You are on page 1of 21

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

net/publication/316817008

Iron and sulfur isotope constraints on redox conditions associated with the 3.2
Ga barite deposits of the Mapepe Formation (Barberton Greenstone Belt,
South Africa)

Article  in  Geochimica et Cosmochimica Acta · May 2017


DOI: 10.1016/j.gca.2017.05.002

CITATIONS READS

2 261

7 authors, including:

Vincent Busigny Johanna Marin-Carbonne


Paris Diderot University University of Lausanne
60 PUBLICATIONS   1,159 CITATIONS    37 PUBLICATIONS   342 CITATIONS   

SEE PROFILE SEE PROFILE

Elodie Muller Claire Rollion-Bard


Institut de Physique du Globe de Paris Institut de Physique du Globe de Paris
19 PUBLICATIONS   29 CITATIONS    117 PUBLICATIONS   1,891 CITATIONS   

SEE PROFILE SEE PROFILE

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

Reconstruction of ancient environments and ecosystems from the geological rock record View project

understanding coral proxies View project

All content following this page was uploaded by Vincent Busigny on 07 December 2018.

The user has requested enhancement of the downloaded file.


Available online at www.sciencedirect.com

ScienceDirect
Geochimica et Cosmochimica Acta 210 (2017) 247–266
www.elsevier.com/locate/gca

Iron and sulfur isotope constraints on redox conditions


associated with the 3.2 Ga barite deposits of the
Mapepe Formation (Barberton Greenstone Belt, South Africa)
Vincent Busigny a,⇑, Johanna Marin-Carbonne b, Elodie Muller a, Pierre Cartigny a,
Claire Rollion-Bard a, Nelly Assayag a, Pascal Philippot a
a
Institut de Physique du Globe de Paris, Sorbonne Paris Cité, Université Paris Diderot, UMR 7154 CNRS, 75238 Paris, France
b
Univ Lyon, UJM Saint Etienne, Laboratoire Magmas et Volcans, CNRS UMR 6524, UCA, IRD, 23 rue du Dr Paul Michelon,
42100 Saint Etienne, France

Received 4 July 2016; accepted in revised form 1 May 2017; Available online 9 May 2017

Abstract

The occurrence of Early Archean barite deposits is intriguing since this type of sediment requires high availability of dis-
solved sulfate (SO2
4 ), the oxidized form of sulfur, although most authors argued that the Archean eon was dominated by
reducing conditions, with low oceanic sulfate concentration (<10 lM) relative to present day levels of 28,000 lM. In order
to better assess the redox state of the paleo-atmosphere and -oceans, we examined Fe and S isotope compositions in a sed-
imentary sequence from the 3.2 Ga-old Mendon and Mapepe formations (Kaapvaal craton, South Africa), recovered from
the drill-core BBDP2 of the Barberton Barite Drilling Project. Major elements were also analyzed to constrain the respective
imprints of detrital vs metasomatic processes, in particular using Al, Ti and K interrelations. Bulk rock Fe isotope compo-
sitions are linked to mineralogy, with d56Fe values varying between 2.04‰ in Fe sulfide-dominated barite beds, to
+2.14‰ in Fe oxide-bearing cherts. d34S values of sulfides vary between 10.84 and +3.56‰, with D33S in a range comprised
between 0.35 and +2.55‰, thus supporting an O2-depleted atmosphere (<105 PAL). Iron isotope variations together with
major element correlations show that, although the sediments experienced a pervasive stage of hydrothermal alteration, the
rocks preserved a primary/authigenic signature predating subsequent hydrothermal stage. Highly positive d56Fe values
recorded in primary Fe-oxides from ferruginous cherts support partial Fe oxidation in a reducing oceanic environment
(O2 < 104 lM), but are incompatible with a model of complete oxidation at the redox boundary of a stratified water column.
Iron oxide precipitation under low O2 levels was likely mediated by anoxygenic photosynthesis, and/or abiotic photo-
oxidation processes. Our results are consistent with global anoxic conditions in the 3.2 Ga-old sediments, implying that
the barite deposits were most likely sourced by atmospheric photolysis of S gases produced by large subaerial volcanic events,
and possibly SO2 4 produced by magmatic SO2 disproportionation in hydrothermal systems.
Ó 2017 Elsevier Ltd. All rights reserved.

Keywords: Iron isotopes; Archean; Paleo-environment; Barite deposits; Anoxic ocean

1. INTRODUCTION

Understanding the evolution of Earth’s surface oxygena-


tion through geological time is crucial for constraining the
⇑ Corresponding author. development of life on Earth. The release and accumulation
E-mail address: busigny@ipgp.fr (V. Busigny).

http://dx.doi.org/10.1016/j.gca.2017.05.002
0016-7037/Ó 2017 Elsevier Ltd. All rights reserved.
248 V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266

of O2 in the ocean and atmosphere imparted strong amount of O2 in the paleoenvironment: (1) a negative Ce
modifications of the geochemical cycles of redox-sensitive anomaly in some greywackes and a ferruginous chert and
elements (e.g. Fe, S, Mo, U, Cr) and their co-evolution with (2) the presence of primary hematite in ferruginous cherts
living organisms (Anbar and Knoll, 2002; Konhauser et al., (jaspilites). Given the apparent lack of secondary/late oxi-
2011; Lyons et al., 2014, 2015). The production of O2 on the dation, the signal was considered as primary (Hofmann,
Earth’s surface is attributed essentially to oxygenic photo- 2005). It could either record a hydrothermal signal being
synthesis, implying that constraints on the timing of early unrelated to surface oxygenation (Hofmann, 2005), or pro-
O2 production could help to establish the development of vide geochemical evidence for some of the earliest traces of
this metabolic activity (Farquhar et al., 2011; see also oxygenic photosynthesis on Earth. The fact that most sam-
Catling, 2014, for a detailed discussion about sources and ples of this previous study did not show Ce anomaly sug-
sinks of atmospheric O2). It is generally considered that gests that the underlying process causing the Ce loss was
ocean-atmosphere system during the Archean period was localized and not representative of a general state of the
depleted in O2 until the Great Oxidation Event (GOE) water column. Interestingly, the possibility for local oxygen
sometime between 2.45 and 2.20 Ga (Holland, 1984, production at that time is further suggested by interpreta-
2002; Bekker et al., 2004). The most compelling evidence tion of microbial mats in the Moodies Group from Barber-
for low concentration of O2 in the Earth’s atmosphere is ton as cyanobacteria fossils (Homann et al., 2015, 2016).
the presence of significant mass-independent fractionation Another important feature of the 3.2 Ga sedimentary rocks
(MIF) recorded in S isotope compositions from Archean from the Fig Tree Group is the presence of extensive sedi-
sedimentary sulfides and sulfates (Farquhar et al., 2000; mentary barite (BaSO4) deposits. This type of sediments
see also review by Johnston, 2011). The presence of S- requires high availability of dissolved sulfate (SO2 4 ), the
MIF has been interpreted as resulting from photochemical oxidized form of sulfur, and may thus point to the presence
reactions of S-species (SO2 and/or OCS) released from vol- of oceanic oxic oases. Alternatively, sulfate could derive
canoes to the atmosphere (e.g. Farquhar et al., 2000; Ueno from photolysis of SO2/OCS in an O2-poor atmosphere
et al., 2009; Thomassot et al., 2015). Most authors concur and subsequent hydrolysis of the products with water
that the generation and preservation of S-MIF signal in (Huston and Logan, 2004; Bao et al., 2007; Roerdink
the Archean rock record requires extremely low atmo- et al., 2012; Muller et al., 2016). Intense subaerial volcanic
spheric O2 concentration, probably less than 105 present activity could for instance lead to massive accumulation of
atmospheric level (PAL; Pavlov and Kasting, 2002). During sulfate in the shallow-ocean (Philippot et al., 2012; Muller
the last decade, attempts were made to track the rise of et al., 2016). Another possibility would be that SO2 4 was
atmospheric O2 using several new or emerging geochemical produced by disproportionation of magmatic SO2 during
proxies such as redox-sensitive metals (Fe, Mo, Re, Cr, U). hydrothermal reactions between water and volcanic rocks
These proxies have highlighted that the rise of atmospheric (e.g. Kusakabe et al., 2000). In the present work, we inves-
O2 may have been more complex than previously thought, tigated the redox state of the paleo-atmosphere and -ocean
and that some fluctuations in the O2 production and accu- by coupling S and Fe isotope compositions recorded in a
mulation may be tracked back several hundred million sedimentary sequence from the 3.2 Ga-old Mendon and
years before O2 became a permanent part of the atmo- Mapepe Formations (Barberton Greenstone Belt, South
sphere (e.g. Anbar et al., 2007; Wille et al., 2007; Frei Africa). The Mapepe Formation represents the lowermost
et al., 2009; Scott et al., 2011; Czaja et al., 2012; Zerkle unit of the Fig Tree Group, South of the Inyoka Fault
et al., 2012; Crowe et al., 2013; Planavsky et al., 2014). (Fig. 1), and contains abundant bedded barites, thus allow-
Alternatively, oxygen could have risen in ocean oases, soil ing us to explore potential variations in the redox condi-
profiles or microbial mats and have no impact on atmo- tions. The Mendon Formation, underlying the Mapepe
spheric O2 levels (e.g. Kasting, 1992; Olson et al., 2013; Formation, belongs to the Onverwacht Group and has been
Lalonde and Konhauser, 2015). Much of the proxy data selected to evaluate the chemical and isotopic effect of late
could therefore have been affected by local effects, unrelated hydrothermal circulation. The advantage of coupling Fe
to atmospheric ‘‘whiffs” of oxygen (Anbar et al., 2007). and S isotopic systems is that Fe isotopes provide informa-
Additionally, the geological record of these proxies is sparse tion about the water column and diagenetic processes
and difficult to correlate between cratons, and fundamental (Rouxel et al., 2005; Yamaguchi et al., 2005; Severmann
questions about their interpretation complicate the picture. et al., 2006, 2008; Staubwasser et al., 2006; Yoshiya et al.,
For instance, the burial and diagenetic history of rocks (e.g. 2012; Johnson et al., 2013; Busigny et al., 2014), whereas
Li et al., 2013) might lead to a false recognition of Archean sulfur can bring insight into atmospheric processes and bio-
oxygenic photosynthesis or oxygenated atmosphere (tracing geochemical cycling in the water column and sedimentary
for instance other oxidants than O2). interface (Farquhar et al., 2000; Pavlov and Kasting,
In a recent study coupling Fe isotope composition and 2002; Bao et al., 2007; Johnston, 2011; Roerdink et al.,
U concentrations in the Manzimnyama Banded Iron For- 2012; Muller et al., 2016).
mations (BIF) from the Fig Tree Group (Barberton Green-
stone Belt, South Africa), Satkoski et al. (2015) proposed 2. GEOLOGICAL BACKGROUND
that the ocean was stratified, with oxic surface waters and
anoxic and ferruginous deep waters, 3.2 Ga ago (i.e. The Barberton Greenstone Belt (BGB) in South Africa
900 Ma before the GOE). In the Fig Tree Group, two other consists of a succession of Archean supracrustal rocks sur-
observations highlighted the possibility for significant rounded and intruded by granitoids (Viljoen and Viljoen,
V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266 249

Fig. 1. (A) Geological map of the Barberton Greenstone Belt, South Africa, showing the location of the drill-core BBDP2 (red star). (B)
Stratigraphic log of BBDP2 drill core after Philippot et al. (2012). Stratigraphic legend: 1, shales with local ferrigenous chert beds (red layer);
2, Felsic volcanoclastic and terrigenous sediments with local barite beds (green layer) and ferrigenous cherts; 3, bedded barite with pyrite
laminates; 4, Laminated black chert with local breccia and vein network; 5, Strongly silicified basaltic komatiites preserving local spinifex
texture (v) and cut across by numerous black and white hydrothermal chert veins. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)

