Next Article in Journal
Formation of the Miaoan Au-Polymetallic Deposit in the Northern Taihang Mountain, North China Craton: Ore Geology, Geochronological and Geochemical Perspectives
Next Article in Special Issue
Involvement of Soil Microorganisms in C, N and P Transformations and Phytotoxicity in Soil from Post-Industrial Areas Treated with Chemical Industry Waste
Previous Article in Journal
CaCl2 as a Mineralizing Agent in Low-Temperature Recycling of Autoclaved Aerated Concrete: Cl-Immobilization by Formation of Chlorellestadite
Previous Article in Special Issue
Bioavailability of Colloidal Iron to Heterotrophic Bacteria in Sediments, and Effects on the Mobility of Colloid-Associated Metal(loid)s
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multiple Sulfur Isotope Evidence for Bacterial Sulfate Reduction and Sulfate Disproportionation Operated in Mesoarchaean Rocks of the Karelian Craton

by
Sergey V. Vysotskiy
1,
Tatyana A. Velivetskaya
1,*,
Aleksandr V. Ignatiev
1,
Aleksandr I. Slabunov
2 and
Anna V. Aseeva
1
1
Far East Geological Institute, Far Eastern Branch of the Russian Academy of Sciences, 690022 Vladivostok, Russia
2
Karelian Research Center, Institute of Geology, Russian Academy of Sciences, 185910 Petrozavodsk, Russia
*
Author to whom correspondence should be addressed.
Submission received: 1 July 2022 / Revised: 2 September 2022 / Accepted: 7 September 2022 / Published: 9 September 2022

Abstract

:
Sulfur isotope in sulfides from the Paleoarchean and the Neoarchean sedimentary rocks evidence microbial sulfur metabolism in Archean sulfur cycle. However, sulfur metabolism for the Mesoarchean interval is less obvious since evidence for a large range in sulfur isotope values has not yet been observed in Mesoarchean samples. We report the results of multiple sulfur isotope measurements for sulfide minerals from ~2.8 Ga sedimentary rocks in the southeastern part of the Karelian Craton. In situ isotope analysis of sulfide grains have been performed using a femtosecond laser-ablation fluorination method. Sulfide samples studied here yielded Δ33S values between −0.3 and +2.7‰ and δ34S values between −10 and +33‰. The Δ33S dataset was interpreted to indicate the incorporation of sulfur from two coexisting sulfur pools, photolytic sulfate and photolytically derived elemental sulfur. We suggest that the relative contributions of these Δ33S different pools to the pyritic sulfur could be controlled by the metabolic activity of coexisting sulfate-reducing and sulfur-disproportionating bacteria during pyrite formation. We therefore suggest the operation of different metabolic pathways of sulfur in Mesoarchean sedimentary environments.

1. Introduction

Bacteria are widespread prokaryotic organisms observed in different modern environments [1,2] and are thought to have already inhabited the Earth in the Paleoarchean [3]. Microbial activity in Paleoarchean (3.6–3.2 Ga) sedimentary environments has been identified on the basis of a combination of mass-dependent and mass-independent fractionation of sulfur isotopes (δ34S and Δ33S values) in sulfides from the Dresser Formation (3.49 Ga) in the Pilbara Craton, Western Australia [4,5,6,7,8,9] and from the Barberton Greenstone Belt, South Africa [5,10,11,12]. The strongly negative δ34S values recorded in sulfides from ancient sediments have been considered as one of the principal arguments for a biological genesis of sulfides [13,14], whereas reactions of thermochemical sulfate reduction have been proposed as an alternative for the process of pyrite formation with negative δ34S values in the Dresser Formation [15].
Studies of sulfide minerals from Neoarchean (2.8–2.5 Ga) rocks of Western Australia, South Africa and Brazil [16,17,18,19,20,21] have also shown large sulfur isotope fractionations (collectively, δ34S data published in these studies range from values of approximately −30‰ to +20‰), that supports the existence of microbial life during the Neoarchean.
In contrast to the relatively wide range of variability for δ34S values in Paleoarchean and Neoarchean sediments, sulfur isotope data that are currently available in the literature for Mesoarchean (3.2–2.8 Ga) sediments have shown a significantly smaller range in δ34S values. For example, Mesoarchean samples from southern Africa and Western Australia yielded δ34S values that predominantly grouped within a range from −3‰ to +7‰ [22,23,24]. Based on the lack of significant fractionations in 34S that is expected in microbially driven processes, Ono and coauthors [24] considered microbial sulfur cycling to be a minor factor for the transfer of sulfur species from the oceanic reservoir into sulfide sulfur in sedimentary rocks of the ~2.9 Ga Mozaan Group of the Pongola Supergroup in Southern Africa. The lack of evidence for large sulfur isotope fractionations has been linked to the limited bacterial sulfur cycling in ancient pyrite-forming environments [25]. However, despite the smaller δ34S variations in Mesoarchean compared to other Archaean intervals, there is still uncertainty with regard to the presence of microorganisms in Mesoarchean oceans since a narrow range in δ34S values in sulfides does not yet indicate the absence of biogenic transformations of sulfur compounds, but there is no evidence of active microbial activity, i.e., observations of significant fractionation in 34S for sedimentary rocks, to trace bacterial cycling of sulfur in the Mesoarchean. Thus, further research is needed to gain clarity on this issue.
There is now a unique opportunity to identify different microbially driven processes and sulfur cycling on the early Earth due to multiple sulfur isotope (32S, 33S, 34S, and 36S) measurements and the existence of mass-independently fractionated sulfur isotope signatures (non-zero Δ33S and Δ36S values) in sediments older than 2.4 Ga [26,27]. The presence of non-zero values for Δ33S and Δ36S in ancient sedimentary rocks may reflect photochemical gas-phase reactions of volcanogenic sulfur dioxide (SO2) in an anoxic atmosphere [28,29], which were responsible for the formation of a negative Δ33S-reservoir of sulfate and a positive Δ33S-reservoir of elemental sulfur [30]. The observations of negative and/or positive Δ33S values in Archean sedimentary pyrite may be helpful for identifying different sources of sulfur that would have been involved in microbially driven processes. Revealing the correlation between Δ33S and Δ36S values may be especially helpful to make a differentiation between mass-dependent and mass-independent processes [31].
In this contribution, we present a study of sulfur isotope composition (δ34S, Δ33S and Δ36S) of sulfides from the Mesoarchean volcanic-hosted massive sulfide occurrences in the southeastern part of the Karelian Craton by using laser-ablation sulfur isotope analyses. This study was performed in order to elucidate the sources of sulfur and the possible participation of bacteria in the transformation of sulfur to sulfides for occurrences of Mesoarchean age. Isotope data obtained in our work and their interpretation unequivocally indicate that microbially driven processes operated in Mesoarchean sedimentary environments, including two types of microbial sulfur metabolism–sulfate reduction and sulfur disproportionation.

