
centimetres in thickness, and were used when possible to describe the slices cut from the sediment
cores. In most cases, grain size was indistinguishable to the eye, except in rare mudstone layers which
contain milimetre-sized grains. Below ca 50 cm depth, the laminae tend to disappear and the sediment
changes to a more massive and darker texture. The water/rock weight ratio decreases from ca 33 in
surface sediments to ca 4 at 20 cm depth and ca 2 at 1 m depth (Fig. 2C). Thus, water is by far the
dominant phase of the sediment and most of the sediment compaction occurs in the first 20 cm below
the sediment surface. This sedimentary facies is very homogeneous and no lateral variation has been
observed in the area of 1 to 5 m of water depth.
The sediment cores C12 and C13 collected in the depression of the lake show a similar
sediment texture (highly gelatinous, laminated and with very fine grain size); however, deformation of
the laminae is observed below ca 30 cm deep (Fig. 2D). In the sediment cores close to the shoreline in
shallow (<1 m) water depth (C9 and C11), the sediment is greyish and grain size can increase up to
millimetres in size. In C9, the sediment texture is close to that of a mudstone and remains relatively
gelatinous and laminated. In C11, the sediment texture is close to a packstone and laminae are not
visible. The thickness of the sediment column is unknown; however, attempts to collect longer
sediment cores (C3, C4 and C5) around 4 to 5 m of water depth suggests that it is thicker than 2.3 m
in this area.
This article is protected by copyright. All rights reserved.
Total carbon content and isotopic composition of the lake sediments
The content of total carbon decreases from ca 20 wt.% in the uppermost sediment to ca 10 wt.% at
depth, except for the sediment cores collected in the depression of the lake (C6) and close to the
shoreline (C9 and C11) (Table S2). In the depression of the lake (C6), the TC content ranges between
20 wt.% and 25 wt.% and is relatively constant with depth. The TC content in C9 and C11 is lower
with values mostly below 10 wt.%. The sediment core C2 shows a local decrease of TC content down
to 3 wt.% at 7 to 14 cm depth. The TOC content measured in the sediment core C6 is high. Values
range mostly between 10 wt.% and 20 wt.% with the highest values in the uppermost sediment.
The isotopic composition of total carbon varies between ca -15‰ and ca +15‰ depending on
the sediment core (Table S2). The highest values are found in C9 and C11. C4 and C6 show the
lowest values. The variations of δ13C values of total carbon between the sediment cores reflect
variable proportions of organic and inorganic carbon. The isotopic compositions of the two carbon
end-members are relatively constant with depth, with average δ13C values of ca -14‰ for organic
carbon and ca +16‰ for inorganic carbon (Table S2). These isotopic compositions are used together
with the isotopic composition of total carbon to determine the TOC content using an isotopic balance
calculation (Quantification of sediment composition section). For sediment core C6, the calculated
and measured TOC content is perfectly consistent, which validates the isotope and mass balance
results (Table S2). Sediment cores C4, C6 and C10 show high TOC contents up to 24 wt.%, which
decrease with depth. The TOC content is lower in the sediment cores close to the shore line (C9 and
C11), with values mostly below 5 wt.%.
Chemical and mineralogical characterization of the sediment cores
Bulk analysis of the sediment cores (XRF)
The mineral content consists mostly of silicon, magnesium, calcium and aluminum (Table S3). Except
for the sediment cores collected at the shoreline (C11) and in the lake depression (C6 and C13), all
sediment cores (i.e. C4, C9 and C10) show a decrease in magnesium content with depth and a
concurrent increase of aluminum and silicon contents, while the calcium content remains relatively
stable. In sediment core C11, the evolution of the chemical composition with depth is approximately
the same, but the aluminum and silicon contents are higher than in C4, C9 and C10. In sediment cores
C6 and C13, the calcium, magnesium, aluminum and silicon contents are relatively stable with depth.
This article is protected by copyright. All rights reserved.
X-ray powder diffraction analyses of sediment cores
Aragonite, hydromagnesite and diopside or clinoenstatite, referred to hereafter as clinopyroxene, are
detected in the uppermost sediment (see Table S4). This mineral assemblage is relatively constant up
to ca 20 cm depth and occurs in all sediment cores collected under a water depth of >1 m (i.e. C2, C4,
C6, C10 and C13). Semi-quantifications with the RIR method indicate contents of aragonite,
hydromagnesite and clinopyroxene of 40 wt.%, 40 wt.% and 20 wt.%, respectively. Minor phases
such as magnetite, dolomite and quartz are sometimes identified with abundances of ca 5 to 10wt.%.
Halite is also occasionally detected but results from precipitation during freeze-drying. Some
sediment cores present a peculiar assemblage; high amounts of alkaline feldspars in C11 close to the
shoreline, calcite and alkaline feldspars between 7 cm and 14 cm depth in C2, magnesite in the
uppermost sediment of C9 and pyrite and dolomite at depth.
In addition to the aragonite–hydromagnesite–clinopyroxene assemblage, clay minerals occur
at depth, characterized by a diffraction peak close to 14.5 angstroms (Å), consistent with the basal
spacing (001) of smectites, chlorite or vermiculite (Fig. 3; the precise characterization of clay
minerals is presented later in the paper). Overall, clays were detected in all sediment cores below ca
10 to 15 cm, except for the sediment cores collected at the shoreline (core C11; 30 cm long) and in the
depression of the lake (cores C6 and C13), in which clays may be absent or below the XRD detection
limit. Concurrent to the occurrence of clays, the content of hydromagnesite decreases with depth until
complete disappearance in the sediment cores sampled between 2 cm and 5 m of water depth (C2, C4,
C9 and C10). In the other cores (C6, C11 and C13), the content of hydromagnesite remains relatively
constant with depth.
Scanning and transmission electron microscopy
Observations with a scanning electron microscope and chemical mapping show Mg-rich minerals
forming platelets, typical of hydromagnesite, together with Ca-rich minerals corresponding to
aragonite (Fig. 4A and B). Associated with the carbonates, a mineral phase rich in silicon, magnesium
and aluminum is recognized (Fig. 4A to C). Observations with transmission electron microscope of
minerals of similar chemical composition highlight the sheet structure of the phase (Fig. 4D and E).
Significant amounts of volcanic glass containing silicon, aluminum and lower amounts of sodium and
potassium, is also observed (Fig. 4F and G). The chemical composition is close to that of anorthoclase
(Na0.7K0.3AlSi3O8) (Fig. 4H). This is consistent with the XRF and XRPD analysis of volcanic pumices
and ash layers of the volcanic catchment (see Appendix S1, Fig. S1, Tables S5 and S6).
This article is protected by copyright. All rights reserved.