1969). Supracrustal rocks from the BGB are termed the et al., 1996). Rocks from the Onverwacht and Fig Tree
Swaziland Supergroup and are subdivided into three strati- Groups experienced low greenschist facies regional meta-
graphic units: the Onverwacht, Fig Tree and Moodies morphism with temperatures not exceeding 350 °C (Xie
Groups (Fig. 1). Samples analyzed in the present work rep- et al., 1997; Toulkeridis et al., 1998; Knauth and Lowe,
resent the transition between the Onverwacht and Fig Tree 2003; Tice et al., 2004; Farber et al., 2015).
Groups. They belong to the Mendon and Mapepe Forma-
tions, which correspond respectively to the upper section of 3. SAMPLES DESCRIPTION
the Onverwacht Group and lower section of the Fig Tree
Group. The Mendon Formation is a dominantly ultramafic In the present study, samples were collected from a
volcanic succession, interpreted to reflect large mantle weathering-free drill-core recovered by the Barberton Bar-
plume magmatic activity (Lowe, 1994). Basaltic komatiites ite Drilling Project (BBDP2; Fig. 1, Philippot et al., 2009,
from Mendon Formation are alternated with bedded 2012). Samples were selected as representative of the diver-
cherts, probably recording deep-water sedimentation dur- sity of lithologies forming the Mendon and Mapepe For-
ing times of magmatic quiescence (Hofmann and Harris, mations. These include basaltic komatiites and
2008). A tuffaceous band from the upper Kromberg Forma- carbonaceous black cherts from Mendon, together with
tion, underlying the Mendon Formation, contains zircons bedded barites, ferruginous cherts, and silicified terrigenous
with a mean age of 3334 ± 3 Ma (Byerly et al., 1996), thus and volcanoclastic sediments from Mapepe.
providing a lower age limit for Mendon Formation. Basaltic komatiites from Mendon display locally spini-
Komatiites and cherts from Mendon were affected by wide- fex textures with olivine and pyroxene pseudomorphed by
spread hydrothermal alteration as illustrated by a petrolog- quartz, carbonate and chlorite (Lowe, 1994). These rocks
ical study of samples from various localities (e.g. Barite were extensively hydrothermally-altered and silicified, and
Valley Syncline, Power Line Road section, Msauli Gorge; occur transected by numerous generations of quartz-
Farber et al., 2015). The Mapepe Formation at the base carbonate and black-chert veins. Three meta-komatiite
of the Fig Tree Group records shallow-water terrigenous samples were studied here: 180-32, 160-12 and 143-7. Sam-
and volcanoclastic sediments, interbedded with felsic vol- ples 180-32 and 143-7 are silicified meta-komatiites, while
canic deposits. The Fig Tree Group sediments were depos- sample 160-12 represents a vein composed of quartz, car-
ited between 3259 ± 3 and 3226 ± 3 Ma, as determined bonate, pyrite and organic material. Basaltic meta-
from zircon of volcanoclastic rocks at the base and the komatiites from Mendon are intercalated and overlain by
top of the group, respectively (Kröner et al., 1991; Byerly finely-laminated black cherts (Fig. 2D, H) containing a sig-
250 V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266

Fig. 2. Selected cm-scale photographs (A-D) and photomicrographs (E-H) illustrating the variety of samples from Mapepe and Mendon
Formations analyzed in the present study. (A) and (E): ferruginous chert layer interleaved with fine-grained volcanoclastic sediments. (B) and
(F): terrigenous and volcanoclastic sediments. (C) and (G): bedded barite with laminated pyrite layers located at the base of the Mapepe
Formation. (D) and (H): laminated black chert of the Mendon Formation. Fe-ox: iron oxides, l-qtz: microquartz, Carb: carbonate, Py:
pyrite, Bar: barite.

nificant content of organic matter between 0.21 and 0.87%. The overlying Mapepe Formation consists of volcan-
Five sedimentary black cherts were analyzed: 181-19, 139-8, oclastic and terrigenous sediments (Fig. 2B) interleaved
114-33, 95-41 and 85-45. They are composed of micro- with local barite beds (Fig. 2C) and ferruginous cherts
quartz, iron carbonate (ankerite), sulfide (pyrite), chlorite (Fig. 2A). Low-grade hydrothermal alteration is wide-
and ganterite (i.e. Ba-muscovite). Iron oxide and chalcopy- spread, resulting in the replacement of many primary con-
rite are found in some samples as minor phases. stituents by silica, ankerite and sericite. It has been
suggested that perhaps 50% of the Mapepe Formation is
V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266 251

composed of altered dacitic erupted ash and volcanoclastic 4. ANALYTICAL METHODS


material and that the associated terrigenous sediments were
derived from uplift and erosion of the Onverwacht Group For each drill-core sample, uniform layers of 0.5–2 cm
together with previously formed units of the lower Fig Tree thickness were selected, cut and powdered to a fraction
Group (Lowe and Nocita, 1999). In the BBDP2 drill-core, <60 lm. The same powders were used for Fe, S and major
the contact between Mendon and Mapepe is an erosional and trace element analyses. Iron isotope compositions were
discontinuity, marked by brecciated tuffaceous material determined using a method thoroughly described in
overlain by barite beds. The barite beds consist of green- Busigny et al. (2014) and only summarized herein. About
colored1 cm-scale barite blades and cauliflowers in associa- 10–35 mg of each sample were acidified by successive mix-
tion with disseminated and bedded pyrites (Fig. 2C). Sec- tures of HF, HNO3 and/or HCl to ensure complete diges-
ondary hydrothermal alteration overprint led to intense tion. For samples containing barite, a solid residue
silicification and partial replacement of barite by (corresponding to barite insoluble fraction) could not be
microquartz. dissolved and was removed by centrifugation. Iron con-
Eight bedded barite samples were investigated: 78-18, tained in the soluble fraction was separated from matrix
78-08, 78-06, 77-5, 77-30, 76-76, 67-95 and 67-93. The sam- elements on anion exchange chromatography in HCl med-
ples located immediately above the Mendon-Mapepe con- ium (Strelow, 1980; Dauphas et al., 2004). Iron concentra-
tact (78-18, 78-08 and 78-06) are generally laminated with tions and isotopic compositions were measured using a
large barite crystal fans alternating with mm-scale pyrite ThermoFischer Neptune MC-ICP-MS (Multiple Collector
layers, while other samples display accumulation of non- Inductively Coupled Plasma Mass Spectrometer). Sample
oriented barite crystals with local needle-like texture solutions were nebulized and introduced using ESI Apex-
(Fig. 2C, G). HF desolvating apparatus. Iron isotopes were resolved
Five samples representing terrigenous and volcanoclas- from argide interferences using the high-resolution mode
tic sediments have been selected: 68-42, 68-05, 56-66, 56- of the Neptune (Weyer and Schwieters, 2003; Schoenberg
30 and 45-99. They are composed of silica shards of various and von Blanckenburg, 2005). Instrumental mass bias was
sizes and large pseudomorphs of barite grains, replaced by corrected by using the conventional sample-standard brack-
microquartz. They are strongly silicified and contain locally eting (SSB) approach (Rouxel et al., 2003; Albarède and
rounded to angular pyrites, and rounded quartz grains. On Beard, 2004). The 56Fe/54Fe and 57Fe/54Fe ratios were
the basis of trace element abundances and modeling, expressed in the d notation in per mil (‰) as:
Hofmann (2005) suggested that hydraulic sorting did not
affect the sedimentary abundances of Rare Earth Elements d56 Fe ¼ ½ð56 Fe=54 FeÞsample =ð56 Fe=54 FeÞstandard  1  1000
(REE) and High-Field Strength Elements (HFSE), imply-
and
ing short distances of sedimentary transport of the detrital
phases. This conclusion is consistent with the dominant
d%57 Fe ¼ ½ð57 Fe=54 FeÞsample =ð57 Fe=54 FeÞstandard  1  1000
angular shape of the sedimentary particles included in our
samples. The angular grains present in our samples corre-
spond to quartz, iron carbonate (siderite and ankerite), iron where the standard is IRMM-014, from the Institute for
oxide, barite needle and pyrite. These grains are embedded Reference Materials and Measurements (Taylor et al.,
in a fine matrix dominated by quartz and ankerite. Samples 1992). The analytical blank was always below 30 ng Fe
68-42 and 68-05 contain local spherule textures composed and represented less than 0.3% of the bulk Fe in the sam-
mainly of quartz and phyllosilicates (i.e. chlorite and seri- ples. Based on replicate analyses of international rock geo-
cite). The spherules are thought to represent distal fallout standards (IF-G, ACE and BCR2), the external precision
formed within an impact plume after large meteoritic and accuracy were always better than 0.06‰ for d56Fe
impact event at 3.24 Ga (Lowe et al., 2003). However, the and 0.12‰ for d57Fe (2SD).
low abundance of spherules in sample 68-05, relative to typ- Multiple sulfur isotope analyses of pyrites and sulfates
ical spherule beds (e.g. Krull-Davatzes et al., 2012; van have been performed on bulk rock powders. The pyrites
Zuilen et al., 2014), likely indicates a re-depositional/ were extracted from bulk sample powders with standard
detrital origin. chromium reduction technique, converting S into Ag2S
Four ferruginous cherts were analyzed: 39-75, 50-25, 52- (Canfield et al., 1986), while sulfates were reacted using
72a and 52-72b (Fig. 2A, E). They are composed of micro- Thode’s solution (Thode et al., 1961) and converted to
quartz with various proportions of ankerite and Fe oxides Ag2S. Ag2S was then converted to SF6 by fluorination,
(present as tiny crystals <2 lm), together with minor Ba- purified first by cryogenic separation and then by gas chro-
rich feldspar or mica (ganterite) and calcite. Samples 52- matography (Labidi et al., 2012). Finally, the purified SF6
72a and 52-72b (ferruginous chert, this one can directly was introduced into a MAT253 isotope-ratio mass spec-
be compared to the study of Satkoski et al., 2015) are trometer (IRMS) for multiple sulfur isotope measurement
two different pieces of the same rock, allowing us to evalu- in dual-inlet mode. All isotopic compositions are reported
ate potential heterogeneities due to finely laminated as per mil (‰) deviations from V-CDT standard using
structures. the conventional notation: d34S = [(34S/32S)sample/
(34S/32S)V-CDT  1]  1000.
1
D33S and D36S are used as a measurement of the devia-
For interpretation of color in Fig. 2, the reader is referred to the tion from the terrestrial mass-dependent fractionation line
web version of this article.
252 V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266

and represent the Mass Independent Fractionations (MIF) tion are extremely depleted in Fe, with Fe/Al ratios between
expressed as (Farquhar et al., 2001): 0.10 and 0.17 (samples 45-99, 67-93, 76-76, 77-30; Table 1).
These four Mapepe samples are not restricted to any given
D33S = 1000  [(1 + d33S/1000)  (1 + d34S/1000)0.515] lithology and occur in bedded barites, as well as terrigenous
D36S = 1000  [(1 + d36S/1000)  (1 + d34S/1000)1.89] and volcanoclastic sediments.
Iron isotope compositions are given in Table 2 and illus-
The analytical reproducibility (2SD) on d34S, D33S and trated as a function of depth in Fig. 4. The d56Fe values in
36
D S, estimated from replicate analyses of the IAEA-S1 Mendon vary in a limited range from 0.15 ± 0.03 to
standard, was better than ±0.2‰, ±0.02‰ and ±0.2‰ +0.50 ± 0.05‰, in comparison with Mapepe samples,
respectively. which show a high variability between 2.04 ± 0.04 and
Whole rock major element concentrations were mea- +2.14 ± 0.07‰. For the latter, there is a strong mineralog-
sured using ICP-OES at the Service d’Analyse des Roches ical control, with the most negative values observed in bed-
et des Mineraux (SARM) of the Centre de Recherches ded barites containing pyrite laminates, while the most
Pétrographiques et Géochimiques (CRPG) of positive values are recorded in ferruginous cherts contain-
Vandoeuvre-Lès-Nancy, France, following the method ing Fe oxides. Mn/Fe ratios display large variations over
described in Carignan et al. (2001). Analytical precision three orders of magnitude, with values from 0.007 to 0.08
and detection limits for major element concentrations are in Mendon and 0.0001 to 0.246 in Mapepe samples
typically better than ±10%. The determination of the major (Table 1). Interestingly these variations are also related to
element composition of bedded barite samples was incom- mineralogy. Low Mn/Fe ratios correspond to samples in
plete because of the difficulty of dissolving barite minerals. which Fe-bearing phases are mostly pyrite or Fe-oxides,
Barite solid residue was thus separated by centrifugation while high Mn/Fe ratios are attributed to carbonates (i.e.
after conventional dissolution of other phases using mixing ankerite, as observed by SEM analyses on thin sections).
of HNO3-HF, and HF-HCl. Only elements with reliable A plot of d56Fe values as a function of Mn/Fe ratio
concentration estimates are provided for these samples. (Fig. 5) shows linear trends clearly linked to mineralogical
variations. It can be emphasized that the strongly
5. RESULTS hydrothermally-altered basaltic komatiites from Mendon
show a large variability in Mn/Fe ratios (0.023–0.080)
5.1. Major element concentrations: a mineralogical control and d56Fe values (0.15 ± 0.03 to +0.49 ± 0.02‰) com-
pared to fresh komatiites, which are characterized by Mn/
Major element concentrations in Mapepe and Mendon Fe  0.015 ± 0.004 (Arndt, 1986) and d56Fe  0.04
samples are given in Table 1 and some general features ± 0.03‰ (Dauphas et al., 2010). Hydrothermal alteration
are illustrated in Fig. 3. SiO2 content shows a large range thus modified both the Fe content and isotope composition
from 17 to 96 wt% (with most samples >60 wt%). SiO2 con- of Mendon komatiites.
tent is inversely correlated with MgO and CaO (Fig. 3A).
This indicates that the composition of these samples is 5.3. Multiple sulfur isotope compositions
dominated by a mixture of quartz and carbonate, in good
agreement with thin section observations. Al2O3 and TiO2 Sulfur isotope compositions of bulk sulfides in Mapepe
contents vary from 0.14 to 4.76 wt%, and 0.02 to 0.66 wt and Mendon samples are shown in Table 2 and illustrated
%, respectively. TiO2 is linearly correlated with Al2O3 and as a function of depth in Fig. 4. d34S values in Mendon dis-
shows two different trends for the Mendon and Mapepe play relatively uniform and positive values between +0.28
Formations (Fig. 3B). K2O shows also a linear correlation and +3.27‰ (±0.02‰, 2SD). In contrast, Mapepe samples
with Al2O3 but with no distinction between Mendon and show a larger variability, with mostly negative d34S values
Mapepe data. Mineralogical analysis indicates that K and from 10.84 to 1.25‰, except one sample with a positive
Al are both carried by mica, suggesting that the variations value of +1.44‰ (sample 56-66). The three samples with
in K and Al contents reflect various proportions of mica in the most negative d34S values show the lowest Fe content
the host rock (Fig. 3C). and Fe/Al ratio (samples 67-93, 76-76, 77-30). D33S is sig-
nificantly different from 0‰ in most samples, with values
5.2. Iron concentration and isotope distribution ranging from 0.15 to +2.55‰ in Mendon and 0.35 to
+0.87‰ in Mapepe (±0.02‰, 2SD). D36S varies from
Iron concentration in Mendon samples covers a wide 2.05 to 0.06‰ in Mendon and from 0.87 to +0.09‰
range from 0.55 to 9.11 wt%, while Mapepe samples display in Mapepe (±0.2‰, 2SD; Table 2). A plot of D36S vs.
even larger variations between 0.07 and 21.38 wt% D33S shows a rough inverse correlation (not shown), with
(Table 1). In order to avoid any dilution effect by carbonate a slope of 0.81 (R2 = 0.38), similar to the trend defined
or quartz precipitation, Fe variations can be evaluated by most sedimentary rocks from the Archean eon
using the Fe/Al molar ratio rather than pure concentration. (slope  0.9; e.g. Farquhar et al., 2000). Fig. 6A displays
The Fe/Al molar ratio ranges from 0.26 to 10.53 in Mendon the relation between D33S and d34S values in Mendon and
and 0.10 to 37.54 in Mapepe. Compared to the average Fe/ Mapepe samples, together with barite data from Mapepe
Al ratio in shales deposited in oxic and suboxic environ- reported by previous studies (Bao et al., 2007; Roerdink
ments (0.5; Taylor and McLennan, 1985; Lyons and et al., 2012; Montinaro et al., 2015). Fig. 6B reports d56Fe
Severmann, 2006), four samples from the Mapepe Forma- vs d34S for Mendon and Mapepe samples. Although no
Table 1
Major element concentrations, total organic carbon content (TOC) and selected molar ratios in Mapepe and Mendon samples from the core BBDP2.
Sample Rock type Deptha (m) SiO2 (wt%) Al2O3 (wt%) Fe2O3 (wt%) MnO (wt%) MgO (wt%) CaO Na2O K2O TiO2 PF Total TOC Mn/Fe Fe/Al
(wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (molar) (molar)