2. Geological Setting

The Karelian Craton (Figure 1A) is the largest block of the Archean crust of the Fennoscandian (Baltic) Shield. It is formed by typical Archean granite–greenstone association and consist of granite–gneiss, greenstone, paragneiss and granulite complexes [32,33,34,35,36].
Three subprovinces are distinguished within the Craton: Vodlozero, Central Karelia and West Karelia, each possessing unique development characteristics during the Archean [35,37,38].
The Paleo- (3.3–3.1 Ga) and Meso-Archean tonalite-trondhjemite-granodiorite (TTG) gneisses and Mesoarchean (3.05–2.8 Ga) greenstone complex formed Vodlozero subprovince [34,42]. The Sumozero-Kenozero greenstone belt is located in the eastern part of it. It is one of the largest greenstone belts of the subprovince. It extends for ~400 km and is up to 50 km in width [34,39,41]. This greenstone belt is a system of the tectonic sheets and consists of slightly metamorphosed volcanic, volcanic-sedimentary, and intrusive rocks. Paleoproterozoic basaltic overlies the Archean sequence [35,41].
The Kamennoozero belt is one of the Sumozero-Kenozero greenstone belt fragments (Figure 1B). It includes two main units (tectonic sheets) with a total thickness of ~5 km [34,39]. The lower unit consists of pillow metabasalts and komatiite showing an affinity to oceanic plateau volcanism. The ages of this lower unit are 2892 ± 130 (by Pb-Pb) and 2916 ± 117 Ma (by Sm-Nd) [39]. The upper unit consists of volcanic basalt–andesite–dacite-rhyolite (BADR)-series rocks, bearing arc signatures, and adelite-series subvolcanic rhyolites. This volcanic rock age is 2875 ± 2 Ma (by U-Pb zircon TIMS) [39]. The unique feature of the upper unit is the presence of sulfide ores [40,43,44], which served as the object of this study.
At the end of the last century, over 10 sulfide deposits were explored in the Kamennoozero belt by exploratory drilling (see Figure 1B). The volcano-sedimentary massive sulfide (VSMS) Leksa deposit studied here is located in the southeastern part of the Kamennoozero belt and occurs in strata of quartz–albite–sericite and shungite-bearing schists [40].
The carbon-bearing schists are made of thin veinlets and lenses of quartz and ferruginous chlorite in a fine-grained matrix built by a mixture of chlorite, quartz, muscovite, and shungite. Quartz–albite–sericite schists are metasomatic rocks, presumably former volcanogenic-sedimentary arc-type rocks [40]. Their mineral composition varies from carbonate–quartz–albite–sericite–chlorite to quartz–pyrite. Carbonate is found almost everywhere, which is characteristic of carbon-dioxide metasomatism.
Sulfide ores constitute several superimposed ore bodies (Figure 2) that formed two main sulfide-bearing sequences totally up to 40 m thick. Host rocks near the ore body underwent an intensive silicification. Sulfides are represented mainly as pyrite forming different morphological shapes (cubic crystals, framboids, rounded globules, and layered concretions), rarer as pyrrhotite, chalcopyrite, sphalerite, and galena. Ores are impregnated, spotted, banded, massive, sometimes with veinlets, and of mainly pyritic composition 30%–50% of which are sulfides. The concentration of polymetals in ores is low (Zn 0.03%–0.24%; Au 0.003–0.06 ppm, Ag up to 0.02–4.00 ppm) [45]. The mineralization is characterized by being stratiform and associated with island-arc sedimentary-volcanogenic felsic rocks. Mafic lavas were not found by drilling, although field observations and the presence of fuchsite in metasomatites indicate their presence somewhere nearby.
Layered or banded, disseminated-banded and disseminated-streaky ore textures are usually typical for distal (distant from the hydrothermal discharge zone) facies of volcanogenic-sedimentary massive deposits [46]. Considering the established metal zonation at many VSMS deposits [47,48], the wells at the Leksa deposit possibly revealed rather low-temperature sulfide ore facies distant from the discharge hydrothermal zone. Based on the lithological classification of the host rock [49,50], we have classified this deposit as bimodal-mafic VSMS. Volcanic rocks of a VSMS deposit of this type are dominated by basalt, andesite, felsic lavas, and pyroclastic rocks. These deposits usually indicate hydrothermal fields related to island-arc volcanoes and backarc spreading centers [51].