Characterization of the clay fraction
The <2 µm clay fraction was extracted from two samples: (i) one from 27 to 30 cm deep in sediment
core C4 which corresponds to a layer rich in clays (Table S4); and (ii) one from 67 to 72 cm deep in
sediment core C6. The XRD patterns acquired on oriented slides are similar for the two samples. It
may be observed that although the clays were Na-saturated, the 001 reflection was at ca 14.5 Å
instead of 12.5 Å as expected for relative humidity conditions normally occurring in the laboratory
(Brindley and Brown, 1980) (Fig. 5A and B). As illustrated further in the manuscript, CEC
measurement carried out on this Na-saturated fraction indicates that the exchanger still contains a
significant amount of magnesium, which therefore explains this value. The swelling of the (001) from
14.5 Å in the air-dried state towards 17 Å in the ethylene glycol state (Fig. 5B) indicates the presence
of smectite or vermiculite (Brindley and Brown, 1980). Powder XRD indicates that the (060) peak
position is 1.53 Å (Fig. 5A), which is characteristic of trioctahedral phyllosilicates (Moore and
Reynolds, 1997). Based on EPMA analysis (for example, no Li and Zn that ruled out hectorite and
sauconite, presence of Al that ruled out stevensite) and the fact that the phyllosilicate is trioctahedral,
the remaining possibilities are saponite and vermiculite (Brindley and Brown, 1980; Debure et al.
2016). Distinguishing between smectite and vermiculite is not easy. However, glycerol solvation on
Mg-saturated samples, led to swelling with a basal reflection at 16.5 Å (Fig. 5C). This is consistent
with the presence of saponite as vermiculite is not supposed to swell under these conditions (Walker,
1958; Brindley, 1966; Suquet et al., 1975; Tosca et al., 2011). The CEC of ca 87.1 meq/100 g is also
consistent with saponite, whereas values of ca 120 meq/100 g are reported for vermiculite (Brindley
and Brown, 1980). In addition, the fact that when heated to 500°C, the sample still showed a 001
reflection expanding to 17.4 Å after ethylene glycol solvation (Fig. 5B) further supports the saponite
hypothesis (Christidis and Koutsopoulou, 2013).
Cation population obtained from CEC measurements is consistent with EPMA results (Table
2). The two methods led to similar Na+ content values, which is the main interlayer cation. Ca2+ and
K+ values are a bit higher by EPMA because part of K+ present in the inner sphere complexes might
be difficult to extract with the hexamine cobalt method (Hadi et al., 2013). Magnesium content
obtained by EPMA was ten times higher than the value obtained by CEC. The structural Mg content
was deduced from the difference between EPMA and CEC measurements.
The mean structural formula was determined assuming that Al was preferentially tetrahedral,
the other cations being octahedral, and that Na, Ca and K were preferentially in the interlayers
(Lerouge et al., 2017). The calculation revealed that Fe was not in tetrahedral sites but in the
This article is protected by copyright. All rights reserved.
octahedral ones. In order to equilibrate the charge, Fe was only in its ferric form (Fe3+). Based on all
of those considerations, the mean structural formula is
K0.01Mg0.12Ca0.02Na0.28(Mg2.22Fe0.18Al0.28□0.32)(Si3.61Al0.39)O10(OH)2 where □ is a vacant site. Due to its
trioctahedral character, this formula is not in agreement with the clay mineralogy nomenclature
defined by Guggenheim (2006) for saponite; however, it is consistent with literature data that reported
vacancies (from 0.15 to 0.56 depending of the studies) in the octahedral position of ferric saponite
(Post, 1984; Kodama et al., 1988; Vincente et al., 1996; Hicks et al., 2014). In addition, the calculated
charge per formula unit (0.57) is below 0.6 which is considered in the nomenclature as the boundary
between saponite and vermiculite. The differences observed in the structural formula could be related
to the occurrence of minor amounts of other clay minerals or minor non-clay impurities (amorphous
and/or crystalline phases) in the samples that have not been identified. The terminology ‘saponite-like
mineral’ will be used in this paper to be consistent with the nomenclature and the characterization.
Quantification of the sediment content
Results of the calculation of the evolution of mineral composition and organic matter content with
depth are shown in Fig. 6 (Table S7). Apart for the sediment core C11, organic matter represents more
than 25 wt.% of the uppermost sediment, with a maximum of 45 wt.% for sediment core C6. In
sediment cores C4 and C10, hydromagnesite decreases with depth while the saponite-like mineral
accumulates to reach up to 33 wt.% at ca 70 cm deep in C10. Concurrently, the content of organic
matter decreases. The content of aragonite is relatively constant around 25 wt.% whereas the
proportion of primary silicates, dominated by alkaline feldspar, increases with depth to reach up to ca
25 wt.%. The sediment core C11 shows much higher amounts of alkaline feldspars than the others
sediment cores, up to 60 wt.%. Compared with C4 and C10, the sediment cores sampled in the
depression of the lake (i.e. C6 and C13) have lower amounts of saponite-like mineral at depth and
relatively stable contents of hydromagnesite.
The calculation should fulfil an internal consistency such that the elemental abundances
corresponding to the calculated mineral contents should equal the measured elemental abundances.
For instance, the sum of silicon content in alkaline feldspar and in the saponite-like mineral, which is
calculated independently using the measured content of potassium and aluminum, respectively, should
not exceed the total measured content of silicon. For sediment cores C4, C6, C10 and C13, the
calculated silicon content exceeded by 10 mol % the measured silicon content for only ca 10% of
samples. These low errors validate the calculations for these sediment cores. In contrast, the
calculation for C9 is not shown due to a lack of internal consistency. Core C9 was collected close to
This article is protected by copyright. All rights reserved.
the shoreline, where detrital influence is high. In this case, considering only alkaline feldspar and
pyroxene as detrital inputs might be an oversimplification.
Sensitivity tests have been performed on several parameters including the TOC to organic
matter weight ratio (±0.1 unit), the empirical formula of clinopyroxene (enstatite or diopside) and the
chemical composition of the saponite-like mineral (±0.1 mol/mol of mineral on the Al, Si and Mg
content). Most of these parameters modify the percentage content of each phase by less than 10%,
except for aragonite (up to 20%).
Thermodynamic model
The chemical compositions of lake waters (Table S8) (Leboulanger et al., 2017; Gerard et al., 2018)
are compared to the stability domains of carbonate and brucite in the system Mg-Ca-C-O-H (Fig. 7A).
The pore waters are close to metastable equilibrium between aragonite and hydromagnesite, and
supersaturated relative to dolomite.
The stability domains of Mg-silicates are represented in the system Mg-Al-Si-O-H,
considering that aluminum behaves conservatively between the Al-bearing phases (Fig. 7B). Chlorites
are not considered in the calculation as their formations are kinetically limited at ambient temperature
and pressure (Meunier, 2005). High pH values together with high activities of Mg2+ and dissolved
silica favour the stability of saponite compared to other aluminosilicates. The chemical composition
of lake water is close to equilibrium with quartz and supersaturated relative to saponite, kerolite, talc
and sepiolite.