V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266


Mapepe Formation
39-75 Ferruginous chert 39.75 87.70 0.14 1.39 0.090 1.33 2.48 0.01 0.03 4.94 98.10 0.03 0.07 6.53
45-99 Terrig. & volc. sed. 45.99 65.69 4.76 0.74 0.006 0.12 0.17 0.06 0.88 0.66 5.95 79.03 0.02 0.01 0.10
50-25 Ferruginous chert 50.25 69.03 1.96 2.70 0.295 4.22 7.66 0.03 0.40 0.15 12.43 98.88 0.02 0.12 0.88
52-72a Ferruginous chert 52.72 94.73 0.17 0.74 0.009 0.13 0.41 0.02 0.03 1.59 97.81 0.02 0.01 2.79
52-72b Ferruginous chert 52.72
56-30 Terrig. & volc. sed. 56.30 79.61 0.65 1.19 0.260 3.23 5.62 0.02 0.15 0.04 8.76 99.53 0.25 1.17
56.66 Terrig. & volc. sed. 56.66 78.24 2.13 1.28 0.213 2.83 4.87 0.03 0.46 0.21 7.91 98.16 0.04 0.19 0.38
67-93 Bedded barite 67.93 0.40 0.08 0.002 0.06 0.07 0.10 0.03 0.12
67-95 Bedded barite 67.95 0.01
68-05 Terrig. & volc. sed. 68.05 61.51 1.63 2.52 0.400 5.20 8.85 0.03 0.25 0.13 13.72 94.24 0.02 0.18 0.98
68-42 Terrig. & volc. sed. 68.42 61.98 0.90 1.47 0.145 2.66 4.44 0.14 0.07 9.83 81.63 0.03 0.11 1.05
76-76 Bedded barite 76.76 0.34 0.07 0.010 0.20 0.29 0.16 0.13
77-30 Bedded barite 77.30 0.26 0.07 0.007 0.11 0.11 0.11 0.17
77-50 Bedded barite 77.50 0.93 18.01 0.004 0.07 0.21 0.09 0.00 12.43
78-06 Bedded barite 78.06 0.34 20.00 0.002 0.00 37.54
78-08 Bedded barite 78.08 0.17 0.31 0.003 0.09 0.13 0.03 0.01 1.16
78-18 Bedded barite 78.18 0.53 21.38 0.003 0.10 0.04 0.00 25.76
Mendon Formation
85-45 Black chert 85.45 77.89 1.94 3.40 0.196 2.82 4.79 0.04 0.44 0.06 9.06 100.64 0.06
95-41 Black chert 95.41 96.05 1.34 0.55 0.004 0.13 0.11 0.02 0.29 0.05 1.46 100.00 0.40 0.01
114-33 Black chert 114.33 84.77 3.01 4.44 0.067 1.04 0.05 0.02 0.73 0.13 4.63 98.88 0.87 0.02
139-80 Black chert 139.80 82.57 1.12 7.81 0.153 1.64 0.15 0.01 0.27 0.05 5.72 99.49 0.25 0.02
143-70 Komatiite 143.70 16.81 2.38 6.79 0.483 12.96 23.02 0.04 0.68 0.07 35.95 99.19 0.08
160-12 Komatiite 160.12 75.07 4.39 4.25 0.089 2.13 3.80 0.05 1.21 0.15 7.93 99.06 0.02 0.02
180-32 Komatiite 180.32 24.38 0.55 9.11 0.434 12.39 19.03 0.04 0.05 0.02 32.83 98.84 0.05
181-19 Black chert 181.19 45.99 3.58 5.58 0.287 7.69 13.45 0.19 0.02 0.12 21.56 98.45 0.06
a
Depth in the drill-core recovered by the Barberton Barite Drilling Project BBDP2. terrig. & volc. sed.: terrigenous and volcanoclastic sediment.

253
254 V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266

clear trend can be observed, the distribution of the data


illustrates a strong petrological control, with the various
lithologies corresponding to separate fields. The most strik-
ing feature is that bedded barites define two different
groups, which are not detected in a D33S-d34S plot. The first
group, with the most negative d34S (10.84 to 5.59‰) and
d56Fe from 0.90 to 0.00‰, corresponds to samples with
the lowest Fe concentration (0.05 wt% Fe; Table 2). The
second group has moderately negative d34S (5.27 to
2.03‰) and extremely negative d56Fe (2.04 to
1.60‰) and corresponds to samples with the highest Fe
concentrations (0.22–14.93 wt%). Two groups of bedded
barite can thus be distinguished based on their Fe concen-
trations, reflecting pyrite content, and its associated Fe
and S isotope compositions.

6. DISCUSSION

6.1. Evidence for detrital and hydrothermal influences

Before deciphering primary authigenic signatures and


reconstructing paleo-environmental conditions associated
with Mapepe and Mendon sediment deposition, it is impor-
tant to evaluate the role of detrital and hydrothermally-
altered components on chemical and isotope compositions
of bulk samples. Detrital influence can be evaluated using
Al and Ti contents since these elements are generally not
mobilized by fluid circulation (Rouxel et al., 2003;
Dauphas et al., 2007), except under exceptionally low pH
(<2; Rusk et al., 2008) or high pressure and temperature
conditions (>1 GPa, 800 °C; Manning, 2006). The correla-
tions observed between Al and Ti in Mapepe and Mendon
Formations are consistent with minimal hydrothermal
alteration for these elements, and support their use as indi-
cators of detrital input. The two different trends defined by
Mapepe and Mendon data indicate that Al and Ti sources
were distinct in these formations (Fig. 3B). Lowe and
Nocita (1999) and Hofmann (2005) suggested that the detri-
tal component included in the Mapepe sedimentary depos-
its derived from uplift and erosion of rocks from the
Onverwacht Group, possibly the Mendon Formation. This
hypothesis was also suggested earlier for the Loenen For-
mation of the Fig Tree Group based on major element
chemistry and petrographic work (Heinrichs, 1980). The
different Al/Ti ratios observed here for Mapepe and Men-
don indicate that Mapepe detrital fraction was not derived
from erosion of the Mendon Formation, and therefore that
another detrital source is required at least in this part of the
oceanic basin.
Contrasting with the two different trends observed
between Al and Ti, plotting Al2O3 vs K2O concentrations
Fig. 3. Major element concentrations (in wt%) showing the main results in a single correlation including samples from both
geochemical features of the samples from Mapepe (square) and Mapepe and Mendon Formations (Fig. 3C). This likely
Mendon (circle) Formations. (A) SiO2 vs CaO: samples are mostly reflects the addition of K by fluid-rock interaction (e.g.
composed of quartz and carbonate. (B) Al2O3 vs TiO2: the two Toulkeridis et al., 1998; Hofmann, 2005). Mineralogical
different trends illustrate that the rocks from Mendon are not the observations coupled with SEM analyses of our samples
source of Mapepe detrital component. (C) K2O vs Al2O3: the single indicate that K is mostly hosted by mica, in the form of ser-
trend for Mapepe and Mendon indicates that all the rocks were
icite and ganterite. The capacity of a rock to incorporate K
hydrothermally altered.
Table 2
Iron and sulfur isotopes data in Mapepe and Mendon samples from the core BBDP2.
Sample Fea d56Fe ±2r d57Fe ±2r Sulfide Sulfate
(wt%) (‰) (‰) (‰) (‰) 34
d S ±2r 33
D S ±2r 36
D S ±2r d34S ±2r D33S ±2r D36S ±2r
(‰) (‰) (‰) (‰) (‰) (‰) (‰) (‰) (‰) (‰) (‰) (‰)

V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266


Mapepe Formation
39-75 0.98 1.20 0.02 1.81 0.06 2.27 0.02 0.31 0.02 0.36 0.20
45-99 0.50 0.17 0.03 0.24 0.08 1.25 0.02 0.06 0.02 0.29 0.20
50-25 1.67 0.22 0.08 0.54 0.01 4.59 0.02 0.40 0.02 0.08 0.20
52-72a 0.51 2.14 0.07 3.33 0.10 4.44 0.02 0.87 0.02 0.87 0.20
52-72b 2.65 0.01 0.04 0.04 0.03
56-30 0.85 0.39 0.04 0.56 0.06 2.47 0.02 0.01 0.02 0.50 0.20
56.66 0.77 0.25 0.05 0.33 0.07 1.44 0.02 0.17 0.02 0.08 0.20
67-93 0.05 0.01 0.05 0.02 0.08 6.82 0.01 0.10 0.02 0.20 0.20 3.92 0.02 0.60 0.02 0.28 0.20
67-95 0.06 0.38 0.03 0.63 0.03 5.59 0.02 0.05 0.02 0.10 0.20
68-05 1.96 0.33 0.03 0.44 0.06 3.47 0.01 0.25 0.02 0.09 0.20
68-42 1.37 0.37 0.05 0.54 0.10 6.16 0.01 0.23 0.02 0.07 0.20
76-76 0.05 0.90 0.02 1.34 0.03 10.84 0.02 0.35 0.02 0.36 0.20 3.62 0.02 0.66 0.02 0.43 0.20
77-30 0.05 0.53 0.04 0.78 0.03 9.08 0.02 0.15 0.02 0.13 0.20 3.95 0.02 0.55 0.02 0.55 0.20
77-50 13.40 1.60 0.03 2.37 0.05 2.03 0.03 0.38 0.02 0.69 0.20
78-06 14.00 2.04 0.04 2.99 0.05 4.22 0.02 0.10 0.02 0.56 0.20
78-08 0.22 1.92 0.05 2.81 0.06 5.27 0.01 0.14 0.02 0.27 0.20 3.01 0.02 0.57 0.02 0.62 0.20
78-18 14.93 1.94 0.02 2.92 0.08 4.61 0.01 0.11 0.02 0.11 0.20 4.69 0.02 0.42 0.02 0.43 0.20
Mendon Formation
85-45 2.25 0.01 0.04 0.05 0.04 0.08 0.01 0.08 0.02 0.64 0.20
95-41 0.40 0.11 0.03 0.26 0.03 1.18 0.01 1.00 0.02 1.31 0.20
114-33 2.66 0.50 0.05 0.78 0.09 2.75 0.02 2.55 0.02 2.05 0.20
139-80 5.04 0.15 0.03 0.23 0.10 0.52 0.03 1.69 0.02 0.93 0.20
143-70 4.80 0.15 0.03 0.23 0.04 2.55 0.03 0.15 0.02 0.34 0.20
160-12 2.69 0.49 0.02 0.80 0.07 0.34 0.01 0.12 0.02 0.06 0.20
180-32 6.47 0.03 0.04 0.00 0.03 3.27 0.02 0.23 0.02 1.03 0.20
181-19 3.93 0.00 0.06 0.02 0.08 0.28 0.02 0.92 0.02 1.25 0.20
a
Iron concentration measured by MC-ICP-MS, after column chemistry.