3. Materials and Methods

Samples were collected from three drill cores: H1, H2, and H3. The cores intersect highly carbonaceous and carbonate–quartz–sericite shales. The choice of samples for the study was guided by the purpose of this research, which is to trace potential sulfur isotope signatures of Mesoarchean life. Therefore, sulfide samples were taken mostly from metasedimentary lenses in ore bodies of the Leksa deposit.
Two textural types of pre-metamorphic pyrite have been identified in these shales. Detailed sample descriptions are in Appendix, Figure A1, Figure A2, Figure A3, Figure A4 and Figure A5. One type is represented by aggregates of nano-micro crystals with different sizes, for which a generic term could be colloform-like pyrite. Colloform pyrite occurs as ellipsoidal aggregates composed of densely packed submicron-sized crystals and also as concentrically laminated grains. They are common in both clay-carbonaceous and carbonate shales. An example of concentrically laminated grains presented in Figure 3A: pyrite layers alternate with layers of marcasite; interstitial spaces of the marcasite crystals are filled with chlorite (chamosite), sericite (muscovite), quartz crystals and graphite. Note that since our study does not address issues related to the conditions favorable for the formation of pyrite or marcasite, in further discussion of the results we use the term pyrite to refer both to pyrite and its polymorph, marcasite.
Another type of pyrite (Figure 3B–D) studied here is represented by euhedral pyrite crystals. They commonly occur as individual microcrystals disseminated in the black carbonaceous shale (Figure 3B), or as thin layers composed of pyrite grains with a size of <100 μm. They also occur as aggregates of a few crystals with a size of ~0.5 mm that have overgrowth texture around concentrically laminated pyrite grains (Figure 3A). In some cases, pyrites were observed as ring-shaped and ball-shaped aggregates (typically less than 20 μm in diameter) of microcrystals, surrounding the inner core comprised of silicate minerals: quartz, chlorite, muscovite, or their mixture (Figure 3C,D).
From pyrite-bearing rock fragments of the drill-cores, polished chips (approximately 1–3 cm length/width and 0.5 cm deep) were prepared for both petrographic observation and isotope analyses. Samples used for isotope analysis were not subjected to additional processing, including etching in HNO3. All analyses were performed at Far East Geological Institute FEB RAS in Russia. More details on sulfide petrography are provided in the Appendix A.

3.1. Sulfur Isotope Analysis

Femtosecond laser-ablation fluorination method described in detail in [52,53] was applied for in situ measurements of δ33S and δ34S values in sulfide minerals. Briefly, an ultraviolet femtosecond laser ablation system (NWR Femtosecond UC with laser Pharos 2mJ-200-PP and harmonics module HE-4Hi-A, supplied by Electro Scientific Industries New Wave Research Division, Portland, OR, USA) was used to produce a laser crater of approximately 80 μm diameter and 40 μm depth in pyrite. Laser-generated aerosol sulfide particles were converted to SF6 (~12–13 nmol) via reaction with BrF5 at 350 °C. The SF6 was purified by means of a cryogenic trap system [53] and introduced into the ion source of a Thermo Scientific MAT 253 isotope ratio mass spectrometer using a gas injection interface developed and described in [52]. Sulfur isotope ratios were measured by monitoring the SF5+ ion currents at mass to charge ratios of m/z 127 (32SF5+), 128 (33SF5+), 129 (34SF5+), and 131(36SF5+).
It should be noted that it is not a problem if the size of the ablation pit is larger than the size of the analyzed sulfide grain because the presence of the matrix rock did not affect the accuracy and precision of analyses [53].
All sulfur isotope data are presented as conventional notation:
δ = R sample R standard 1 ,
where Rsample and Rstandard are the 34S/32S or 33S/32S ratios for the sample and standard, respectively, and reported relative to Vienna Cañon Diablo Troilite (V-CDT). For describing sulfur isotope anomalies (mass-independent sulfur isotope fractionations), capital delta notation was used which is defined as follows [54]:
Δ 33 S = δ 33 S δ 34 + 1 0.515 1
Δ 36 S = δ 36 S δ 34 + 1 1.90 1
Three International Atomic Energy Agency standards (IAEA S1, S2, and S3) were used to calibrate the SF6 reference gas and in-house standard. Analytical precision of ±0.2‰ (1σ) for δ34S values, ±0.03‰ (1σ) for Δ33S values, and ±0.27‰ (1σ) for Δ36S values were estimated from the long-term (approximately 1-year-long period) reproducibility of in-house standard [53].

3.2. Scanning Electron Microscopy and Energy Dispersion X-ray Spectroscopy Analysis

  • Scanning electron microscopy (SEM) was used to examine the detailed mineralogical and textural characteristics of sulfides on polished and some unpolished surface areas. Backscattered electron (BSE) and secondary electron (SE) imagery, as well as qualitative energy dispersive X-ray spectroscopy (EDS) analysis, was carried out on gold or carbon coated samples using a dual-beam TESCAN LYRA 3 XMH (Schottky cathode) Oxford AZtec Energy EDS system.
  • Electron imagery was performed at variable acceleration voltage (20–30 kV) and beam current (9–14 nA). Up to 30 kV and 14 nA was used for EDS analysis to ensure sufficiently high peak count rates for accurate determination of characteristic element-specific X-ray emission lines.
  • Quantitative X-ray spectroscopy spot analysis was performed using a JEOL JXA 8100 electron probe micro-analyser with three wave spectrometers and one energy dispersive spectrometers (Oxford Instruments Inca, Abingdon, UK), under a resolution of 137 eV MnKα, an accelerating voltage of 20 kV and a measure current of 1 × 10−8 A. Prior to analysis the samples were coated with a 20 nm carbon film. The beam was fully focused to give a spot size of about 1 μm with a measure current 1 × 10−8 A. Pure metals, glasses, and minerals analyzed through other methods were used as standards along with Oxford Instruments standards. Total Fe is equivalent to Fe+2 in calculations.
  • The qualitative phase analysis of sulfides was determined by X-ray diffraction in the powder by Rigaku MiniFlex II (Rigaku, Japan) (XRD) at the Laboratory of X-Ray Methods of the Analytical Center of the Far East Geological Institute, Far Eastern Branch of the Russian Academy of Sciences (FEGI FEB RAS) in Vladivostok.