DISCUSSION
Sediment composition and distribution
Three main sedimentary facies were observed within the lake, which respond to a shore to basin
central lake transect. Close to the shoreline, the greyish facies dominated by millimetre-size grains is
relatively poor in organic matter and rich in primary silicates, alkaline feldspars and pyroxene, which
highlights high detrital input from the volcanic catchment. In the area of the lake, 1 to 5 m in depth,
the gelatinous facies is dominated by organic matter and indicates low lateral variation. The detrital
components decrease with distance from the lake shore. The laminated uppermost sediment is mostly
This article is protected by copyright. All rights reserved.
formed of aragonite and hydromagnesite whereas, at depth, the sediment texture becomes massive and
dark, and the hydromagnesite is replaced by the saponite-like mineral. The layers containing coarser
grains and enriched in calcite and alkaline feldspar with less organic carbon (sediment core C2)
reflects a local higher energy detrital input to the lake margin. In the central lake depression, soft-
sediment deformation is interpreted as slump events (Alsop and Marco, 2011). In this area, the
apparently undeformed uppermost sediment is also rich in organic matter, aragonite and
hydromagnesite but hydromagnesite persists at depth and the saponite-like mineral occurs in much
lower amounts. No clay mineral has been detected in the uppermost part of sediment cores suggesting
low amounts of detrital clays inherited from the weathering of the surrounding volcanic rocks.
Overall, the sediment deposited below 1 m of water depth is singular by its extremely high
content of organic matter (up to 45 wt.%), a consequence of intense primary productivity in the lake,
and by the anomalously positive carbon isotope signature of both carbonates and organic matter,
which probably resulted from a combination of volcanic CO2 input and methanogenesis. The
sediment collected between 1 m and 5 m of water depth is the most representative of the sedimentary
deposits observed within the lake. It is used hereafter as a case study to develop a better understanding
of the co-occurrences of carbonates and Mg-silicates.
Thermodynamic and kinetic controls on carbonates and saponite-like mineral formation
Lake water composition is close to equilibrium with aragonite and hydromagnesite and the presence
of these two carbonates in the uppermost sediment indicates that they precipitate relatively rapidly
from lake water, either in the water column or at the sediment–water interface. The occurrence of
aragonite is consistent with a Mg/Ca molar ratio of >10 of the lake water, which has been shown to
favour the formation of aragonite rather than calcite (Simkiss, 1964; Taft, 1967; Debure et al., 2013a
and b; Zeyen et al., 2017) Although thermodynamics suggests that magnesite and dolomite are the
most stable Mg-bearing carbonates at surface temperature and pressure, kinetics exerts a strong
primary control on their formation (e.g. Sayles, 1973; Arvidson and MacKenzie, 1999; Gautier et al.,
2014), which may explain why they are rarely observed.
Despite supersaturation of the lake water relative to saponite, kerolite, talc and sepiolite, none
of these phases were detected in the mineral assemblage of the uppermost sediments, suggesting that
they do not precipitate from the lake water. Precipitation of Al-free Mg-silicates from lake water is
nonetheless possible, as exemplified by the magnesium depletion in Lake Turkana attributed to the
This article is protected by copyright. All rights reserved.
precipitation of stevensite (Yuretich and Cerling, 1983; Cerling, 1996). Tosca (2015) proposed a
critical supersaturation allowing Al-free Mg-silicates to overcome kinetic barriers and to precipitate
directly from the water column. In the Dziani Dzaha, the removal of magnesium by hydromagnesite
precipitation likely hinders the achievement of this critical supersaturation, preventing the
homogeneous nucleation of Al-free Mg-silicates in the water column. Regarding the Mg-
aluminosilicates, they are expected to form mostly through heterogeneous nucleation (e.g. Tosca,
2015; Tosca and Wright, 2015), because of the low solubility of aluminum in most cases and
especially in alkaline lakes (2.2 10-6 mol L-1 at pH 9). In this process, detrital Al-bearing silicates act
as mineral precursors, lowering the supersaturation required for the formation of aluminosilicates. In
the Dziani Dzaha, alkaline feldspars and volcanic glasses are the likely Al-bearing precursors,
allowing the formation of the saponite-like mineral in the sediment.
Achievement of high pH values in the water column is one of the chief factors that enable
supersaturation relative to saponite, aragonite and hydromagnesite. Consistently, Mg-silicates
associated with carbonate sediments have been widely reported in alkaline lakes and evaporative
basins of pH values >9 (Van Denburgh, 1975; Yuretich and Cerling, 1983; Darragi and Tardy, 1987;
Banfield et al., 1991; Hay et al., 1991; Hover et al., 1999; Martini et al., 2002; Bristow et al., 2009).
As for other alkaline and hypereutrophic lakes (Lòpez-Archilla et al., 2004), the intense primary
productivity in the Dziani Dzaha is the most likely mechanism responsible for these pH values,
according to the simplified reaction:
HCO3- + H2O = “CH2O” + OH- + O2 (3)
The development of massive primary productivity could be linked to the input of volcanic CO2 within
the lake, which probably supplies primary productivity with carbon.
In addition to pH, magnesium and silica concentrations are critical for the formation of
saponite. The magnesium concentration of ca 4 mM is comparable to the ones of other alkaline lakes
producing Mg-silicates (Van Denburgh, 1975; Yuretich and Cerling, 1983; Darragi and Tardy, 1987;
Banfield et al., 1991). However, due to the competition for magnesium between Mg-silicates and Mg-
carbonates, silica might have a greater influence than magnesium on the formation of saponite (Zeyen
et al., 2017). The silica concentrations (ca 0.3 mM) are around two orders of magnitude higher than
the usual concentration of surface oceans (<2 µM; Treguer et al., 1995), where the efficiency of
diatoms in precipitating silica drastically lowers the concentration (e.g. Siever, 1992; Racki and
Cordey, 2000). Because no spring inflows have been identified, these high silica concentrations are
attributed to the weathering of volcanic terrains in a diatom-poor lacustrine environment.
This article is protected by copyright. All rights reserved.
Early diagenetic evolution of sediment composition
The evolution of the sediment composition with depth is the likely result of early diagenetic reactions.
The appearance of the saponite-like mineral only at depth, in spite of its supersaturation in the lake
water, can be explained by the intrinsic kinetics limit of saponite precipitation (e.g. Tosca, 2015) and
by the kinetics of dissolution of the Al-bearing detrital phase. The mineralization of organic matter is
evidenced by the decrease of TOC content with depth. In the right proportions, the CO2 produced by
the mineralization of organic matter could account for a decrease of pH destabilizing hydromagnesite
while allowing both aragonite and the saponite-like mineral to remain supersaturated. In addition, a
genetic relationship may exist between hydromagnesite and the saponite-like mineral and affect the
pH. Calculation of mineral stability domains (Fig. 7B) shows that the precipitation of saponite from
supersaturated water tends to decrease the pH values, favouring the destabilization of hydromagnesite,
which, in return, could supply with Mg the formation of saponite (Fig. 8). In addition to organic
matter mineralization, CO2-rich volcanic gases seeping from the volcanic basement could contribute
to the pH decrease of pore waters.