255
256 V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266

Fig. 4. Variations in Fe/Al molar ratio, Fe (d56Fe in ‰ relative to IRMM-014) and S isotope compositions (d34S relative to V-CDT, and D33S
in ‰) as a function of depth in the drill-core for Mapepe (squares) and Mendon (circles) Formations. The red dotted line corresponds to the
erosional contact between Mendon and Mapepe Formations. The grey line on Fe/Al plot represents the average ratio in shales deposited in
oxic and suboxic environments (0.5; Taylor and McLennan, 1985; Lyons and Severmann, 2006). Other gray vertical lines represent
references at 0‰ for d56Fe, d34S and D33S, respectively. Error bars are smaller than symbols. (For interpretation of the references to color in
this figure legend, the reader is referred to the web version of this article.)

Fig. 5. Bulk rock Mn/Fe molar ratio versus Fe isotope composition (d56Fe), showing that Fe isotope compositions are controlled by
mineralogy. (A) The correlation labeled Fe-carb/Fe-ox refers to Fe-carbonate/Fe-oxyhydroxide in Mapepe samples with well-preserved, non-
hydrothermally altered signatures. The grey area represents samples (basaltic komatiites, black cherts, some bedded barite and terrigenous
and volcanoclastic sediments) with low Fe contents and hydrothermally modified d56Fe values. (B) Zoom on the hydrothermal-alteration
trend shown in A. Data from Mendon and Mapepe are represented as circles and squares, respectively. Error bars are smaller than symbols.

is dependent on the Al content of the initial rock, before mica. Other evidence in support of a strong fluid influence
hydrothermal alteration. In other words, the more a rock in these rocks was noted in previous studies. For instance,
was Al-rich, the more it is capable of incorporating K into multi-isotope dating (Pb-Pb, Sm-Nd and Rb-Sr) of the
V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266 257

Fig. 6. (A) Multiple S isotope composition of bulk-rock sulfides in samples from Mapepe (squares) and Mendon (circles) Formations. Sulfate
S-isotope data (open diamonds) are from Bao et al. (2007), Roerdink et al. (2012) and Montinaro et al. (2015). Data of sulfides disseminated in
bedded barites are from Roerdink et al. (2013). Error bars are smaller than symbols. The two arrows represent potential processes explaining
S isotope variations in sulfides from bedded barites, as well as terrigenous and volcanoclastic sediments: (1) microbial reduction from a sulfate
pool in seawater, followed by (2) a mixing with hydrothermally-derived juvenile sulfur (Roerdink et al., 2013; see Section 6.2. for more
details). (B) Iron versus sulfur isotope compositions for Mapepe and Mendon samples; symbols as in (A).

silicic carbonates from Onverwacht and Fig Tree Groups sive hydrothermal alteration process affecting both Mapepe
yielded an age of 2.7 Ga, interpreted as post- and Mendon sedimentary successions, a conclusion in good
depositional isotope resetting related to fluid circulation agreement with previous studies.
(Toulkeridis et al., 1998). In situ O and Si isotope measure- An important question to address is the potential impact
ments coupled to fluid inclusion analysis of chert from the of fluid overprint on the bulk Fe and S isotope composi-
Onverwacht Group were interpreted as a late-stage fluid- tions. Plotting Mn/Fe vs d56Fe can provide valuable insight
rock exchanges at temperature between 200 and 300 °C into this question. As shown in Fig. 5B (gray area), black
and depth of 1–4 km (Marin-Carbonne et al., 2011). In cherts and basaltic komatiites from Mendon display a
other volcanic rocks and cherts from the Barberton Green- strong correlation with five samples from Mapepe
stone Belt (Theespruit, Hooggenoeg and Kromberg Forma- (R2 = 0.85), including bedded barites (76-76, 77-30 and
tions), positive correlations were found between bulk rock 67-93), and terrigenous and volcanoclastic sediments (68-
SiO2 contents and O and Si isotopes, and were interpreted 42 and 45-99). This correlation likely reflects a fluid-
as evidence for pervasive silicification by large volumes of induced overprint as it includes various lithologies and
Si-rich fluids (Abraham et al., 2011). Silicification of these stratigraphic intervals. Importantly, the three bedded barite
volcanic rocks produced a depletion in elements commonly samples involved in this correlation have the lowest Fe con-
mobile during fluid-rock interaction (e.g. Co, Ni, Cu, Zn, tents (0.07–0.08 wt% Fe2O3). This suggests that they were
Na), but an enrichment in K, Rb and Ba (Hofmann and more sensitive to the effect of Fe-bearing fluids compared
Harris, 2008). Accordingly, the well-defined K-Al correla- to other barite samples, which have higher Fe contents
tion reported here (Fig. 3C) is interpreted to reflect a perva- (0.3–21 wt% Fe2O3). More specifically, we interpret the
258 V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266

fluid-influenced trend shown in Fig. 5B as a mixing between 6.2. Assessing primary mineral composition
Fe-bearing carbonates and sulfides, this suggestion being
again supported by petrographic observations. In these Although some of the rock samples from Mapepe and
fluid-influenced samples, both sulfides and carbonates were Mendon were strongly overprinted by hydrothermal fluids,
either precipitated or reset by isotope exchange during late the Fe isotope composition of several samples appears to
hydrothermal event. SEM analyses of sulfides in these sam- preserve a primary authigenic signature. The distinction
ples show that they contain negligible amount of Mn. Thus between both types of samples can be made from Fig. 5,
hydrothermally altered sulfides can be approximated from where the trend highlighted by the gray area (containing
the correlation plotted in Fig. 5B at Mn/Fe  0 and d56Fe basaltic komatiites, black cherts, some bedded barites as
of +0.46 ± 0.27‰. The Fe/Mn ratios of carbonate mea- well as terrigenous and volcanoclastic sediments) is inter-
sured by SEM analyses of hydrothermally altered samples preted as hydrothermally-modified d56Fe values (see discus-
show relatively large variations between 0.08 and 0.16. sion above in Section 6.1). In contrast, several samples from
From the trend in Fig. 5B, this Fe/Mn ratio corresponds the Mapepe Formation stand apart from this hydrothermal
to d56Fe values of Fe carbonates of 0.59 ± 0.35‰. Inter- trend and are suggested to record primary geochemical sig-
estingly, the Fe isotope fractionation between pyrite and nature from textural, petrological and geochemical evi-
Fe-carbonate (D56Fepyrite-carb) in these hydrothermally dence. For instance, ferruginous cherts (39-75, 50-25, 52-
altered samples is around 1.05‰, a value similar to that 72) and several terrigenous and volcanoclastic sediments
determined by density functional theory calculation for pyr- (56-30, 56-66, 68-05) display a linear correlation
ite and siderite at 300 °C (1.3‰; Blanchard et al., 2009). (R2 = 0.903) in a Mn/Fe vs d56Fe diagram (Fig. 5A), which
The good agreement between our measurements and data is clearly distinct from the hydrothermal alteration trend.
predicted from theoretical calculation supports our Petrographic observations indicate that Fe in ferruginous
conclusion that Fe isotope compositions of sulfides and cherts is carried by tiny crystals (<2 lm) of Fe oxides and
Fe-carbonates were modified by late fluid overprint at ankerite included in a cherty matrix. Interestingly, the rocks
300 °C. with the highest proportion of Fe oxides correspond to the
Sulfide S-isotope data of strongly hydrothermally- most positive d56Fe values. In contrast, iron in terrigenous
altered Mapepe samples (i.e., plotting in the gray area in and volcanoclastic sediments exclusively occurs in ankerite,
Mn/Fe vs d56Fe, Fig. 5B) show the largest range of d34S val- with no visible trace of Fe oxides. The linear correlation in
ues between 10.84 and 1.25‰ (Table 2) and lowest Fe Fig. 5A between Mn/Fe and d56Fe thus represents a mixing
and sulfide contents. This contrasts with the more restricted trend between Fe oxide and ankerite (noted Fe-carb in
range of d34S values (3.7 ± 2.4‰, 2r, n = 9) defined by Fig. 5A). Since Fe oxide contains limited amount of Mn rel-
other Mapepe samples. The Mapepe samples with low Fe ative to Fe, their Mn/Fe ratio can reasonably be considered
and sulfide contents could therefore have been modified to be 0. The d56Fe values of these oxides can then be
by late fluid overprint. However, this is inconsistent with derived from the linear correlation, providing a value of
the fact that their d34S vs D33S values define distinct isotopic +2.19 ± 0.69‰ (Fig. 5A). The positive d56Fe values in fer-
arrays related to rock lithology and no evidence for ruginous cherts are positively correlated with the SiO2 con-
hydrothermal influence such as isotope homogenization tent, thus suggesting that silicification protected primary Fe
(Fig. 6A). In other words, hydrothermally altered Mapepe oxides – and their Fe isotope composition – from reductive
samples identified from Mn/Fe-d56Fe relationships show, dissolution. Together with textural observations, this indi-
in a d34S-D33S diagram, the same general trends than other cates that silicification may have acted as an impermeable
samples, with strong petrological control. This suggests that barrier inhibiting cation and isotope exchanges relatively
S isotope compositions were not homogenized on large early in the rock history. The Fe isotope composition of
scale and preserved pristine signatures inherited from Fe-carbonate matching Fe-oxides in Fig. 5A can also be
depositional environment, a suggestion consistent with deduced from their Mn/Fe ratios. Considering a mean
Roerdink et al. (2013). If true, then it implies that sulfides Mn/Fe value of 0.21, it yields a d56Fecarbonate value of
in hydrothermally altered Mapepe samples did not 0.50 ± 0.52‰ (Fig. 5A). As detailed below, we suggest
precipitate during fluid overprint but rather experienced that this value reflects the primary signature of Fe carbon-
Fe isotope exchanges during fluid-rock interaction (while ates, either formed in the water column or, most likely, dur-
S isotopes remained unaffected). This may have been ing early diagenetic processes in the sediment.
possible for instance if hydrothermal fluids were S-free (or The primary vs hydrothermal origin of the carbonates
at least S-poor). Finally, mass-independent signatures, contained in Mapepe terrigenous and volcanoclastic sedi-
reflected by D33S values, are significantly different from ments has been the subject of discussion. From Sm-Nd,
0‰ for all Mendon and Mapepe samples (Fig. 6A). Rb-Sr and Pb-Pb isotopes systematic, Toulkeridis et al.
This indicates that the source of S input to the rocks (1998) suggested that Mapepe sedimentary rocks represent
was not strictly of mantle origin but rather experienced primary carbonates, which have experienced silicification
surface (atmospheric and oceanic) cycling before being and sericitization during a late hydrothermal event 2.7 Ga
reintroduced into the crust by fluid infiltration processes. ago. Yet part of the carbonates in Mapepe likely experi-
We thus emphasize that 33S-anomalies can be preserved in enced chemical and isotopic exchanges and possibly recrys-
sulfide minerals despite evidence of strong hydrothermal- tallization through hydrothermalism, while others may
alteration overprint. have been preserved, for instance due to early silicification
V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266 259