4. Results

Sulfur isotope data are presented in Appendix A, Table A1 and displayed in Figure 4 as plots of δ34S vs. Δ33S values.
Colloform pyrite shows both negative and positive δ34S values; a total δ34S range is from −10.2‰ to +27.5‰ (Figure 4). In situ analyses revealed strongly heterogenous δ34S values, ranging from −6.7‰ to +27.5‰ in concentrically laminated pyrite grains, and relatively homogenous δ34S values that range from −10.2‰ to −9.5‰ and from −6.6‰ to +7.0‰ in non-laminated pyrite grains (Appendix A, Table A1). It is also evident that colloform pyrite studied here is mass-independently fractionated, with Δ33S values that mainly vary between −0.3‰ and +0.5‰. The exception is relatively large positive Δ33S = +1.4‰ measured in the core of the concentrically laminated pyrite (see Appendix A, Figure A4 and Figure A5).
The euhedral pyrite crystals have showed a wide spread of δ34S values and also revealed the presence of mass-independently fractionated sulfur. More specifically, disseminated pyrite exhibited intergranular heterogeneity for δ34S and Δ33S values that vary between −7.0‰ and +32.7‰ and between −0.5‰ and +1.5‰, respectively (Figure 4). These data are similar to those obtained for concentrically laminated pyrite grains. In contrast, euhedral pyrite overgrowths observed around colloform pyrite are relatively homogeneous in δ34S and Δ33S values that vary between +4.6‰ and +7.0‰ and between +1.5‰ and +2.7‰, respectively (Figure 4). Two analysis of euhedral pyrite observed as ring-shaped and ball-shaped aggregates of microcrystals have showed δ34S = +11.0‰ and +9.9‰, Δ33S = +2.7‰ and +1.7‰ (Table A1, Figure 4).
A subset of the euhedral pyrite samples was measured for 36S/32S isotope ratios to understand the relationship between Δ33S and Δ36S values. These pyrite grains have negative Δ36S values which range from −1.3 to −2.5‰ (Table A1), and also show a negative correlation between Δ33S and Δ36S with a slope of ≈−1.15 (Figure 5).

5. Discussion

As pointed out earlier in the introduction, the range of variability for δ34S values of sulfides from the Mesoarchean rocks represented by sedimentary units in southern Africa [22,23,24] and Western Australia [22] is significantly smaller compared to other Archaean intervals. Sulfur isotope data reported in this study, however, reveal a much wider range of δ34S values (−10.2‰ to +32.7‰) in the 2.8 Ga old rocks of the Karelian Craton (Figure 6). Figure 4 also illustrates one feature of the sulfur isotope record identified in our study, which is the presence of relatively high ∆33S values between +1.5‰ and +2.7‰ in the euhedral pyrite grains and the relatively low Δ33S values mostly between −0.30‰ and +0.44‰ for the colloform pyrite.
These observations aroused an obvious research interest regarding the possible pathways to differentiate the sulfur isotope ratios between these texturally distinct types of pyrite. Moreover, can the sulfur isotope data be explained in terms of known microbial processes and what these data tell us about the Mesoarchean microbial life? We suggest that considering what we know today about pyrite formation in modern sedimentary environments and sulfur isotope fractionation associated with microbial processes may help in answering these questions.
A possible mechanism of pyrite formation has been studied in several experimental works [55,56,57]. The formation of low-temperature sedimentary pyrite can occur in an anoxic environment and is mainly initiated by the reaction of dissolved H2S with reactive iron to yield non-crystalline monosulfide:
Fe 2 + + H 2 S FeS + 2 H +
The produced FeS is a transit product in the process of pyrite formation. There are two ways to transform FeS to pyrite: by the reaction of FeS with elemental sulfur
FeS + S 0 FeS 2 ,
or by the reaction of FeS with H2S
FeS + H 2 S FeS 2 + H 2 .

5.1. Sulfur Pathways Recorded by the Studied Colloform Pyrite

In our work it can be expected, as a first approximation, that dissolved H2S which is needed for pyrite formation in marine sediments could be produced primarily from sulfate dissolved in sea water by microbial metabolism that operated at the time of pyrite deposition. This is because the wide range from −6.7‰ to +27.5‰ and strong isotopic heterogeneity in δ34S values observed at micrometer scale in our sample of concentrically laminated pyrite grains (Figure A4, Appendix A) could be attributed to biological sulfate reduction, as the most likely mechanism to produce very large fractionation of sulfur isotopes [13].
We note that colloform pyrite studied here by in situ isotope analysis displays both positive (up ~1.4‰) and negative (−0.30‰) ∆33S values (see Figure 4). Distinct mass-independent sulfur isotope signals may indicate the transformations of sulfur species in the oxygen-poor atmosphere via photochemical reactions of volcanogenic SO2 [26]. These ultraviolet-driven reactions should, according to photochemical experiments [58,59,60,61,62,63,64,65], produce sulfate aerosols with a negative ∆33S signature and elemental sulfur aerosols with a positive ∆33S signature. As a result of SO2 atmospheric photochemistry, two compositionally and isotopically distinguished photolytic sulfur pools can be generated and maintained in an anoxic atmosphere. The absence of free molecular oxygen in the atmosphere prevents the oxidation of elemental sulfur to sulfate, and hence the homogenization of the mixing of different pools. An anoxic atmosphere favors the coexistence of various photolytic sulfur reservoirs and transfers the photolytic components out of the atmosphere to Earth’s surface. Once atmospherically derived elemental sulfur and sulfate rained out into aquatic environments, they can then be transformed into pyrite which inherits the isotope anomaly. Since the samples of colloform pyrite studied here archived opposite signs of the sulfur isotope anomaly, we can suggest photolytic seawater sulfate and photolytic elemental sulfur to be the principal sulfur source during precipitation of the pyrite grains with colloform texture.
According to the above consideration, isotope characteristics of colloform pyrite samples (i.e., non-zero ∆33S values, the broad range and also strong isotopic heterogeneity in δ34S values) may be associated with photolytic origin of its sulfur and would be consistent with forming pyrite via microbially mediated pathways, namely photolytic sulfate converted to H2S through microbial sulfate reduction. However, it needs to be ascertained if unusually high δ34S values of the studied pyrite can be associated solely with microbial sulfate reduction.