Changes of palaeoenvironmental conditions throughout the lake history are less consistent
than early diagenetic reactions to explain the evolution of the sediment composition. If the decreasing
amount of hydromagnesite with depth was due to lower Mg concentrations in the palaeo-lake waters
and/or lower pH values than currently observed, saponite supersaturation would also be lower, which
is not compatible with the increasing amount of saponite-like mineral at depth. The diagenetic
scenario is thus the most likely, even though it remains to be validated. This would require the
assessment of the saturation state of pore waters and the quantitative evaluation, with reactive
transport modelling, of the diagenetic reactions occurring in the sediment, including saponite
precipitation, hydromagnesite destabilization and organic matter mineralization.
Implications for the formation and distribution of Mg-silicates of the South Atlantic rift basins
Magnesium-silicates (stevensite, kerolite and talc) have been identified in lacustrine Cretaceous
carbonate reservoir rocks of the South Atlantic (Bertani and Carozzi, 1985a, 1985b; Rehim et al.,
1986; Wright, 2012; Tosca and Wright, 2015); their formation was suggested to occur in lakes where
volcanic terrains predominated (Cerling, 1994; Wright, 2012). However, the source of Si and Mg to
account for the amount of Mg-silicate found in those rocks remains under discussion. Wright (2012)
suggested that spring inflows, in addition to the weathering of volcanic terrains, might have been
required to enable the formation of both carbonate and Mg-silicates.
This article is protected by copyright. All rights reserved.
The study case of the Dziani Dzaha suggests that high silica and magnesium concentrations
can be reached in lakes through the weathering of volcanic alkaline terrains, without necessarily
involving external inputs such as spring or hydrothermal inflows. The development of an intense
primary productivity, sustained both by the magmatic CO2 and by the nutrients supplied by the
volcanic terrains (such as iron, potassium and phosphorous), is most likely to be responsible for the
high pH values. Thus, a genetic link may exist between volcanic alkaline terrains, organic-rich
sediment and Mg-silicates. In the Doushantuo Formation (635 to 551 Ma), saponite is associated with
TOC contents reaching up to 4 wt.% (Bristow et al., 2009). Using a carbon to organic matter ratio of
1.9 (Broadbent, 1953) and considering 66 wt.% loss of initial organic carbon by thermal maturation
(Tissot and Welte, 1984), the initial content of organic matter in the Doushantuo sediments may have
been of ca 23 wt.%, which is comparable to the Dziani Dzaha sediments. The deposition of lacustrine
carbonates containing more than 20 wt.% of organic matter could have required a significant source
of carbon in addition to the atmospheric CO2. In the Dziani Dzaha, the volcanic CO2 may have played
a major role by fuelling both the primary productivity and carbonate formation. Similarly, the
Cretaceous lakes at the origin of the South Atlantic carbonate reservoir rocks could have been
influenced by CO2-rich magmatic inflows during the continental rifting, as the continental crust
becomes thinner and the influence of mantle fluids increases.
The absence of Al-free Mg-silicates in the Dziani Dzaha sediment, whereas they prevail in the
carbonate reservoirs of the South Atlantic, supports two mutually compatible hypotheses as to the
factors controlling the presence or absence of Al in Mg-silicates. At the local scale, this supports the
‘clay-zoning’ scenario of Millot et al. (1970), who suggested that the presence of detrital Al-bearing
silicates at the basin margin allows Al-rich clay minerals to form, whereas, towards the basin centre,
lower amounts of detrital materials indirectly promote the homogeneous nucleation of Al-free Mg-
silicates (Weaver and Beck, 1977; Jones and Weir, 1983; Jones, 1986; Calvo et al., 1999; Deocampo
et al., 2009; Galán and Pozo, 2011). Accordingly, the occurrence of a saponite-like mineral in the
Dziani Dzaha is consistent with the presence of detrital component in the sediments and the
prevalence of Al-free Mg-silicates in the South Atlantic carbonate reservoirs would rather indicate
large lake systems with open water lake deposits free from detrital inputs.
At the regional scale, the formation of saponite or Al-free Mg-silicates may also reflect the
nature of the volcanic terrains hosting the lacustrine environments. At the early stage of continental
rifting, low mantle uplift results in low partial melting, producing alkaline magma (e.g. White and
McKenzie, 1989; Wilson, 1992). Below the thick continental lithosphere, the slow cooling of magmas
generates highly differentiated basalts enriched in alkaline elements, aluminum and silicon, i.e. a
This article is protected by copyright. All rights reserved.
phonolite/alkaline composition comparable to the one of magmas at the origin of the Mayotte Islands.
In contrast, during later stages of continental rifts, higher partial melting and lower magmatic
differentiation produce tholeiitic basalts with lower aluminum content. The relatively low availability
of aluminum in the Cretaceous lakes at the origin of the South Atlantic carbonate reservoirs may
indicate lacustrine environments formed during the late stage of continental rifting. This is consistent
with the suggestion of Tosca and Wright (2015) that the water chemistry of these lakes may have been
influenced by the serpentinization of exhumed mantle at the transition between continental and
oceanic crusts.
CONCLUSIONS
The Dziani Dzaha sediment exhibits a particular evolution of composition in the first metre of
sediment, which can be explained by early diagenetic reactions within the lake sediments; these are
mainly gelatinous and exhibit homogeneous facies distribution with local increases in detrital inputs
close to the shore. Mineralization of organic matter within the sediment, possibly associated with
inputs of volcanic CO2, can explain the destabilization of hydromagnesite and the decrease of organic
matter content with depth, while aragonite remains stable. Kinetic limitation prevents the formation of
the saponite-like mineral in the uppermost sediment, such that it only accumulates below 15 cm depth.
Reactive transport modelling is now needed to allow a quantitative evaluation of the role of all
diagenetic processes occurring in these sediments: saponite precipitation, hydromagnesite
destabilization, organic matter mineralization and CO2-rich gas inputs.
The formation of the saponite-like mineral in the Dziani Dzaha results from the high silica
activities and the high pH values of the lake waters. This water chemistry is attributed to the delivery
in a restricted lake system of chemical elements, among which nutrients, produced by the weathering
of volcanic terrains, which favours the development of an intense primary productivity. The bubbling
of magmatic gases in the Dzaini Dzaha suggests that magmatic CO2 may fuel the primary productivity
and the carbonate precipitation. Thus, this study suggests that a genetic link may exist between
volcanic contexts and carbonate sediments rich in organic matter and Mg-silicates, which may be key
in understanding their formation throughout the fossil record. As an example, the migration through
the crust of mantle-derived fluids during continental rifting could have influenced the formation of the
Cretaceous South Atlantic carbonate reservoir rocks. In these rocks, the formation of Al-free Mg-
silicates rather than saponite may be indicative of low aluminum availability, favoured in large
lacustrine environments hosted in poorly differentiated volcanic terrains commonly emplaced during
late stages of continental rifting.