(Hofmann, 2005). In the present work, the two different (Roerdink et al., 2012, 2013; Philippot et al., 2012;
trends interpreted as reflecting primary and hydrothermal Montinaro et al., 2015; Muller et al., 2016). At least two
features (Fig. 5) are also supported by two other geochem- scenarios could explain the different D33S values observed
ical arguments. First, the partitioning behavior of Mn and in sulfides and barite. First, we could consider that barite
Fe in carbonates is temperature dependent. An experimen- formed in the water column and recorded dissolved sulfate
tal study demonstrated that Mn is preferentially incorpo- D33S values, while pyrite formed in a more complex path-
rated in carbonates relative to Fe with decreasing way including several S sources with distinct D33S values
temperature (Dromgoole and Walter, 1990). Considering (e.g. Roerdink et al., 2013). For instance, pyrite formation
a fluid of similar composition (either deep Archean seawa- could require sulfur addition to a pyrite precursor, such as
ter or hydrothermal fluid), high Mn/Fe are indeed mea- iron monosulfide (FeS, mackinawite) or iron polysulfide
sured among inferred primary carbonates, supposed to be (Fe3S4, greigite) (Rickard and Luther, 2007; Farquhar
precipitated at lower temperature. Second, bulk rock anal- et al., 2013; Marin-Carbonne et al., 2014). In the present
yses for Mapepe terrigenous and volcanoclastic sediments study, Mapepe samples show a large range of d34Spyrite val-
show a bimodal distribution interpreted as a mixture ues, from 10.84 to +1.44‰, and both negative and posi-
between Ba-poor primary (0.08, 0.35 and 3.19 wt% for sam- tive 33S-anomalies, thus indicating that different sources
ples 56-3, 56-66 and 68-05, respectively) and Ba-rich sec- of sulfur are involved in sulfide formation. From quadruple
ondary (i.e. hydrothermally-altered) minerals (12.82 and S isotope analyses by secondary ion mass spectrometry on
11.74 wt% for samples 45-99 and 68-42). Assuming Ba con- disseminated pyrite in barite beds from Mapepe, it was sug-
tent can be used as a proxy for the degree of hydrothermal gested that these pyrites resulted from a mixing between S
overprint, then the high and low Ba contents are in perfect derived from microbial sulfate reduction and juvenile man-
agreement with the conclusions deduced from the two tle sulfide (Roerdink et al., 2013). Our data on sulfides from
trends observed in Fig. 5. Furthermore, if all minerals were bedded barites and volcanoclastic and terrigenous sedi-
secondary, the bimodal distribution would not exist. ments fall on the same trend in a D33S-d34S diagram
Another primary signature is likely recorded in Fe-rich (Fig. 6A) and could thus be explained similarly by mixing
bedded barites. These rocks contain 18–21 wt% Fe2O3, various sources. Alternatively, a decoupling between barite
except one sample (78-08) with 0.3 wt% (Table 1). The and pyrite D33S values in Mapepe sediments could also be
main Fe-bearing mineral is pyrite, which generally occurs produced if pyrite formed through water column and sedi-
as thin layers (mm-scale) alternating with beds of barite mentary processes, while barite would result from atmo-
(Fig. 2C). These specific samples were selected for their spheric precursor. In a subaerial volcanic plume, barium
abundant pyrite layers with parallel and undeformed could react with sulfate aerosols to form microscopic barite,
structures, believed to be pristine sedimentary feature. In which then will remain insoluble and be transferred to sed-
a Mn/Fe-d56Fe diagram, these sulfides cluster around a iments (Muller et al., 2016). This could explain the system-
mean d56Fe value of 1.88 ± 0.38‰. This d56Fe value is atic MIF signature of Archean sulfate (Muller et al., 2016).
significantly different from the Fe sulfide signature of Finally, the highest D33S values of Mapepe are recorded in
+0.46 ± 0.27‰ associated with the hydrothermal- ferruginous cherts, with positive values up to +0.87‰
alteration trend, and could therefore reflect a pristine (Fig. 6A). If, as suggested above, early silicification pro-
d56Fesulfide composition. The particularly high Fe content tected Fe sulfides from late fluid-induced re-equilibration,
of these rocks makes Fe isotope resetting by late fluid then these strong MIF-S signatures should also be consid-
circulation unlikely. ered as primary. Roerdink et al. (2013) obtained similar sig-
The S isotope signatures determined in the present study natures in sulfides from cherts and proposed that positive
for sulfides and barite from Mapepe sediments are similar D33S values reflect sulfides derived from elemental sulfur
to those measured in previous studies (see Fig. 6; Bao metabolism (e.g. Philippot et al., 2007) when dissolved sul-
et al., 2007; Roerdink et al., 2012, 2013; Philippot et al., fate concentration were too low for microbial sulfate reduc-
2012; Montinaro et al., 2015; Muller et al., 2016). Like tion to occur. This conclusion is also compatible with our
for Fe isotopes, a pristine sulfide S-isotope composition is data, implying that microbial processes, either sulfates
expected for sulfide-rich bedded barites because these sul- and/or elemental sulfur reduction, possibly played an
fide layers are finely laminated and their high abundance important role in determining S isotope composition of
implies that their isotope compositions were hardly modi- the 3.2 Ga Mapepe sedimentary rocks.
fied by hydrothermal fluids. At first glance, a primary com-
position of these sulfides may appear incompatible with the 6.3. Implications for paleo-environmental conditions
observation that pyrite and barite in these samples do not
carry similar D33S values (Fig. 6). If pyrite and barite were The Fe isotope signature recorded in the most pristine
secondary, either formed or re-equilibrated, they would minerals (Fe oxides, carbonates and sulfides), deciphered
have the same D33S and D36S values, which is not the case. above, can be used to reconstruct paleo-redox conditions
Also barite samples have the typical negative D33S values in the Mapepe depositional environment. A schematic
inferred for Archean seawater sulfate (e.g. Bao et al., model is presented in Fig. 7. A well-accepted feature for
2007; Shen et al., 2009), and hydrothermalism would shift the Archean ocean is that deep waters were anoxic and fer-
their D33S towards higher values. It must be noted that bar- ruginous (e.g. Holland, 1984; Poulton and Canfield, 2011).
ite and sulfide D33S signatures could be pristine but not The main source of Fe to the ocean was hydrothermal sys-
identical because of different mechanisms of formation tems associated with submarine volcanic activity (Kump
260 V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266

Fig. 7. Schematic model illustrating Fe geochemical cycle in 3.2 Ga-old ferruginous ocean, recorded by the Mapepe sedimentary rocks. Fe
(II)aq is released from hydrothermal system to the ocean and diffuses from deep to shallow waters. In the photic zone, Fe(II)aq is oxidized to Fe
(III)aq and precipitates as Fe-(oxy)hydroxide. Iron precipitates sink in the water column down to the sediment, and are trapped as ferruginous
cherts. Iron carbonates are formed either in the water column, or more likely during sediment diagenesis from porewater buffered by high
concentration of Fe(II) (initially derived from hydrothermal system).

and Seyfried, 2005). Iron present as Fe(II) in minerals oxides suggest that O2 content in the photic zone was lower
forming the ocean crust was released to hydrothermal fluids than 104 lmol/L. This O2 content is particularly low and
during alteration processes under anoxic conditions and provides evidence for anoxic conditions at the time of fer-
injected to the ocean (Fig. 7). It is generally assumed that ruginous chert deposition, even in shallow water of the pho-
the isotopic composition of hydrothermal Fe(II) has not tic zone. The mechanism of Fe oxidation in Archean ocean
varied significantly over Earth’s history (Johnson et al., has been highly debated in the literature (see review in
2008) and has probably been similar to modern hydrother- Bekker et al., 2010). This may have resulted from either
mal vent systems with d56Fe values mostly between 0.5 (1) interaction with O2 produced by oxygenic photosynthe-
and 0‰ (Sharma et al., 2001; Beard et al., 2003; sis (Klein and Beukes, 1989), (2) anoxygenic photosynthesis
Severmann et al., 2004; Rouxel et al., 2008). Dissolved Fe using Fe(II) rather than water as electron donor and pro-
(II) accumulated in the deep ocean would then have dif- ducing Fe(III) rather than O2 (Konhauser et al., 2002;
fused to shallow water to be oxidized to Fe(III) in the pho- Kappler et al., 2005) or (3) photo-oxidation processes
tic zone. Fe(III) is insoluble at circumneutral pH and would (Braterman et al., 1983). The low O2 content determined
thus have precipitated as iron oxyhydroxides (Fe(OH)3). from the present data may be in favor of hypothesis 2 or
Iron oxyhydroxide particles would have sunk in the water 3 (for similar conclusions on other geological formations,
column before being incorporated into the sediment see Li et al., 2013; Czaja et al., 2013). However, oxidation
(Fig. 7). The highly positive d56Fe value recorded in Fe oxi- of Fe(II) by O2 (hypothesis 1) cannot be precluded since
des from Mapepe Formation (+2.2‰) is compatible with O2 could have been produced, used as an oxidizing agent
partial oxidation of aqueous Fe(II) with d56Fe  0‰, con- but not be accumulated in the environment due to higher
sidering the 2.5–4‰ isotope fractionation between oxides consumption relative to production. Another interesting
and aqueous Fe(II) (Johnson et al., 2002; Welch et al., result is that Fe isotope composition of carbonates from
2003; Anbar et al., 2005; Balci et al., 2006; Wu et al., Mapepe samples, interpreted to record primary signature
2011). Using dispersion-reaction modeling of the Fe(II) oxi- (i.e. d56Fecarb  0.50‰), is compatible both with a precip-
dation developed in previous studies (Czaja et al., 2012, itation from aqueous Fe(II) with d56Fe  0‰ (Fig. 7) and
2013; Li et al., 2013), we can estimate the concentration with isotope fractionation between Fe-carbonate and aque-
of O2 dissolved in the photic zone (Fig. 8). This model pre- ous Fe(II) of 0.26 to 0.70‰, as determined experimen-
dicts that under high O2 content in the photic zone (>1 lM) tally (Wiesli et al., 2004). If Fe-carbonates precipitated from
– i.e. redox stratified ocean – iron is completely oxidized at early diagenetic reactions, as suggested by earlier work in
the redox boundary, thus producing Fe oxide precipitates Archean rocks (Craddock and Dauphas, 2011; Dauphas
with d56FeFe(OH)3  0‰. In contrast, under low O2 content and Kasting, 2011; Johnson et al., 2013), this would imply
in the photic zone (<1 lM), Fe is only partially oxidized that deep ocean waters were Fe-rich and buffered sedimen-
and the isotope fractionation is expressed in Fe oxide par- tary pore waters in Mapepe depositional environment at
ticles, with the highest d56FeFe(OH)3 values corresponding d56Fe  0‰.
to the lowest dissolved O2 content (Fig. 8). The extremely Although the Fe isotope composition of ferruginous
positive d56Fe value of +2.2‰ estimated for Mapepe Fe cherts indicates that anoxic conditions prevailed in the
V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266 261

Fig. 8. Comparison between the Fe isotope composition of primary Fe oxides recorded in Mapepe ferruginous cherts (+2.2‰) and the
results of a dispersion-reaction modeling of the Fe(II) oxidation developed in previous studies (Czaja et al., 2012, 2013). The model was
calculated for two different depths of the water column (200 and 500 m) and a range of Fe isotope fractionation between Fe(II) and Fe(III),
from 2.5 to 4‰ (Wu et al., 2011). It shows that dissolved O2 content in the photic zone was lower than 104 lmol/L, thus demonstrating
anoxic conditions.

photic zone, this does not constrain directly the depositional nature would witness a modification of the Fe redox cycling
environment of bedded barites since they are not exactly in the water column, due to extensive Fe oxidation by O2
located on the same stratigraphic interval than cherts. production in shallow water (Rouxel et al., 2005). In mod-
Accordingly, one could consider the possibility of rapid ern analogues of Archean ocean, Fe oxidation at the
fluctuation of dissolved O2 concentration in the water col- chemocline leaves a residual dissolved Fe(II) enriched in
umn, with significant increase at the time of bedded barite light isotopes, due to precipitation of Fe(III) particles
formation and subsequent decrease during cherts deposi- enriched in heavy isotopes (Malinovsky et al., 2005;
tion. The only reliable Fe isotope signature of bedded barite Teutsch et al., 2009; Busigny et al., 2014). Dissolved sulfate
that can be interpreted in terms of paleo-environmental is reduced into sulfide, inducing Fe sulfide precipitation
conditions is carried by Fe sulfides, which show a mean (due to significant dissolved Fe concentration and low Fe
d56Fe value of 1.88‰. Three potential scenarios are gen- sulfide solubility). As a consequence, the light Fe isotope
erally proposed to explain negative d56Fe values of sedi- signature of the residual dissolved Fe(II) is trapped into sul-
mentary pyrites: (1) Dissimilatory Iron Reduction (DIR) fide precipitates, which is ultimately transferred to the sed-
during sediment diagenesis (Johnson et al., 2008; iments. Iron isotope composition of pyrite thus mirrors the
Heimann et al., 2010), (2) water column Fe oxidative light isotope signature inherited from the Fe oxidation cycle
cycling (Rouxel et al., 2005; Busigny et al., 2014), and (3) (Busigny et al., 2014). In the Mapepe Formation, the heavy
kinetic isotope fractionation associated with pyrite forma- Fe isotope composition recorded in Fe oxides from ferrug-
tion in a Fe-rich aqueous system (Guilbaud et al., 2011). inous cherts indicate a very limited Fe oxidation, with resid-
These three scenarios are examined below and compared ual dissolved Fe(II) remaining unchanged relative to mean
to data obtained on bedded barites from Mapepe Forma- ocean Fe derived from hydrothermal system (i.e. d56-
tion. In the case of DIR (scenario 1), partial reduction of Fe  0‰). This suggests that the negative d56Fe of Mapepe
Fe(III) oxide releases Fe(II) enriched in light isotopes (by pyrite does not record Fe redox variation related to an
3‰; Crosby et al., 2007) to sedimentary porewater increase in dissolved O2 content of shallow water, thus rul-
(Severmann et al., 2006). Subsequent pyrite precipitation ing out scenario 2. The third and last scenario to be consid-
from porewater may then record this light Fe isotope signa- ered is the possibility for a kinetic Fe isotope fractionation
ture (Johnson et al., 2008). An important assumption of associated with pyrite precipitation in iron-rich aqueous
this model is that Fe(III) reduction must be incomplete system. Experimental studies showed that a depletion of
since high -or complete- reduction would produce aqueous 2.2‰ in pyrite can be produced during formation from
Fe(II) with a d56Fe value similar to initial Fe(III) oxide. In aqueous Fe(II) (Guilbaud et al., 2011). This has never been
Mapepe bedded barites, we found no evidence for residual evidenced in any natural system but clearly requires more
Fe oxides. Additionally, if primary Fe oxides with d56Fe of specific study, where both pyrite and associated fluid can
2.2‰ (similar to those of cherts) would have been par- be collected. However, we note that this hypothesis is con-
tially reduced, aqueous Fe(II) should have been >0‰, sistent with a Fe(II) source of d56Fe  0‰ in the water col-
which is strongly different from the values observed for pyr- umn or sediment, as deduced from the primary Fe oxides
ites in bedded barites. Thus we conclude that the negative and carbonates in Mapepe samples. If correct, it implies
d56Fe values of Mapepe pyrite were not inherited from that the pool of Fe was dominating over S, allowing the
DIR microbial activity. In scenario 2, the pyrite isotope sig- expression of Fe isotope fractionation.
262 V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266