5.2. Colloform Pyrite—A Result of Microbial Sulfate Reduction Alone?

Based on current knowledge of sulfur isotope fractionations between sulfate and sulfide by bacterial sulfate reduction, the produced H2S is 32S-enriched vs. parent sulfate (i.e., δ34Ssulfide < δ34Ssulfate) [13,66,67,68,69,70]. If continued sulfate reduction takes place within a restricted system (in sediment pore waters), the concentration of sulfate in the system falls and Rayleigh fractionation occurs. As a result, the common sulfur isotopic trend observed for both remaining sulfate and pyrite formed is that δ34S values progressively increase as sulfate is consumed by sulfate-reducing microorganisms [71,72]. Therefore, the simplest explanation for the high-δ34S values (up to +27.5‰) observed in the concentrically laminated pyrite grains would be restricted-reservoir effects. In this situation, formation of these grains may take place along or below the sediment-water interface, not within the marine water column with a relatively large pool of available sulfate in comparison to sediment pore water, and thus Rayleigh fractionation is unlikely to occur. A sedimentary origin for the pyrite grains is also supported by preservation of their textural characteristics. Pyrite grains with concentric texture have been reported for Precambrian sedimentary successions and extensively reviewed [73]. This type of pyrite is interpreted to have formed in shallow-marine environments through accretion of pyrite mud [74,75], so the growing and aggregation of microcrystals from the inner core to the outer concentric layers can be explained by geometrical selection [76].
Concentrically laminated pyrite grains from sedimentary rocks of the Witwatersrand Basin in the central Kaapvaal Craton (South Africa) have been interpreted by Agangi and coauthors [77] to have formed at the water–sediment interface and their δ34S values that range from −12.2‰ to +12.6‰ have been considered as a result of sulfur isotope fractionation during sulfate reduction accompanied by Rayleigh fractionation in a closed system (at constant sulfate−sulfide fractionation). According to these authors, more than 80% of available sulfate must be consumed by sulfate reducers, which may lead to formation of pyrite with the highest (+12.6‰) δ34S values (under the assumption that an initial δ34S of the fluid ~−1‰). In our work, concentrically laminated pyrite grains from the Leksa deposit in the eastern part of Karelian Craton have been observed to have the δ34S values as high as +27.5‰, that is ~15‰ higher in comparison to Witwatersrand concentric pyrite [77]. Although one feasible explanation we propose for the strongly 34S-enriched sulfides is microbial processes, it remains unclear whether such isotopic enrichments can be achieved only through microbial sulfate reduction, even in restricted pore waters.
To calculate the magnitude of the 34S isotopic enrichment in sulfide during bacterial sulfate reduction undergoing Rayleigh-type distillation in a closed system, the equation from [78] was used. We made some assumptions regarding the initial conditions in our calculations. They are: (1) starting concentration of marine sulfate was taken as 200 μM, following the conclusions of [79] that “oceanic Archean sulfate concentrations were <200 μM”, although much lower sulfate levels ≈80 μM [80] and even <10 μM are also possible [21,81]; (2) the initial marine water sulfate had a δ34S value of ~0‰ [82]; (3) the fractionation associated with sulfate reduction was taken as 10‰ (i.e., δ34Ssulfate − δ34Ssulfide ≈10‰), since the maximum depletion of 34S into pyrite observed in our study is ~−10‰. If other mechanisms did not operate, the calculations show that more than 98% of the total sulfate in the system should be consumed by bacteria to explain pyrite highly enriched in 34S up to ~ +27‰. In this case the residual sulfate concentration may expect to be <4 μM, which is too low to enable bacterial sulfate reduction because threshold concentrations for sulfate uptake have been detected at ~5 to 20 μM sulfate [83]. Thus, the current model does not provide an explanation for the high δ34S values measured in concentrically laminated pyrite grains.
To elucidate the possible reasons for the formation of highly 34S-enriched pyrite, other cases can be considered. It is possible that the sulfur isotopic composition of Archean seawater sulfate could be higher than 0‰. Then a Rayleigh distillation model will require the δ34S value of seawater sulfate to be not less than ~+10‰ to generate pyrite with δ34S ≈ +27‰ upon removal of 85% of sulfate by sulfate-reducing microorganisms in the hypothetically closed system. These considerations are in line with [84] that the δ34S of Archean seawater sulfate may be in the range of +6‰ to +16‰. However, this is not the case for our samples since our model calculations with large sulfur isotope fractionations between hydrogen sulfide and sulfate of more than 25% should imply high sulfate concentrations of the Mesorchean ocean, but there exists no evidence to support this resulting deduction.