This article is protected by copyright. All rights reserved.
ACKNOWLEDGEMENTS
This work was supported by IPGP, TOTAL (project FR00008189), Agence Nationale de la Recherche
(France) (ANR DZIANI, grant number ANR-13-BS06-0001), Total Corporate Foundation (project
755 DZAHA) and one INSU-INTERRVIE grant (grant number AO2013-785992). The Deep Carbon
Observatory community is thanked here for several informative discussions on the importance of
studying environments related to magmatic CO2 emissions. The authors also wish to thank their
colleagues (C. Leboulanger, C. Bernard, P. Got, E. Fouilland, M. Bouvy, E. Le Floch, V. Grossi and
D. Sala) for their support and assistance during sampling campaigns on Dziani Dzaha. Last but not
least, they thank the Air Austral Airline Company and Alexandra and Laurent at the ‘Les Couleurs’
Guest House in Mayotte for their valuable assistance and support.
REFERENCES
Alsop, G.I. and Marco, S. (2011) Soft-sediment deformation within seismogenic
slumps of the Dead Sea Basin. Journal of Structural Geology, 33, 433-457.
Arvidson, R.S. and Mackenzie, F.T. (1999) The dolomite problem; control of
precipitation kinetics by temperature and saturation state. American Journal of Science, 299,
257-288.
Assayag, N., Rivé, K., Ader, M., Jézéquel, D. and Agrinier, P. (2006) Improved
method for isotopic and quantitative analysis of dissolved inorganic carbon in natural water
samples. Rapid Communications in Mass Spectrometry, 20, 2243-2251.
Banfield, J.F., Jones, B.F. and Veblen, D.R. (1991) An AEM-TEM study of
weathering and diagenesis, Abert Lake, Oregon, II. Diagenetic modification of the
sedimentary assemblage. Geochimica and Cosmochimica Acta, 55, 2795–2810.
Bertani, R.T. and Carozzi, A.V. (1985a) Lagoa feia formation (Lower Cretaceous),
Campos basin, offshore Brazil: rift valley stage lacustrine carbonate reservoirs-I. Journal of
Petroleum Geology, 8, 37-58.
Bertani, R.T. and Carozzi, A.V. (1985b) Lagoa feia formation (Lower Cretaceous),
Campos basin, offshore Brazil: rift valley type lacustrine carbonate reservoirs-II. Journal of
Petroleum Geology, 8, 199-220.
Bertil, D. and Regnoult, J.M. (1998) Seismotectonics of Madagascar. Tectonophysics,
294, 57-74.
Blanc, P., Lassin, A., Piantone, P., Azaroual, M., Jacquemet, N., Fabbri, A. and
This article is protected by copyright. All rights reserved.
Gaucher, E.C. (2012) Thermoddem: A geochemical database focused on low temperature
water/rock interactions and waste materials. Applied Geochemistry, 27(10), 2107-2116.
Brindley, G.W. (1966) Ethylene Glycol and Glycerol Complexes of Smectites and
Vermiculites. Clay Minerals 6, 237-259.
Brindley, G.W. and Brown, G. (1980) Crystal Structures of Clay Minerals and their
X-Ray Identification. Mineralogical Society Monograph 5, Mineralogical Society of Great
Britain and Ireland, London, 495 pp.
Bristow, T.F., Kennedy, M.J., Derkowski, A., Droser, M.L., Jiang, G.Q. and Creaser,
R.A. (2009) Mineralogical constraints on the palaeoenvironments of the Ediacaran
Doushantuo Formation. Proceedings of the National Academy of Sciences of the USA, 106,
13190–13195.
Broadbent, F.E. (1953) The soil organic fraction. Advances in Agronomy, 5, 153-183.
Calvo, J.P., Blanc-Valleron, M.M., Rodriguez-Arandia, J.P., Rouchy, J.M. and Sanz,
M.E. (1999) Authigenic clay minerals in continental evaporitic environments. In:
Palaeoweathering, Palaeosurfaces and Related Continental Deposits (Eds M. Thiry and R.
Simon-Coinçon R), International Association of Sedimentologists, Special Publications, 27,
129–151.
Cerling, T.E. (1996) Pore water chemistry of an alkaline lake: Lake Turkana, Kenya.
In: Limnology, Climatology and Paleoclimatology of the East African Lakes (Eds T.C.
Johnson and E.O. Odada), pp. 225–240. Gordon and Breach, Amsterdam,.
Chamley, H. (1989) Clay Sedimentology. Springer,-Verlag, Berlin, Heidelberg, 623
pp.
Christidis, G.E. and Koutsopoulou, E. (2013) A simple approach to the identification
of trioctahedral smectites by X-ray diffraction. Clay Minerals, 48, 687-696.
Coltorti, M., Bonadiman, C., Hinton, R.W., Siena, F. and Upton, B.G.J. (1999)
Carbonatite metasomatism of the oceanic upper mantle: evidence from clinopyroxenes and
glasses in ultramafic xenoliths of Grande Comore, Indian Ocean. Journal of Petrology, 40,
133-165.
Darragi, F. and Tardy, Y. (1987) Authigenic trioctahedral smectites controlling pH,
alkalinity, silica and magnesium concentrations in alkaline lakes. Chemical Geology, 63, 59–
72.
Debure, M., De Windt, L., Frugier, P. and Gin, S. (2013a) HLW glass dissolution in
the presence of magnesium carbonate: Diffusion cell experiment and coupled modeling of
diffusion and geochemical interactions. Journal of Nuclear Materials, 443, 507-521.
Debure, M., Frugier, P., De Windt, L. and Gin, S. (2013b) Dolomite effect on
borosilicate glass alteration. Applied Geochemistry, 33, 237-251.
This article is protected by copyright. All rights reserved.
Debure, M., De Windt, L., Frugier, P., Gin, S. and Vieillard, P. (2016) Mineralogy
and thermodynamic properties of magnesium phyllosilicates formed during the alteration of a
simplified nuclear glass. Journal of Nuclear Materials, 475, 255-265.
Deocampo, D.M., Cuadros, J., Wing-Dudek, T., Olives, J. and Amouric, M. (2009)
Saline lake diagenesis as revealed by coupled mineralogy and geochemistry of multiple
ultrafine clay phases: Pliocene Olduvai Gorge, Tanzania. American Journal of Science, 309,
834–858.