To summarize, the Fe isotope composition in Fe oxides volcanic rocks (Gamo et al., 1997; Kusakabe et al., 2000)
from ferruginous cherts provides indirect evidence for but the extent of this process in the Archean remains
anoxic conditions, with dissolved O2 concentration <104 - unknown (Holland, 1999; Bao et al., 2007). Instead, sulfate
lmol/L in the photic zone of Mapepe depositional environ- may derive from SO2-OCS photolysis in an O2-poor atmo-
ment. Anoxic conditions are also supported by pyrite d56Fe sphere, followed by hydrolysis of the chemical products
values and associated mineralogical observations. The sig- (Pavlov and Kasting, 2002; Huston and Logan, 2004). This
nificant mass-independent signature recorded in S isotopes may have been promoted by enhanced subaerial volcanism
is also indicative of highly reducing conditions in the atmo- specifically during this early Archean period (Philippot
sphere. Our conclusions contrast with those of a recent et al., 2012; Muller et al., 2016). Barium in seawater was
study coupling Fe isotope data and U concentrations in probably derived from hydrothermal alteration of mafic
the Manzimnyama Banded Iron Formations from the Fig and ultramafic rocks on the seafloor. The very low solubil-
Tree Group (Satkoski et al., 2015). In this previous study, ity of barite (Hannor, 2000) requires that Ba2+ and SO2 4
the authors proposed that the ocean was stratified, with cannot be transported together in solution. We propose
oxic surface waters and anoxic and ferruginous deep- that the oceanic basin was stratified, with sulfate-rich shal-
waters. The origin of the discrepancy between this previous low water and barium-rich deep water. The size of these
study and the results presented here is still unclear since the two layers could have been controlled by the fluxes of sub-
samples are from the same geological formation (i.e. Fig aerial volcanic activity (SO2 4 flux to shallow water) and
Tree Group) and are of similar ages (3.25–3.3 Ga). We hydrothermal circulation (Ba2+ flux to deep water). Barium
explored in detail the differences between the two sample sulfate would have been precipitated at the boundary
sets. Based on GPS coordinates, we calculated that the between these shallow and deep waters until one of the
drill-core site of the present work (core BBDP2, Latitude: reservoirs was exhausted. An alternative to this model
25°540 24.800 S, Longitude: 31°030 23.900 E) and the one of the would be that barium could be expelled by subaerial vol-
previous study (core BARB4, Latitude: 25°540 24.5500 S, Lon- canic plume of dacitic to felsic magma and react rapidly
gitude: 31°060 18.8200 E) are apart from more than 3 km. with sulfate aerosols in the atmosphere, producing
Local lateral variability of the sedimentary depositional microparticles of barite (see Muller et al., 2016). These
environment could thus potentially explain the discrepancy insoluble microparticles would be transferred and accumu-
between the two studies. For instance, ‘‘oxygen oases” with lated in sedimentary rocks. Following a model of enhanced
accumulation of O2 in proximal environments associated subaerial volcanism for explaining the increase in sulfate
with shallow water have been proposed for Archean times, concentrations in the basin, one would expect that the
while most of the ocean could have been anoxic and reduc- influx of elemental S should also have been increased. This
ing (Kasting, 1992; Olson et al., 2013). This hypothesis can- could have resulted in abundant pyrite with the opposite
not be tested from available data since there is no precise sign in D33S, but not necessarily in the same stratigraphic
dating for any of the two sedimentary sections. Moreover, level as the sulfate enrichments, in particular because trans-
no stratigraphic correlation can be made between the two port and preservation of these S isotope anomalies may be
cores due to strongly faulted blocks in the area and the fact decoupled, a feature very well-established for Archean sed-
that the contact between Mapepe and underlying Mendon imentary record (e.g. Farquhar and Wing, 2003). In the pre-
Formation is not conformable (e.g. Heinrichs and sent dataset, positive D33S values in pyrites were found in
Reimer, 1977). We thus conclude that lateral variability in one bedded barite sample from Mapepe, as well as in fer-
the depositional environment is a possibility for explaining ruginous cherts (Fig. 6A). In situ S isotope studies of
the differences in lithologies and geochemical models Mapepe pyrites also identified positive D33S values and neg-
derived in our study and the one of Satkoski et al. (2015). ative d34S in terrigenous and volcanoclastic sediments
A last possibility would lie in the assumptions made in (Philippot et al., 2012) and in some layered and massive
the case of the previous study by Satkoski et al. (2015), in pyrite bands of barite-rich samples (Roerdink et al.,
which Fe isotope composition of bulk rock was supposed 2013). Finally, it is interesting to note that barite with pos-
to represent primary Fe oxides. Although the presence of itive D33S matching the sulfides with negative values were
siderite (FeCO3) was noted by the authors, no detailed min- detected in several bedded barite beds, although in small
eralogical description of the samples was provided and the quantities (Muller et al., 2016).
proportion of siderite in Manzimnyama samples is
unknown. Assuming the presence of siderite in their sam- 7. CONCLUSION
ples, the isotopic composition of Fe oxides could have been
underestimated and the O2 concentration of the photic zone The occurrence of Early Archean barite deposits
overestimated. This possibility could be easily tested by requires high sulfate accumulation, thus raising the ques-
determining siderite content in Manzimnyama samples tion if free O2 was an important component of the atmo-
and compare the values to bulk Fe isotope compositions. sphere/ocean system about 1 billion years before the
The present findings also imply that high sulfate concen- Great Oxidation Event. The crux of this issue lies in the
tration in the ocean, necessary for bedded-barite deposi- degree and complexity of the modifications experienced
tion, was not produced by sustained oxic conditions in by the rocks during diagenesis, metamorphism and
the Mapepe paleo-environment. Part of the sulfate required hydrothermal alteration processes.
for BaSO4 precipitation could have been generated by mag- Here, we combined Fe and S isotope analyses and major
matic SO2 disproportionation during hydrothermalism of element chemistry with detailed petrological work in order
V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266 263

to distinguish primary signatures from post-depositional, Beard B. L., Johnson C. M., VonDamm K. L. and Poulson R. L.
fluid-induced, overprint. This allowed us to determine reli- (2003) Iron isotope constraints on Fe cycling and mass balance
able primary signature for Fe oxides, Fe carbonates and Fe in oxygenated Earth oceans. Geology 31, 629–632.
sulfides. A global model for Fe isotope geochemical cycle Bekker A., Holland H. D., Wang P. L., Rumble D., Stein H. J.,
Hannah J. L., Coetzee L. L. and Beukes N. J. (2004) Dating the
suggests that O2 concentrations in the photic zone must
rise of atmospheric oxygen. Nature 427, 117–120.
have been lower than 104 lmol/L during the chert deposi- Bekker A., Slack J. F., Planavsky N., Krapez B., Hofmann A.,
tion. The presence of mass-independent signature recorded Konhauser K. O. and Rouxel O. J. (2010) Iron Formation: the
by S isotopes in barite and sulfide supports that the atmo- sedimentary product of a complex interplay among mantle,
sphere was also anoxic, with O2 < 0.001% of the present tectonic, oceanic, and biospheric processes. Econ. Geol. 105,
atmospheric level (Pavlov and Kasting, 2002), thus indicat- 467–508.
ing global anoxic conditions of the Archean ocean at Braterman P. S., Cairns-Smith A. G. and Sloper R. W. (1983)
3.2 Ga. We conclude that barite deposit must have been Photooxidation of hydrated Fe2+: significance for banded iron
formed under anoxic conditions, probably related to a sul- formations. Nature 303, 163–164.
fate flux formed by photodissociation of abundant sub- Blanchard M., Poitrasson F., Méheut M., Lazzeri M., Mauri F.
and Baland E. (2009) Iron isotope fractionation between pyrite
aerial volcanic gases.
(FeS2), hematite (Fe2O3) and siderite (FeCO3): a first-principles
density functional theory study. Geochim. Cosmochim. Acta 73,
ACKNOWLEDGMENTS 6565–6578.
Busigny V., Planavsky N. J., Jézéquel D., Crowe S., Louvat P.,
Colleagues from the Isotope Geochemistry Laboratories in Moureau J., Viollier E. and Lyons T. W. (2014) Iron isotopes in
IPGP are thanked for fruitful discussions, particularly Wafa Abou- an Archean ocean analogue. Geochim. Cosmochim. Acta 133,
chami, Magali Bonifacie and Emilie Thomassot. Pascale Louvat 443–462.
and Julien Moureau are acknowledged for their technical assistance Byerly G. R., Kroner A., Lowe D. R., Todt W. and Walsh M. M.
for MC-ICP-MS analyses. Marc Quintin is thanked for making (1996) Prolonged magmatism and time constraints for sediment
thin sections of all samples. We highly appreciated the constructive deposition in the early Archean Barberton greenstone belt:
comments of four anonymous reviewers and the editor that helped evidence from the Upper Onverwacht and Fig Tree groups.
to improve the quality of the manuscript. This work was funded by Precamb. Res. 78, 125–138.
the National Program of Planetology (PNP) 2015 of the Institut Canfield D., Raiswell R., Westrich J., Reaves C. and Berner R.
National des Sciences de l’Univers and the UnivEarths Labex pro- (1986) The use of chromium reduction in the analysis of
gram at Sorbonne Paris Cité (ANR-10-LABX-0023 and ANR-11- reduced inorganic sulfur in sediments and shales. Chem. Geol.
IDEX-0005-02). 54, 149–155.
Carignan J., Hild P., Mevelle G., Morel J. and Yeghicheyan D.
(2001) Routine analyses of trace element in geological samples
using flow injection and low pressure on-line liquid chromatog-
REFERENCES raphy coupled to ICP-MS study of geochemical reference
materials BR, DR-N, UB-N, AN-G and GH. Geostand.
Abraham K., Hofmann A., Foley S. F., Cardinal D., Harris C., Newslett. 25, 187–198.
Barth M. G. and André L. (2011) Coupled silicon–oxygen Catling, D.C., 2014. The great oxidation event transition. In:
isotope fractionation traces Archaean silicification. Earth Holland, H.D., Turekian, K.K. (Eds.), Treatise on Geochem-
Planet. Sci. Lett. 301, 222–230. istry, second ed., vol. 6. Elsevier, Oxford, pp. 177–195.
Albarède F. and Beard B. L. (2004) Analytical methods for non- Craddock P. R. and Dauphas N. (2011) Iron and carbon isotope
traditional isotopes. In Geochemistry of Non-traditional Stable evidence for microbial iron respiration throughout the Archean.
Isotopes (eds. C. M. Johnson, B. L. Beard and F. Albarède). Earth Planet. Sci. Lett. 303, 121–132.
Mineralogical Society of America and Geochemical Society, Crosby H. A., Roden E. E., Johnson C. M. and Beard B. L. (2007)
Washington DC., pp. 113–152. The mechanisms of iron isotope fractionation produced during
Anbar A. D. and Knoll A. H. (2002) Proterozoic ocean chemistry dissimilatory Fe(III) reduction by Shewanella putrefaciens and
and evolution: a bioinorganic bridge? Science 297, 1137– Geobacter sulfurreducens. Geobiology 5, 169–189.
1142. Crowe S. A., Døssing L. N., Beukes N. J., Bau M., Kruger S. J.,
Anbar A. D., Jarzecki A. A. and Spiro T. G. (2005) Theoretical Frei R. and Canfield D. E. (2013) Atmospheric Oxygen 3.0
investigation of iron isotope fractionation between Fe(H2O)3+ 6 billion years ago. Nature 501, 535–538.
and Fe(H2O)2+ 6 : implications for iron stable isotope geochem- Czaja A. D., Johnson C. M., Roden E. E., Beard B. L., Voegelin A.
istry. Geochim. Cosmochim. Acta 69, 825–837. R., Nägler T. F., Beukes N. J. and Wille M. (2012) Evidence for
Anbar A. D., Duan Y., Lyons T. W., Arnold G. L., Kendall B., free oxygen in the Neoarchean ocean based on coupled iron-
Creaser R. A., Kaufman A. J., Gordon G. W., Scott C., Garvin molybdenum isotope fractionation. Geochim. Cosmochim. Acta
J. and Buick R. (2007) A whiff of oxygen before the Great 86, 118–137.
Oxidation Event? Science 317, 1903–1906. Czaja A. D., Johnson C. M., Beard B. L., Roden E. E., Li W. and
Arndt N. T. (1986) Differentiation of komatiite flows. J. Petrol. 27, Moorbath S. (2013) Biological Fe oxidation controlled depo-
279–301. sition of banded iron formation in the ca. 3770 Ma Isua
Balci N., Bullen T. D., Witte-Lien K., Shanks W. C., Motelica M. Supracrustal Belt (West Greenland). Earth Planet. Sci. Lett.
and Mandernack K. W. (2006) Iron isotope fractionation 363, 192–203.
during microbially stimulated Fe(II) oxidation and Fe(III) Dauphas N., Janney P. E., Mendybaev R. A., Wadhwa M., Richter
precipitation. Geochim. Cosmochim. Acta 70, 622–639. F., Davis A. M., van Zuilen M., Hines R. and Foley C. N.
Bao H., Rumble D. and Lowe D. R. (2007) The five stable isotope (2004) Chromatographic separation and multicollection-
compositions of Fig Tree barites: implications on sulphur cycle ICPMS analysis of iron – investigating mass-dependent and -
in ca. 3.2 Ga oceans. Geochim. Cosmochim. Acta 71, 4868–4879. independent isotope effects. Anal. Chem. 76, 5855–5863.
264 V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266