5.3. Colloform Pyrite: Evidence of Sulfur Disproportionation Accompanying Sulfate Reduction

Here we demonstrate that observed high δ34S values in colloform pyrite can be explained by mixing two different pools of sulfate available simultaneously for sulfate-reducing bacteria at the site of pyrite formation. One of these pools was dissolved atmospherically derived photolytic sulfate, and the other pool was biologically produced sulfate, namely as a result of microbial disproportionation of atmospherically derived insoluble elemental sulfur. Microbial elemental sulfur disproportionation in marine sediments served as the source of sulfate and hydrogen sulfide in the pore water [85,86]:
4 S 0 + 4 H 2 O 3 H 2 S + SO 4 2 + 2 H +
It was shown that bacterial sulfur disproportionation results in generating sulfide that is depleted in 34S only up to 7‰ while sulfate is enriched in 34S up to 30‰ relative to starting sulfur [86,87,88,89,90,91,92]. Then, the combination of microbial sulfur disproportionation and microbial sulfate reduction can be proposed to explain highly positive δ34S values of pyrite studied here. If sulfate reduction takes place and sulfate derived from elemental sulfur is continuously added to the reservoir, pyrite produced from the sulfate mixture can reach a high positive δ34S value even before the concentration of available sulfate in the pore water drops to a threshold of ~5 to 20 μM [83] and pyrite formation ceases. This interpretation suggests that pyrite was formed in sediments high in available elemental sulfur for microbial processes where elemental sulfur delivered from the atmosphere consistently fell to the bottom and accumulated in sediments, forming the elemental sulfur pool. Metabolism of elemental sulfur by bacteria resulted in the formation and maintenance of a biologically derived sulfate pool, which mixed with the atmospherically derived sulfate pool that contained within pore waters of the sediment. This process was accompanied by the metabolism of sulfate sulfur through sulfate-reducing bacteria to form pyrite. If so, disproportionation of atmospherically derived elemental sulfur in the sediment would act as a source of additional sulfate in the processes of continuous loss of atmospherically derived sulfate by bacteria out of the pore waters in the closed system. This would be running into the changes in δ34S values that can be observed in different pyrite grains because their δ34S values should be related directly to the concentration of available biologically and atmospherically derived sulfate in pore water. Pyrite exhibiting negative δ34S values would be expected if a greater proportion of total sulfate reservoir would be atmospherically derived sulfate with δ34S values of ~0‰. Pyrite exhibiting relatively high δ34S values would be apparent if the δ34S values of sulfate reservoir would have been locally increased due to metabolism by sulfate-reducing bacteria, where the loss of the sulfate in a closed system would be replenished from a pool of atmospherically derived elemental sulfur through bacterial metabolism of elemental sulfur.
It should be noted that a combination of sulfate reduction and sulfur disproportionation has been proposed to explain strong 34S depletions in sedimentary sulfides [86,93]. Indeed, these processes may be responsible for formation of depleted rather than enriched 34S sulfides, but only if the mechanism of sulfide oxidation into elemental sulfur is involved [86,93], implying the presence of an oxidizing environment. We consider such an environment unlikely to exist at the time of deposition of pyrite studied here, because it is inconsistent with the observed significant mass-independent isotopic signals in pyrite, generation and preservation of which requires an anoxic depositional environment [84].
If our assumption linking formation of colloform pyrite and two different metabolic pathways of sulfur is true, colloform pyrite should also be highly inhomogeneous at micrometer scale in the δ34S values. This is because the magnitude sulfur isotope fractionation induced by bacteria, and hence pyrite δ34S value, depends on microbial sulfate reduction rates, which in turn can be related directly to the concentration of available dissolved sulfate as well as organic matter. Therefore, in microenvironments where several biologically mediated mechanisms operate simultaneously, produced pyrite can be remarkably heterogeneous in δ34S, which was observed for concentrically laminated pyrite grains studied here.
Proceeding from our assumption of a biogenic origin for pyrite genesis, the presence of two different metabolic pathways should be responsible not only for the local variations in δ34S values, but also for Δ33S values, which is the case for the samples of colloform pyrite. Concentrically laminated pyrite grains are characterized by a heterogeneous distribution of nonzero Δ33S values as well as the presence of zero Δ33S values.
The observed at micrometer scale changes in Δ33S values of pyrite grains cannot be explained by microbial activity and associated Rayleigh fractionation of sulfur isotopes in restricted pore waters. It is because bacterial sulfate reduction and sulfur disproportionation cannot induce sulfur isotope mass-independent fractionation; the transformation of sulfur by these processes cannot significantly change Δ33S of sulfur reservoirs. In other words, as bacterial sulfate reduction with Rayleigh distillation progressively occurs within a Δ33S-bearing reservoir, the Δ33S magnitude of the remaining sulfate reservoir remains unchanged and, thus, pyrite formed should be homogeneous in Δ33S values and identical to the Δ33S reservoir.
Strong heterogeneity in Δ33S at the micrometer scale that we observe for colloform pyrite can be explained by relatively rapid alteration of Δ33S for sulfur reservoir due to mixing in pore waters at least two sulfate reservoirs that should carry isotopic distinguishing Δ33S signatures. In the present study it is suggested that the negative Δ33S reservoir of sulfate available for use by sulfate-reducing microorganisms in pore water could have been diluted by positive Δ33S sulfate, which was produced through bacterial sulfur disproportionation of atmospherically derived elemental sulfur accumulated in sediments. In this case, the relative bacterial sulfur disproportionation rates would locally influence and alter the integrated Δ33S signature of dissolved sulfate in pore waters. Thereby a change in the Δ33S value of local sulfate reservoir gives rise to a change in the pyrite from negative to positive Δ33S values, including zero Δ33S values. Microbial activity can be seen as a mixer if the amount of elemental sulfur was not a limiting factor for pyrite formation. This means that the presence of Δ33S values close to zero within concentrically laminated pyrite grains studied here does not necessitate an origin of sulfur from magmatic inputs but may result from mixing of two Δ33S-distinct sulfate reservoirs. It is conceivable that the simultaneous operation of two different metabolic pathways led to the erasing of the atmospheric signature in the pyrite studied here.
An additional issue with the existence of mass-independent isotopic signals in our samples arises: whether the observed Δ33S values in the sample of colloform pyrite are attributable to the photolytic origin or it may be due to kinetic sulfur isotope fractionation associated with metabolic pathways.
According to experimental and modeling studies, the positive difference between Δ33S values of sulfide and sulfate up to 0.2‰ are produced by microbial sulfate reduction (if only large isotope fractionations in δ34S of >30‰) and slightly negative or positive shifts in Δ33S values are associated with microbial sulfur disproportionation [31,92,94,95,96]. In our work isotopic data obtained for the grains with concentric texture are very difficult to associate with a purely kinetic sulfur isotope fractionation, since the magnitude detected for Δ33S values is too large (>0.3‰) to account for the shift in Δ33S due to microbial sulfate reduction. According to the above discussion, microbial sulfate reduction in combination with microbial sulfur disproportionation can explain both the highest δ34S values (+27‰) and strong variability of the δ34S and Δ33S values measured in concentrically laminated pyrite grains at the micrometer scale.