Galán, E. and Pozo, M. (2011) Palygorskite and sepiolite deposits in continental
environments. Description, genetic patterns and sedimentary settings. In: Developments in
Palygorskite–Sepiolite Research (Eds E. Galán and A. Singer), pp. 125-173. Elsevier,
Amsterdam.
Gautier, Q., Bénézeth, P., Mavromatis, V. and Schott, J. (2014) Hydromagnesite
solubility product and growth kinetics in aqueous solution from 25 to 75°C. Geochimica et
Cosmochimica Acta, 138, 1-20.
Gérard, E., De Goeyse, S., Hugoni, M., Agogue, H., Richard, L., Milesi, V., Guyot,
F., Lecourt, L., Borensztajn, S., Leclerc, T., Joseph, M., Sarazin, G., Jezequel, D.,
Leboulanger, C. and Ader, M. (2018) Key role of Alphaproteobacteria and Cyanobacteria in
the formation of stromatolites of Lake Dziani Dzaha (Mayotte, Western Indian Ocean).
Front. Microbiol. 9, 796. doi: 10.3389/fmicb.2018.00796.
Guggenheim, S., Adams, J.M., Bain, D.C., Bergaya, F., Brigatti, M.F., Drits, V.A.,
Formoso, M.L.L., Galán, E., Kogure, T. and Stanjek, H. (2006). Summary of
recommendations of nomenclature committees relevant to clay mineralogy: Report of the
Association Internationale pour l’Etude des Argiles (AIPEA) Nomenclature Committee for
2006. Clay Minerals, 41, 863-877.
Hadi, J., Tournassat, C., Ignatiadis, I., Greneche, J.M. and Charlet, L. (2013)
Modelling CEC variations versus structural iron reduction levels in dioctahedral smectites.
Existing approaches, new data and model refinements. Journal of Colloid and Interface
Science, 407, 397-409.
Hay, R.L., Guldman, S.G., Matthews, J.C., Lander, R.H., Duffin, M.E. and Kyser,
T.K. (1991) Clay mineral diagenesis in core KM-3 of Searles Lake, California. Clays and
Clay Minerals, 39, 84-96.
Hicks, L.J., Bridges, J.C. and Gurman, S.J. (2014) Ferric saponite and serpentine in
the nakhlite martian meteorites. Geochim. Cosmochim. Ac., 136, 194-210.
Hover, V.C., Walter, L.M., Peacor, D.R. and Martini, A.M. (1999) Mg-smectite
authigenesis in a marine evaporative environment, Salina Ometepec, Baja California. Clays
and Clay Minerals, 47, 252–268.
This article is protected by copyright. All rights reserved.
Huang, J., Chu, X., Lyons, T.W., Planavsky, N.J. and Wen, H. (2013) A new look at
saponite formation and its implications for early animal records in the Ediacaran of South
China. Geobiology, 11, 3-14.
Jones, B.F. (1986) Clay mineral diagenesis in lacustrine sediments. In: Studies in
Diagenesis (Ed. F.A. Mumpton), United States Geological Survey Bulletin, 1578, 291–300.
Jones, B.F. and Weir, A.H. (1983) Clay minerals of Lake Abert, an alkaline, saline
lake. Clays and Clay Minerals, 31, 161-172.
Kodama, H., De Kimpe, C.R. and Dejou, J. (1988) Ferrian saponite in a gabbro
saprolite at Mont Mégantic, Quebec. Clays Clay Miner., 36, 102-110.
Lebeau, O., Busigny, V., Chaduteau, C. and Ader, M. (2014) Organic matter removal
for the analysis of carbon and oxygen isotope compositions of siderite. Chemical Geology,
372, 54-61.
Leboulanger, C., Agogué, H., Bernard, C., Bouvy, M., Carré, C., Cellamare, M.,
Duval, C., Fouilland, E,m Got, P., Intertaglia, L., Lavergne, C., Le Floc’h, E., Roques, C. and
Sarazin, G. (2017) Microbial Diversity and Cyanobacterial Production in Dziani Dzaha
Crater Lake, a Unique Tropical Thalassohaline Environment. PLoS ONE 12, e0168879.
doi.org/10.1371/journal.pone.0168879
Lerouge, C., Gaboreau, S., Grangeon, S., Claret, F., Warmont, F., Jenni, A., Cloet, V.
and Mäder, U. (2017) In situ interactions between Opalinus Clay and Low Alkali Concrete,
Physics and Chemistry of the Earth, 99, 3-21.
Lewis Jr., W.M. (2011) Global primary production of lakes: 19th Baldi Memorial
Lecture. Inland Waters, 1, 1-28.
Lòpez-Archilla, A.I., Moreira, D., López-García, P. and Guerrero, C. (2004)
Phytoplankton diversity and cyanobacterial dominance in a hypereutrophic shallow lake with
biologically produced alkaline pH. Extremophiles, 8, 109-115.
Martini, A.M., Walter, L.M., Lyons, T.W., Hover, V.C. and Hansen, J. (2002)
Significance of early diagenetic water-rock interactions in a modern marine
siliciclastic/evaporite environment: Salina Ometepec, Baja California. Geol. Soc. Am. Bull.
114, 1055–1069.
Meunier A. (2005) Clays. Springer-Verlag, Berlin, Heidelberg, 472 pp.
Millot G. (1970) Geology of Clays. Springer-Verlag, Berlin, Heidelberg, 430 pp.
Mougenot, D., Hernandez, J. and Virlogeux, P. (1989) Structure et volcanisme d'un
rift sous-marin; le fosse des Kerimbas (marge nord-mozambique). Bulletin de la Société
Géologique de France, 2, 401-410.
This article is protected by copyright. All rights reserved.
Moore, D.M. and Reynolds, R.C. (1989) X-ray Diffraction and the Identification and
Analysis of Clay Minerals. Oxford University Press, Oxford, 332 pp.
Nougier, J., Cantagrel, J.M. and Karche, J.P. (1986) The Comores archipelago in the
western Indian Ocean: volcanology, geochronology and geodynamic setting. Journal of
African Earth Sciences, 5, 135-145.
Orsini, L. and Remy, J.C. (1976) Utilisation du chlorure de cobaltihexamine pour la
détermination simultanée de la capacité d'échange et des bases échangeables des sols. Sci.
Sol, 4, 269-275.
Ostrom, M.E. (1961) Separation of clay minerals from carbonate rocks by using acid.
Journal of Sedimentary Research, 31, 123-129.