Dauphas N., van Zuilen M., Busigny V., Lepland A., Wadhwa M. tectonic, hydrothermal and surface processes during mid-
and Janney P. E. (2007) Iron isotope, major and trace element Archean times. Precamb. Res. 143, 23–49.
characterization of early Archean supracrustal rocks from SW Hofmann A. and Harris C. (2008) Silica alteration zones in the
Greenland: protolith identification and metamorphic overprint. Barberton greenstone belt: a window into subseafloor processes
Geochim. Cosmochim. Acta 71, 4745–4770. 3.5–3.3 Ga ago. Chem. Geol. 257, 224–242.
Dauphas N., Teng F.-Z. and Arndt N. T. (2010) Magnesium and Holland H. D. (1984) The Chemical Evolution of the Atmosphere
iron isotopes in 2.7 Ga Alexo komatiites: mantle signatures, no and Oceans. Princeton University Press, Princeton, p. 582.
evidence for Soret diffusion, and identification of diffusive Holland H. D. (1999) When did the Earth’s atmosphere become
transport in zoned olivine. Geochim. Cosmochim. Acta 74, oxic? A Reply. Geochem. News 100, 20–22.
3274–3291. Holland H. D. (2002) Volcanic gases, black smokers, and the Great
Dauphas N. and Kasting J. K. (2011) Low pCO2 in the pore water, Oxidation Event. Geochim. Cosmochim. Acta 66, 3811–3826.
not in the Archean air. Nature 474, E1–E3. http://dx.doi.org/ Homann M., Heubeck C., Airo A. and Tice M. M. (2015)
10.1038/nature09960. Morphological adaptations of 3.22 Ga-old tufted microbial
Dromgoole E. L. and Walter L. M. (1990) Iron and manganese mats to Archean coastal habitats (Moddies Group, Barberton
incorporation into calcite: effects of growth kinetics, tempera- Greenstone Belt, South Africa). Precamb. Res. 266, 47–64.
ture and solution chemistry. Chem. Geol. 81, 311–336. Homann M., Heubeck C., Bontognali T. R., Bouvier A.-S.,
Farber K., Dziggel A., Meyer F. M., Prochaska W., Hofmann A. Baumgartner L. P. and Airo A. (2016) Evidence for cavity-
and Harris C. (2015) Fluid inclusion analysis of silicified dwelling microbial life in 3.22 Ga tidal deposits. Geology 44,
Paleoarchean oceanic crust – a record of Archean seawater? 51–54.
Precamb. Res. 266, 150–164. Huston D. L. and Logan G. A. (2004) Barite, BIFs and bugs:
Farquhar J., Bao H. and Thiemens M. (2000) Atmospheric evidence for the evolution of the Earth’s early hydrosphere.
influence of Earth’s earliest sulfur cycle. Science 289, 756–758. Earth Planet. Sci. Lett. 220, 41–55.
Farquhar J., Savarino J., Airieau S. and Thiemens M. H. (2001) Johnson C. M., Skulan J. L., Beard B. L., Sun H., Nealson K. H.
Observation of wavelength sensitive mass-independent sulfur and Braterman P. S. (2002) Isotopic fractionation between Fe
isotope effects during SO2 photolysis: implications for the early (III) and Fe(II) in aqueous solutions. Earth Planet. Sci. Lett.
atmosphere. J. Geophys. Res. 106, 32829–32839. 195, 141–153.
Farquhar J. and Wing B. A. (2003) Multiple sulfur isotopes and the Johnson C. M., Beard B. L. and Roden E. E. (2008) The iron
evolution of the atmosphere. Earth Planet. Sci. Lett. 213, isotope fingerprints of redox and biogeochemical cycling in
1–13. modern and ancient Earth. Annu. Rev. Earth Planet. Sci. 36,
Farquhar J., Zerkle A. L. and Bekker A. (2011) Geological 457–493.
constraints on the origin of oxygenic photosynthesis. Photo- Johnson C. M., Ludois J., Beard B. L., Beukes N. J. and Heimann
synth. Res. 107, 11–36. A. (2013) Iron formation carbonates: paleoceanographic proxy
Farquhar J., Cliff J., Zerkle A. L., Kamyshny A., Poulton S. W., or recorder of microbial diagenesis? Geology 41, 1147–1150.
Claire M., Adams D. and Harms B. (2013) Pathways for Johnston D. T. (2011) Multiple sulfur isotopes and the evolution of
Neoarchean pyrite formation constrained by mass-independent Earth’s surface sulfur cycle. Earth-Sci. Rev. 106, 161–183.
sulfur isotopes. Proc. Natl. Acad. Sci. 110, 17638–17643. Kappler A., Pasquero C., Konhauser K. O. and Newman D. K.
Frei R., Gaucher C., Poulton S. W. and Canfield D. E. (2009) (2005) Deposition of banded iron formations by anoxygenic
Fluctuations in Precambrian atmospheric oxygenation recorded phototrophic Fe(II)-oxidizing bacteria. Geology 33, 865–868.
by chromium isotopes. Nature 461, 250–254. Kasting J. F. (1992) Models relating to Proterozoic atmospheric
Gamo T., Okamura K., Charlou J. L., Urabe T., Auzende J. M., and oceanic chemistry. In The Proterozoic Biosphere, A
Ishibashi J., Shitashima K. and Chiba H. (1997) Acidic and Multidisciplinary Study (eds. J. W. Schopf and C. Klein).
sulfate-rich hydrothermal fluids from the Manus back-arc Cambridge University Press, Cambridge, UK, pp. 1185–1187.
basin, Papua New Guinea. Geology 25, 139–142. Klein C. and Beukes N. J. (1989) Geochemistry and sedimentology
Guilbaud R., Butler I. B. and Ellam R. M. (2011) Abiotic pyrite of a facies transition from limestone to iron-formation depo-
formation produces a large Fe isotope fractionation. Science sition in the early Proterozoic Transvaal Supergroup, South
332, 1548–1551. Africa. Econ. Geol. 84, 1733–1774.
Hannor, J.S., 2000. Barite-celestine geochemistry and environ- Knauth L. P. and Lowe D. R. (2003) High Archean climatic
ments of formation. In: Alpers, C.N., Jambor, J.L., Nordstrom, temperatures inferred from oxygen isotope geochemistry of
D.K. (Eds.), Reviews in Mineralogy and Geochemistry: Sulfate cherts in the 3.5 Ga Swaziland Supergroup, South Africa. Geol.
Minerals. Crystallography, Geochemistry and Environmental Soc. Am. Bull. 115, 566–580.
Significance. Mineralogical Society of America, Washington Konhauser K. O., Hamade T., Raiswell R., Morris R. C., Ferris F.
DC, pp. 193–275. G., Southam G. and Canfield D. E. (2002) Could bacteria have
Heinrichs T. and Reimer T. O. (1977) A sedimentary barite deposit formed the Precambrian banded iron formations? Geology 30,
from the Archean Fig Tree Group of the Barberton Mountain 1079–1082.
Mand (South Africa). Econ. Geol. 72, 1426–1441. Konhauser K. O., Lalonde S. V., Planavsky N. J., Pecoits E.,
Heinrichs T. (1980) Lithostratigraphische Untersuchungen in der Lyons T. W., Mojzsis S. J., Rouxel O. J., Barley M. E., Rosiere
Fig Tree Gruppe des Barberton Greenstone Belt zwischen C., Fralick P. W., Kump L. R. and Bekker A. (2011) Aerobic
Umsoli und Lomati (Südafrika). Geottinger Arbeiten zur bacterial pyrite oxidation and acid rock drainage during the
Geologie und Paleaontologie 22, 118. Great Oxidation Event. Nature 478, 369–373.
Heimann A., Johnson C. M., Beard B. L., Valley J. W., Roden E. Kröner A., Byerly G. R. and Lowe D. R. (1991) Chronology of
E., Spicuzza M. J. and Beukes N. J. (2010) Fe, C, and O isotope early Archaean granite-greenstone evolution in the Barberton
compositions of banded iron formation carbonates demon- Mountain Land, South Africa, based on precise dating by single
strate a major role for dissimilatory iron reduction in 2.5 Ga zircon evaporation. Earth Planet. Sci. Lett. 103, 41–54.
marine environments. Earth Planet. Sci. Lett. 294, 8–18. Krull-Davatzes A. E., Lowe D. R. and Byerly G. R. (2012)
Hofmann A. (2005) The geochemistry of sedimentary rocks from Mineralogy and diagenesis of 3.24 Ga meteorite impact
the Fig Tree Group, Barberton greenstone belt: implications for spherules. Precamb. Res. 196, 128–148.
V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266 265