5.4. Euhedral Pyrite: Evidence of Sulfur Disproportionation Accompanying Sulfate Reduction

The hypotheses presented here make it possible to understand what the reason is for a drastic increase in the Δ33S values of up to +2.7‰ in euhedral pyrite grains developed around the colloform pyrite (see Figure 4) and why these euhedral grains are so homogeneous in both δ34S and Δ33S values. These questions deserve special consideration because euhedral pyrites were formed later than the colloform pyrite but yield higher Δ33S values compared to colloform pyrite formed at an early stage of diagenesis. It would seem that an additional source of sulfur carrying a positive signal appeared in the system, for example, due to the opening of the system during late diagenesis. This scenario assumes infiltration of atmospherically derived sulfur carrying Δ33S > 0‰ towards the site of pyrite formation, which is unlikely because elemental sulfur is an insoluble component of sediments. Instead of this we consider a closed system where the amount of atmospherically derived elemental sulfur being accumulated in the sediment was sufficient to form pyrite. Very distinct multiple sulfur isotopic signals of euhedral and colloform pyrite would simply be the result of the exhaustion of atmospherically derived sulfate within pore waters at the time when formation of euhedral pyrite occurred within sediments.
Based on the formation mechanism of early and late diagenetic pyrites, the sequence of processes that may be responsible for the isotopic features observed in different pyrite generations is proposed as follows:
(1) During early diagenesis, the main mechanism of pyrite formation with concentric texture could, as considered above, be closely related to the transformation of photolytic sulfates and elemental sulfur to pyritic sulfur by microbial processes, and therefore these pyrite grains showed the large range of variability for both δ34S and Δ33S values;
(2) As diagenesis continues during burial, diffusion of dissolved sulfate into marine sediments from overlying water decreases thereby limiting sulfate reducers activity. In this situation, the residual photolytic elemental sulfur could be a dominant source for the formation of pyrite via reactions (R1−R3). Pyrite crystallization by this mechanism occurs only as aggregates of microcrystals ~1 μm in diameter [55];
(3) With the burial of sediment, the processes of recrystallization and overgrowth of the microcrystals can result in the formation of large euhedral pyrites [97,98,99]. Accordingly, the euhedral pyrites issued by these processes were observed around the concentrically laminated pyrite grains. The remarkable intergranular and intragranular homogeneity in both δ34S and Δ33S values as well as large positive magnitude of Δ33S values inherited from photolytic elemental sulfur are reasonably expected in these pyrites.
It is worthwhile saying that the strongly positive Δ33S observed here for euhedral pyrite grains undoubtedly reflects the presence of mass-independently fractionated sulfur. Atmospheric origin of this sulfur was proven by including Δ36S in the analysis. It was revealed that the values of Δ33S linearly correlate with Δ36S with a slope of ca. −1.15 (Figure 5), indicating an atmospheric source because the obtained slope differs from a slope of −6.85 (Figure A6), produced by mass-dependent microbially mediated processes [31], but is in good agreement with the Archean trend with a slope of −0.9 produced by mass-independently fractionated sulfur isotopes in Archean rocks.
Besides the concentrically laminated pyrite studied here, the single euhedral crystals or aggregates of a few crystals developed in fine-grained sedimentary rocks can also clearly indicate that sulfate-reducing and sulfur-disproportionating bacteria could have played an essential role at the time of deposition of these grains. This is because the large range of δ34S values (−7.0‰ to +32.7‰) observed for the disseminated pyrite grains (see Figure 4) is consistent with the range in δ34S variability from concentrically laminated pyrite (−6.7‰ to +24.5‰), and therefore could be explained by similar arguments that the formation of pyrite sulfur was closely related to the microbial processes of sulfate reduction, sulfur disproportionation, or a combination of both. The preservation of both positive and negative Δ33S values in the euhedral pyrite crystals also implies the presence of at least two sources of sulfur for the formation of pyrite, that is, atmospherically derived sulfate carrying Δ33S < 0‰ and elemental sulfur carrying Δ33S > 0‰. An additional point to emphasize is that most of the disseminated pyrite grains analyzed in this work have moderate Δ33S values ranging roughly between −0.3‰ and 0.5‰ (Figure 4). The exception is the cluster of crystals that form the ring-shaped or tubular structures (see Figure 3); they showed distinctively higher Δ33S values of 1.65‰ and 2.65‰ (Figure 4). We associate these structures with the fossilization of microorganisms. Similar structures considered as evidence of microfossils have been described in Paleorchaean rocks of Western Australia [100], in Mesorchaean VSMS deposit in the Pilbara of Western Australia [101], in the Witwatersrand System of South Africa [102] and in Phanerozoic sediments [103]. An exhaustive review of microstructures that have been considered as Archean microfossilsis has recently been published by [104]. Experimental studies have shown that macroscopic structures of bacteria may be preserved by pyritization. This is shown by the presence of amorphous FeS on both the inner and outer surfaces of sulfate-reducing bacteria and by the immobilization of FeS around the bacterial microcolony [105]. Once FeS is formed it can transform into pyrite by the reaction of FeS with H2S or S0 (reactions R2-R3).
Based on the fact that Δ33S values measured for pyritized fossils in this study were as high as +2.65‰, we are certain that elemental sulfur was the dominant source for the processes that are responsible for the pyritization of microorganisms. This, of course, does not mean that pyritized fossils were necessarily disproportionating microorganisms, but may be considered in favor of the latter.

5.5. General Insights

In summary, the mutually complementary facts considered in this study as evidence of the integrated activity of sulfate-reducing and sulfur-disproportionating bacteria rather than sulfate reducers alone at the time of pyrite formation are the following:
(1) The similarity of the isotopic composition of microfossils (average δ34S ≈ +10‰ and Δ33S ≈ +2‰) to that measured both in the core of concentrically laminated pyrite (δ34S ≈ +7.8‰ and Δ33S ≈ +1.4‰) and in the euhedral pyrite surrounding it (average δ34S ≈ +5‰ and Δ33S ≈ +1.8‰);
(2) The similar range of δ34S variations, spanning between −7.0‰ and +32.7‰, for both dispersed and concentrically laminated pyrite that is too large to be interpreted as microbial sulfate reduction.
The importance of microbial sulfur disproportionation for early Archean pyrite formation is also evidenced by the multiple sulfur isotope data of Philippot and coauthors [4]. They suggested that the microscopic sulfides with negative δ34S and positive Δ33S values in barite deposit (Dresser Formation, Western Australia) are likely due to the activity at this time of the coexisting sulfur-disproportionating, sulfur-reducing and sulfate-reducing microorganisms. However, an alternative origin for this negative relationship between Δ33S and δ34S in microscopic pyrite has been proposed by Bao and his team [106]; that is, the mixing scenario where sulfides derived from S0 (produced by the 193 nm SO2 photolysis) and sulfides derived from seawater sulfate could be mixed in different proportions. Moreover, it is still possible that that the puzzle of isotopic effects will fit if sulfur-disproportionating bacteria that can also reduce sulfate were involved in the generation of the micropyrites [107].
Although the fact of the existence of microorganisms in the Archaean is well established today, less is known about the communities of bacteria which can utilize sulfur species for their metabolism in an Archaean environment. If the isotope data for sedimentary pyrite studied here indicate the coexistence of sulfur-disproportionating and sulfate-reducing microorganisms in the Mesoarchean, the sedimentary circumstances should be expected to favor the growth of both groups of these organisms. It is important that bacterial sulfur disproportionation is more sensitive to the hydrogen sulfide concentrations than bacterial sulfate reduction. In experimental studies [87,108] it was evidenced that utilization of elemental sulfur by disproportionating bacteria occurs only in the presence of a hydrogen sulfide scavenger because disproportionation of elemental sulfur is inhibited as the hydrogen sulfide concentrations raised >1 mM. In early anoxic oceanic environments with abundant Fe2+ in the Archean oceans [109], the presence of Fe2+ in pore water could act as a scavenger, buffering the hydrogen sulfide concentrations in pore water to low levels, even with rapid rates of bacterial sulfate reduction. Shallow lagoons could, according to [106], be favorable habitats for sulfur disporportionating bacteria before the rise of atmospheric oxygen. In addition, the mineral textures of pyrites studied here supports the assumption that the pyrite grains would have been formed in shallow water conditions in a zone away from the area where hydrothermal fluids vented onto the sea floor and where temperatures would have been low enough for the development of microbial activity.