Pace, A., Bourillot, R., Bouton, A., Vennin, E., Galaup, S., Bundeleva, I., Patrier, P.,
Dupraz, C., Thomazo, C., Sansjofre, P., Yokoyama, Y., Franceschi, M., Anguy, Y., Pigot, L.,
Virgone, A. and Visscher, P.T. (2016) Microbial and diagenetic steps leading to the
mineralisation of Great Salt Lake microbialites. Scientific Reports, 6:31495. doi:
10.1038/srep31495
Parkhurst, D.L. and Appelo, C.A.J. (2013) Description of input and examples for
PHREEQC version 3: a computer program for speciation, batch-reaction, one-dimensional
transport, and inverse geochemical calculations. U.S. Geological Survey Techniques and
Methods, Book 6, A43, 497 pp. Available at https://pubs.usgs.gov/tm/06/a43
Pelleter, A.A, Caro, M., Cordier, C., Bachélery, P., Nehlig, P., Debeuf, D. and
Arnaud, N. (2014) Melilite-bearing lavas in Mayotte (France): An insight into the mantle
source below the Comores. Lithos, 208-209, 281-297.
Post, J.L. (1984) Saponite from near Ballarat, California. Clays Clay Miner., 32, 147.
Racki, G. and Cordey, F. (2000) Radiolarian palaeoecology and radiolarites: is the
present the key to the past? Earth-Science Reviews, 52, 83-120.
Rehim, H., Mizusaki, A.M.P., Carvalho, M.D. and Monteiro, M. (1986) Talco e
estevensita na formação lagoa feia da bacia de campos--possíveis implicações no ambiente
deposicional. Anais do XXXIV Congresso Brasileiro de Geologia. Goiânia, Goiás, 1986, 1,
416-425.
Sayles, F.L. and Fyfe, W.S. (1973) The crystallization of magnesite from aqueous
solution. Geochimica et Cosmochimica Acta, 37, 87-96.
Setti, M., Marinoni, L. and Lopez-Galindo, A. (2004) Mineralogical and geochemical
characteristics (major, minor, trace elements and REE) of detrital and authigenic clay
minerals in a Cenozoic sequence from Ross Sea, Antarctica. Clay Minerals, 39, 405–421.
Siever, R. (1992) The silica cycle in the Precambrian. Geochimica et Cosmochimica
Acta, 56, 3265-3272.
This article is protected by copyright. All rights reserved.
Simkiss, K. (1964) Variation of the crystallization form of calcium carbonate from
artificial sea water. Nature, 201, 492-493.
Stoessell, R.K. (1988) 25°C and 1 atm dissolution experiments of sepiolite and
kerolite. Geochim. et Cosmochim. Ac. 52, 365-374.
Suquet, H., de la Calle, C. and Pezerat, H. (1975) Swelling and Structural
Organization of Saponite. Clays and Clay Minerals, 23, 1-9.
Taft, W.H. (1967) Physical chemistry of formation of carbonates. In: Carbonate
Rocks – Physical and Chemical Aspects (Eds G.V. Chilingar, H.J. Bissell and R.W.
Fairbridge), pp.151-167. Elsevier, Amsterdam.
Tissot, B.P. and Welte, D.H. (1984) Diagenesis, catagenesis and metagenesis of
organic matter. In: Petroleum Formation and Occurrence, 2nd edn, pp. 69-73. Springer-
Verlag, Berlin, Heidelberg.
Tosca, N.J. (2015) Geochemical pathways to Mg-clay formation. In: Magnesian
Clays: Characterization, Origins and Applications (Eds M. Pozo and E. Galán). AIPEA
Special Publications, 2, 283–329.
Tosca, N.J., Macdonald, F.A., Strauss, J.V., Johnston, D.T. and Knoll, A.H. (2011)
Sedimentary talc in Neoproterozoic carbonate successions. Earth and Planetary Science
Letters, 306, 11-22.
Tosca, N.J. and Wright, V.P. (2015) Diagenetic pathways linked to labile Mg-clays in
lacustrine carbonate reservoirs: a model for the origin of secondary porosity in the Cretaceous
pre-salt Barra Velha Formation, offshore Brazil. In: Reservoir Quality of Clastic and
Carbonate Rocks: Analysis, Modelling and Prediction (Eds P.J. Armitage, A.R. Butcher, J.M.
Churchill, A.E. Csoma, C. Hollis, R.H. Lander, J.E. Omma and R.H. Worden), Geological
Society, London, Special Publications, 435, 33-46.
Tréguer, P., Nelson, D.M., Van Bennekom, A.J., DeMaster, D.J., Leynaert, A. and
Quéguiner, B. (1995) The Silica Balance in the World Ocean: A Reestimate. Science, 268,
375-379.
Van Denburgh, A.S. (1975) Solute Balance at Abert and Summer Lake, South-central
Oregon. Professional Paper 502-C. US Government Printing Office, Washington, 28 pp.
Vicente, M.A., Suárez, M., López-González, J.D.D. and Banares-Munoz, M.A. (1996)
Characterization, surface area, and porosity analyses of the solids obtained by acid leaching
of a saponite. Langmuir, 12, 566-572.
Visser J.W. and de Wolff P.M. (1964) Absolute Intensities. Report 641.109,
Technisch Physische Dienst, Delft, Netherlands.
Walker, G.F. (1958) Reactions of expanding-lattice clay minerals with glycerol and
ethylene glycol. Clay Minerals Bulletin, 3, 302-313.
This article is protected by copyright. All rights reserved.
Weaver, C.E. (1989) Clays, Muds, and Shales. Elsevier Science, Amsterdam, 818 pp.
Weaver, C.E. and Beck, K.C. (1977) Miocene of the S.E. United States: A model for
chemical sedimentation in a peri-marine environment. Sedimentary Geology, 17, 1-234.
White, R. and McKenzie, D. (1989) Magmatism at rift zones: the generation of
volcanic continental margins and flood basalts. Journal of Geophysical Research: Solid
Earth, 94, 7685-7729.
Wilson, M. (1992) Magmatism and continental rifting during the opening of the South
Atlantic Ocean: a consequence of Lower Cretaceous super-plume activity? In: Magmatism
and the Causes of Continental Break-up (Eds B.C. Storey, T. Alabaster, R.J. Pankhurst),
Geological Society, London, Special Publications, 68, 241-255.
Wright, V.P. (2012) Lacustrine carbonates in rift settings: the interaction of volcanic
and microbial processes on carbonate deposition. In: Advances in Carbonate Exploration and
Reservoir Analysis (Eds J. Garland, J.E. Neilson, S.E. Laubach and K.J. Whidden),
Geological Society, London, Special Publications, 370, 39-47.
Wright, V.P. and Barnett, A.J. (2015) An abiotic model for the development of
textures in some South Atlantic early Cretaceous lacustrine carbonates. In: Microbial
Carbonates in Space and Time: Implications for Global Exploration and Production (Eds
D.W.J. Bosence, K.A. Gibbons, D.P. Le Heron, W.A. Morgan, T. Pritchard and B.A.
Vining). Geological Society, London, Special Publications, 418, 209–219.
Yuretich, R.F. and Cerling, T.E. (1983) Hydrogeochemistry of Lake Turkana, Kenya:
Mass balance and mineral reactions in an alkaline lake. Geochimica and Cosmochimica Acta,
47, 1099–1109.