Kump L. R. and Seyfried W. E. (2005) Hydrothermal Fe fluxes atmosphere. Proc. Natl. Acad. Sci.. http://dx.doi.org/10.1073/
during the Precambrian: effect of low oceanic sulfate concen- pnas.1520522113.
trations and low hydrostatic pressure on the composition of Olson S. L., Kump L. R. and Kasting J. F. (2013) Quantifying the
black smokers. Earth Planet. Sci. Lett. 235, 654–662. areal extent and dissolved oxygen concentrations of Archean
Kusakabe M., Komoda Y., Takano B. and Abiko T. (2000) Sulfur oxygen oases. Chem. Geol. 362, 35–43.
isotopic effects in the disproportionation reaction of sulfur Pavlov A. A. and Kasting J. F. (2002) Mass-independent fraction-
dioxide in hydrothermal fluids: implications for the d 34S ation of sulfur isotopes in Archean sediments: strong evidence
variations of dissolved bisulfate and elemental sulfur from for an anoxic Archean atmosphere. Astrobiology 2, 27–41.
active crater lakes. J. Volcanol. Geoth. Res. 97, 287–307. Philippot P., van Zuilen M., Lepot K., Thomazo C., Farquhar J.
Labidi J., Cartigny P., Birck J. L., Assayag N. and Bourrand J. J. and Van Kranendonk M. J. (2007) Early Archean microorgan-
(2012) Determination of multiple sulfur isotopes in glasses: a isms preferred elemental sulfur, not sulfate. Science 317, 1534–
reappraisal of the MORB d 34S. Chem. Geol. 334, 189– 1537.
198. Philippot P., Van Kranendonk M., Van Zuilen M., Lepot L.,
Lalonde S. V. and Konhauser K. O. (2015) Benthic perspective on Rividi N., Teitler Y., Thomazo C., Blanc-Valleron M.-M.,
Earth’s oldest evidence for oxygenic photosynthesis. Proc. Natl. Rouchy J.-M., Grosch E. and de Wit M. (2009) Early traces of
Acad. Sci. 112, 995–1000. life investigations in drilling Archean hydrothermal and sedi-
Li W., Czaja A. D., Van Kranendonk M. J., Beard B. L., Roden E. mentary rocks of the Pilbara Craton, Western Australia and
E. and Johnson C. M. (2013) An anoxic, Fe(II)-rich, U-poor Barberton Greenstone Belt, South Africa. C.R. Palevol. 8, 649–
ocean 3.46 billion years ago. Geochim. Cosmochim. Acta 120, 663.
65–79. Philippot P., van Zuilen M. and Rollion-Bard C. (2012) Variations
Lowe D. R. (1994) Accretionary history of the Archean Barberton in atmospheric sulphur chemistry on early Earth linked to
Greenstone Belt (3.55-3.22 Ga), southern Africa. Geology 22, volcanic activity. Nat. Geosci. 5, 668–674.
1099–1102. Planavsky N. J., Asael D., Hofmann A., Reinhard C. T., Lalonde
Lowe D. R. and Nocita B. W. (1999) Foreland basin sedimentation S. V., Knudsen A., Wang X., Ossa Ossa. F., Pecoits E., Smith
in the Mapepe Formation, southern-facies Fig Tree Group. In A. J. B., Beukes N. J., Bekker A., Johnson T. M., Konhauser
Geologic Evolution of the Barberton Greenstone Belt, South K. O., Lyons T. W. and Rouxel O. J. (2014) Evidence for
Africa (eds. D. R. Lowe and G. R. Buerly). Geological Society oxygenic photosynthesis half a billion years before the Great
of America. Oxidation Event. Nat. Geosci. 7, 283–286.
Lowe D. R., Byerly G. R., Kyte F. T., Shukolyukov A., Asaro F. Poulton S. W. and Canfield D. E. (2011) Ferruginous conditions: a
and Krull A. (2003) Spherule beds 3.47-3.24 billion years old in dominant feature of the ocean through Earth’s history.
the Barberton Greenstone Belt, South Africa: a record of large Elements 7, 107–112.
meteorite impacts and their influence on early crustal and Rickard D. and Luther G. W. (2007) Chemistry of iron sulfides.
biological evolution. Astrobiology 3, 7–48. Chem. Rev. 107, 514–562.
Lyons T. W. and Severmann S. (2006) A critical look at iron Roerdink D. L., Mason P. R. D., Farquhar J. and Reimer T.
paleoredox proxies: new insights from modern euxinic marine (2012) Multiple sulphur isotopes in Paleoarchean barites
basins. Geochim. Cosmochim. Acta 70, 5698–5722. identify an important role for microbial sulfate reduction in
Lyons T. W., Reinhard C. T. and Planavsky N. J. (2014) The rise the early marine environment. Earth Planet. Sci. Lett. 331–332,
of oxygen in Earth’s early ocean and atmosphere. Nature 506, 177–186.
307–315. Roerdink D. L., Mason P. R. D., Whitehouse M. J. and Reimer T.
Lyons T. W., Fike D. A. and Zerkle A. (2015) Emerging (2013) High-resolution quadruple sulfur isotope analyses of 3.2
biogeochemical views of Earth’s ancient microbial worlds. Ga pyrite from the Barberton Greenstone Belt in South Africa
Elements 11, 415–421. reveal distinct environmental controls on sulfide isotopic arrays.
Malinovsky D. N., Rodyushkin I. V., Scherbakova E. P., Ponter Geochim. Cosmochim. Acta 117, 203–215.
C., Ohlander B. and Ingri J. (2005) Fractionation of Fe isotopes Rouxel O. J., Dobbek N., Ludden J. and Fouquet Y. (2003) Iron
as a result of redox processes in a basin. Geochem. Int. 43, 797– isotope fractionation during oceanic crust alteration. Chem.
803. Geol. 202, 155–182.
Manning C. (2006) Mobilizing aluminum in crustal and mantle Rouxel O. J., Bekker A. and Edwards K. J. (2005) Iron isotope
fluids. J. Geochem. Explor. 89, 251–253. constraints on the Archean and Paleoproterozoic ocean redox
Marin-Carbonne J., Chaussidon M., Boiron M.-C. and Robert F. state. Science 307, 1088–1091.
(2011) A combined in situ oxygen, silicon isotopic and fluid Rouxel O. J., Shanks W. C., Bach W. and Edwards K. J. (2008)
inclusion study of a chert sample from Onverwacht Group (3.35 Integrated Fe- and D-isotope study of seafloor hydrothermal
Ga, South Africa): new constraints on fluid circulation. Chem. vents at East Pacific Rise 9–10°N. Chem. Geol. 252, 214–227.
Geol. 286, 59–71. Rusk B. G., Lowers H. A. and Reed M. H. (2008) Trace elements
Marin-Carbonne J., Rollion-Bard C., Bekker A., Rouxel O., in hydrothermal quartz: relationships to cathodoluminescent
Agangi A., Cavalazzi B., Wohlgemuth-Ueberwasser C. C., textures and insight into vein formation. Geology 36, 547–550.
Hofmann A. and McKeegan K. D. (2014) Coupled Fe and S Satkoski A. M., Beukes N. J., Li W., Beard B. L. and Johnson C.
isotope variations in pyrite nodules from Archean shale. Earth M. (2015) A redox-stratified ocean 3.2 billion years ago. Earth
Planet. Sci. Lett. 392, 67–79. Planet. Sci. Lett. 430, 43–53.
Montinaro A., Strauss H., Mason P. R. D., Roerdink D., Münker Schoenberg R. and von Blanckenburg F. (2005) An assessment of
C., Schwarz-Schampera U., Arndt N. T., Farquhar J., Beukes the accuracy of stable Fe isotope ratio measurements on
N. J., Gutzmer J. and Peters M. (2015) Paleoarchean sulfur samples with organic and inorganic matrices by high-resolution
cycling: multiple sulfur isotope constraints from the Barberton multicollector ICP-MS. Int. J. Mass Spectrom. 242, 257–275.
Greenstone Belt, South Africa. Precamb. Res. 267, 311–322. Scott C. T., Bekker A., Reinhard C. T., Schnetger B., Krapez B.,
Muller E., Philippot P., Rollion-Bard C. and Cartigny P. (2016) Rumble D. and Lyons T. W. (2011) Late Archean euxinic
Multiple sulfur-isotope signatures in Archean sulfates and their conditions before the rise of atmospheric oxygen. Geology 39,
implications for the chemistry and dynamic of the early 119–122.
266 V. Busigny et al. / Geochimica et Cosmochimica Acta 210 (2017) 247–266

Severmann S., Johnson C. M., Beard B. L., German C. R., Ueno Y., Johnson M. S., Danielache S. O., Eskebjerg C., Pandey
Edmonds H. N., Chiba H. and Green D. R. H. (2004) The effect A. and Yoshida N. (2009) Geological sulfur isotopes indicate
of plume processes on the Fe isotope composition of hydrother- elevated OCS in the Archean atmosphere, solving faint young
mally derived Fe in the deep ocean as inferred from the sun paradox. Proc. Natl. Acad. Sci. 106, 14784–14789.
Rainbow vent site, Mid-Atlantic Ridge, 36°14’N. Earth Planet. Van Zuilen M. A., Philippot P., Whitehouse M. J. and Lepland A.
Sci. Lett. 225, 63–76. (2014) Sulfur isotope mass-independent fractionation in impact
Severmann S., Johnson C. M., Beard B. L. and McManus J. (2006) deposits of the 3.2 billion-year-old Mapepe Formation, Bar-
The effect of early diagenesis on the Fe isotope compositions of berton Greenstone Belt, South Africa. Geochim. Cosmochim.
porewaters and authigenic minerals in continental margin Acta 142, 429–441.
sediments. Geochim. Cosmochim. Acta 70, 2006–2022. Viljoen M. J. and Viljoen R. P. (1969) An introduction to the
Severmann S., Lyons T. W., Anbar A. D., McManus J. and geology of the Barberton granite-greenstone terrain. Geol. Soc.
Gordon G. (2008) Modern iron isotope perspective on the S. Afr. Spec. Publ. 2, 9–28.
benthic iron shuttle and the redox evolution of ancient oceans. Welch S. A., Beard B. L., Johnson C. M. and Braterman P. S.
Geology 36, 487–490. (2003) Kinetic and equilibrium Fe isotope fractionation
Sharma M., Polizzotto M. and Anbar A. D. (2001) Iron isotopes in between aqueous Fe(II) and Fe(III). Geochim. Cosmochim.
hot springs along the Juan de Fuca Ridge. Earth Planet. Sci. Acta 67, 4231–4250.
Lett. 194, 39–51. Weyer S. and Schwieters J. B. (2003) High precision Fe isotope
Shen Y., Farquhar J., Masterson A., Kaufman A. J. and Buick R. measurements with high mass resolution MC-ICPMS. Int. J.
(2009) Evaluating the role of microbial sulfate reduction in the Mass Spectrom. 226, 355–368.
early Archean using quadruple isotope systematics. Earth Wiesli R. A., Beard B. L. and Johnson C. M. (2004) Experimental
Planet. Sci. Lett. 279, 383–391. determination of Fe isotope fractionation between aqueous Fe
Staubwasser M., von Blanckenburg F. and Schoenberg R. (2006) (II), siderite and ‘‘green rust” in abiotic systems. Chem. Geol.
Iron isotopes in the early marine diagenetic iron cycle. Geology 211, 343–362.
34, 629–632. Wille M., Kramers J. D., Nägler T. F., Beukes N. J., Schröder S.,
Strelow F. W. E. (1980) Improved separation of iron from copper Meisel Th., Lacassie J. P. and Voegelin A. R. (2007) Evidence
and other elements by anion-exchange chromatography on a for a gradual rise of oxygen between 2.6 and 2.5 Ga from Mo
4% cross-linked resin with high concentrations of hydrochloric isotopes and Re-PGE signatures in shales. Geochim. Cos-
acid. Talanta 27, 727–732. mochim. Acta 71, 2417–2435.
Taylor S. R. and McLennan S. M. (1985) The Continental Crust: Wu L., Beard B., Roden E. E. and Johnson C. M. (2011) Stable
Its Composition and Evolution. Blackwell, Oxford. iron isotope fractionation between aqueous Fe(II) and hydrous
Taylor P. D. P., Maeck R. and De Bievre P. (1992) Determination ferric oxide. Environ. Sci. Technol. 45, 1847–1852.
of the absolute isotopic composition and atomic weight of a Xie X., Byerly G. R. and Ferrell, Jr., R. E. (1997) IIb trioctahedral
reference sample of natural iron. Int. J. Mass Spectrom. 121, chlorite from the Barberton greenstone belt: crystal structure
111–125. and rock composition constraints with implications to geother-
Teutsch N., Schmid M., Müller B., Halliday A. N., Bürgmann H. mometry. Contrib. Mineral. Petrol. 126, 275–291.
and Wehrli B. (2009) Large iron isotope fractionation at the Yamaguchi K. E., Johnson C. M., Beard B. L. and Ohmoto H.
oxic-anoxic boundary in Lake Nyos. Earth Planet. Sci. Lett. (2005) Biogeochemical cycling of iron in the Archean-Paleo-
285, 52–60. proterozoic Earth: Constraints from iron isotope variations in
Thomassot E., O’Neil J., Francis D., Cartigny P. and Wing B. A. sedimentary rocks from the Kaapvaal and Pilbara Cratons.
(2015) Atmospheric record in the Hadean Eon from multiple Chem. Geol. 218, 135–169.
sulfur isotope measurements in Nuvvuagittuq Greenstone Belt Yoshiya K., Nishizawa M., Sawaki Y., Ueno Y., Komiya T.,
(Nunavik, Quebec). Proc. Natl. Acad. Sci. 112, 707–712. Yamada K., Yoshida N., Hirata T., Wada H. and Maruyama
Thode H., Monster J. and Dunford H. (1961) Sulphur isotope S. (2012) In situ iron isotope analyses of pyrite and organic
geochemistry. Geochim. Cosmochim. Acta 25, 159–174. carbon isotope ratios in the Fortescue Group: metabolic
Tice M. M., Bostick B. C. and Lowe D. R. (2004) Thermal history variations of a Late Archean ecosystem. Precamb. Res. 212,
of the 3.5_3.2 Ga Onverwacht and Fig Tree Groups, Barberton 169–193.
greenstone belt, South Africa, inferred by Raman microspec- Zerkle A. L., Claire M. W., Domagal-Goldman S. D., Farquhar J.
troscopy of carbonaceous material. Geology 32, 37–40. and Poulton S. W. (2012) A bistable organic-rich atmosphere
Toulkeridis T., Goldstein S. L., Clauer N., Kröner A., Todt W. and on the Neoarchaean Earth. Nat. Geosci. 5, 359–363.
Schidlowski M. (1998) Sm–Nd, Rb–Sr and Pb–Pb dating of
silicic carbonates from the early Archaean Barberton green-
Associate editor: Janne Blichert-Toft
stone belt, South Africa. Evidence for post-depositional
isotopic resetting at low temperature. Precamb. Res. 92, 129–
144.

View publication stats

You might also like