Zeyen, N., Benzerara, K., Li, J., Groleau, A., Balan, E., Estève, I., Tavera, R.,
Moreira, D. and López-García, P. (2015) Formation of low-T hydrated silicates in modern
microbialites from Mexico and implications for microbial fossilization. Frontiers in Earth
Science, 3, 64. doi: 10.3389/feart.2015.00064
Zeyen, N., Daval, D., Lopez-Garcia, P., Moreira, D., Gaillardet, J. and Benzerara, K.
(2017) Geochemical Conditions Allowing the Formation of Modern Lacustrine Microbialites.
Procedia Earth and Planetary Science, 17, 380-383.
Zinke, J., Reijmer, J.J.G. and Thomassin, B.A. (2003) Systems tracts sedimentology
in the lagoon of Mayotte associated with the Holocene transgression. Sedimentary Geology,
160, 57-784.
This article is protected by copyright. All rights reserved.
FIGURE CAPTIONS
Fig. 1. Location map of the study site (A) and (B) and bathymetric map of the Dziani Dzaha
(C) showing bubbling gas sites (red circles), microbialites (grey circles) and sediment core
collection locations (red triangles).
Fig. 2. (A) North–south topographic profile of the lake, showing an ideal representation of
lake sediments. The location of sediment cores is projected onto the topographic profile
which crosses the 18 m deep depression. The insert (white box) shows the topographic profile
of the entire volcanic crater. Vertical exaggeration is 4x. (B) Picture of C12 at 15 to 38 cm
depth showing varves composed of dark and light brown layers. No disturbance of the
sediment pile is observed. (C) Picture of C13 at 44 to 61 cm depth showing deformation
structures. (D) Water/rock weight ratio as function of depth in C12.
Fig. 3. X-ray diffraction pattern of C9 at 54 to 58 cm depth (a.u. – arbitrary unit). Only the
most representative diffraction peaks of aragonite (dark blue), hydromagnesite (light blue),
clinopyroxene (orange) and clay minerals (red) are shown. Values of d-spacing (Ångström)
are inserted for clay minerals.
Fig. 4. Secondary electron (SE) image (A) and chemical composition map of Mg, Si and Ca
(B) and Al (C) showing hydromagnesite (blue), aragonite (yellow) and clay minerals (pink-
blue) of C1 at 22 to 23 cm depth. Transmission electron microscopy (TEM) image (D) and
energy dispersive X-ray fluorescence (EDX) spectrum (E) of clay minerals of C4 at 24 to 27
cm depth. SE image (F), EDX spectrum (G) and quantitative chemical composition (H) of
volcanic glass of C9 at 10 to 14 cm depth. The red crosses indicate the position of the EDX
analyses. Correlation between the Na and Cl signals in the EDX spectrum (E) indicates NaCl
contamination.
This article is protected by copyright. All rights reserved.
Fig. 5. X-ray diffraction (XRD) patterns of the <2 µm clay fraction extracted from C4 at 27 to
30 cm depth acquired on (A) an un-oriented slide; (B) oriented slide in the air-dried state
(black line), ethylene glycol state (blue line), after heating at 500°C and ethylene glycol
salvation (red line); and (C) after Mg saturation (green line) coupled to glycerol solvation
during three days (purple line).
Fig. 6. Mineral and organic matter content (wt.%) as function of depth in C11, C4, C10, C6
and C13. The location of sediment cores is shown on the topographic profile of the lake (Fig.
2A).
Fig. 7. Mineral stability domains in the Ca-Mg-C-O-H and Mg-Si-O-H systems at 30°C and 1
bar. The chemical composition of Dziani Dzaha waters (Leboulanger et al., 2017; Gérard et
al., 2018) is represented with a black box. (A) Ca-Mg-C-O-H system. Magnesite is not taken
into account in the calculation. Aragonite is considered rather than calcite, consistent with the
observations. The grey dotted lines indicate metastable equilibria between dolomite, huntite,
aragonite, hydromagnesite and brucite. The full black lines indicate metastable equilibria
between brucite, aragonite and hydromagnesite when dolomite and huntite are not considered
in the calculation. (B) Mg-Si-O-H system. The full black lines represent metastable equilibria
between Mg-aluminosilicates, when chlorites are not considered in the calculation. The
dotted lines are fluid compositions at equilibrium with talc, sepiolite and kerolite, i.e. three
Al-free Mg-silicates. The thermodynamic data of kerolite are from Stoessell (1988). The
dashed-dotted line corresponds to the critical supersaturation for homogenous nucleation of
Mg-silicates suggested by Tosca (2015). Silica activities at equilibrium with quartz and
amorphous silica are shown (vertical dotted lines).
Fig. 8. Model of the early diagenetic formation of the saponite-like mineral. Dashed arrows
show the possible feedback effect between the destabilization of hydromagnesite and the
precipitation of the saponite-like mineral.
This article is protected by copyright. All rights reserved.
Table 1. Method for quantification of mineral phases.
Table 2. Refined composition of saponite-like mineral in C4 (in weight %). EMPA – electron
microprobe analysis; CEC – cation exchange capacity
Fig. S1. Volcanic catchment: (A) ash layers; (B) fragments of solid/igneous rocks in ash
layers.
This article is protected by copyright. All rights reserved.
Mineral
Chemical formula
Calculation*
Halite
NaCl
Halite = [Cl]
Alkaline
feldspars
Na0.7K0.3AlSi3O8
Alk feld = [K]/0.3
Saponite
K0.01Mg0.12Ca0.02Na0.28(Mg2.22Fe0.18Al0.28)
(Si3.61Al0.39)O10(OH)2
Sap = ([Al] – Alk feld)/0.67
Clinopyroxene
(CayMg1-y)SiO3
Px = [Si] – 3*Alk feld – 3.61*Sap
Hydromagnesite
Mg5(CO3)4(OH)2·(H2O)4
Hydromag = ([Mg] – 2.34*Sap –
(1-y)*Px)/5
Aragonite
CaCO3
Ara = ([Ca] – y*Px – 0.02*Sap)
Pyrite
FeS2
Pyr = [S]/2
Magnetite
Fe3O4
Mag = ([Fe] – Pyr – 0.18*Sap)/3
*[X] are the concentrations in mol.% of Cl, K, Al, Si, Mg, Ca, S and Fe measured with XRF.
Sample
EPMA results
CEC results
Al
Ca
Fe
K
Mg
Mn
Na
Si
Na
K
Ca
Mg
C4
3.65
0.20
2.01
0.07
11.62
0.07
0.70
20.78
0.65
0.02
0.04
0.15
This article is protected by copyright. All rights reserved.
This article is protected by copyright. All rights reserved.
This article is protected by copyright. All rights reserved.
This article is protected by copyright. All rights reserved.
This article is protected by copyright. All rights reserved.