6. Conclusions

A sulfur isotope study of sedimentary pyrite from Mesoarchean rocks of the Karelian Craton revealed signatures of mass-independent isotope fractionation and wide δ34S variations in pyrite grains. Variations in the magnitude of ∆33S values, in the range of −0.3‰ to +2.7‰, were close to those recorded in the rocks of similar ages in Southern Africa and Western Australia [22,23,24]. The negative and positive ∆33S values observed in the present study were interpreted as a result of transferring photolytic sulfate and elemental sulfur to the sulfide sulfur in these samples; preservation of mass-independent signals in sedimentary rocks from the Karelian Craton supports the influence of atmospheric photochemistry on the sulfur cycling in the Mesoarchean [22].
The most striking features of the isotope data reported here are the remarkably large mass-dependent sulfur isotope fractionations where the δ34S values extend from −10‰ to +32‰ and exceed previously reported δ34S variations in Mesoarchean samples. In situ measurements of sulfide grains also reveal strong heterogeneity for both δ34S and Δ33S values. We argue that the source of these isotope effects was microbial processes, namely bacterial sulfur reduction in combination with microbial sulfur disproportionation. The present results can enhance understanding of sulfur cycling and microbial life in the Mesoarchean.

Author Contributions

Conceptualization, T.A.V. and S.V.V.; methodology, T.A.V. and A.V.I.; investigation, A.V.I. and A.V.A.; writing—original draft preparation, T.A.V., S.V.V. and A.V.A.; writing—review and editing, S.V.V. and A.I.S.; supervision, S.V.V.; project administration, S.V.V.; funding acquisition, S.V.V. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation [grant number 21-17-00076, https://rscf.ru/project/21-17-00076/] (accessed on 12 April 2021).

Data Availability Statement

All reported results have been obtained during our research.

Acknowledgments

We are grateful to Lyudmila Vladimirovna Kuleshevich who is gone by now. We also appreciate reviewers for the valuable comments that helped to improve the manuscript.

Conflicts of Interest

We declare that this manuscript is original, has not been published before and is not currently being considered for publication elsewhere. The authors declare no conflict of interest.

Appendix A. Sulfide Petrography

Several sulfide occurrence styles can be distinguished at the Leksa deposit based on mineral association, texture, and lithology of the host-rocks. We use the classification emphasizing the most important characteristics.
Pyrite dominates in massive and semi-massive ores of the Leksa deposit. Initially the ore consisted of large (up to 1 mm) and more isometric crystals of massive pyrite Figure A1B), sometimes with small chalcopyrite, sphalerite, and galena inclusions. This pyrite resulted from iron sulfide rapid massive precipitation from a high-temperature hydrothermal solution. The following metamorphism led to cleavage and partial recrystallization of primary pyrite. This resulted in the formation of a fine-grained texture composed of many different-sized, mainly xenomorphic, crystals (Figure A1C,D), sometimes containing a large amount of silicate inclusions (porous pyrite). This texture is well observed after sample etching using 70% nitric acid (HNO3).
Pyrite precipitation textures and forms in poor disseminated ores and host-rocks are different. These ores consist of dispersed pyrite in the form of single crystals, microcrystalline aggregates of irregular shape, microspheres and subspherical formations, framboids (Figure A2). Pyrite is frequently confined to high-C rock segments. Pyrite crystallized from near the bottom or interstitial waters of sedimentary host-rocks.
Collomorphous forms of pyrite are found in carbonaceous sediments. They form irregular lenses, oval and ellipsoid globular aggregates with a clearly visible lamination that is sometimes concentric (Figure A3). Their matrix is represented by dendrite-like pyrite forming fine-grained structures of intimate intergrowths of iron sulfide thin acicular and bladed crystals with silicate minerals (quartz, chlorite, muscovite, graphite etc.). These intergrowths often have a radial fibrous or interwoven-fibrous texture, which possibly indicates that crystals grow fast in viscous, low-mobility glazer masses due to high crystallization rate of iron sulfide. According to XRD data collomorphous aggregates contain pyrite and marcasite. The further coalescence of dendrite-like crystals led to occurrence of reticulate (E,F) massive porous (C,D) textures.
The figure shows three different microstructures indicating the consecutive order of crystallization stages. Early pyrite (Figure A3E–G) consists of submillimeter acicular or box-like crystals and agglomerates of pyrite 2–20 μm in size with an abundant silicate inclusion (dendrite-like pyrite). Early pyrite is replaced with massive pyrite (Figure A3B–D) containing only small silicate inclusions. Massive pyrite is composed of microcrystals of early pyrite cemented by a later pyrite filling the gaps and replacing silicate minerals. Late-stage pyrite is found as idiomorphic grains without inclusions filling the cracks; it overgrows in the form of a crown collomorphous aggregate associated with quartz.
In situ analyses revealed strongly heterogenous δ34S values, ranging from −6.7‰ to +27.5‰ in concentrically laminated pyrite grains. The Δ33S values mainly vary between −0.3‰ and +0.5‰. The exception is relatively large positive Δ33S = +1.4‰ measured in the core of the concentrically laminated pyrite.
Euhedral pyrite crystals around the concentrically laminated pyrite are relatively homogeneous in δ34S and Δ33S values, ranging from +4.6‰ to +7.0‰ and from +1.55‰ to +1.93‰, respectively.
Carbonaceous shales contain spheroids (vesicles) of pyrite (Figure A5). Morphological analysis of their walls shows that vesicles are fractal. Large spheres 10–20 mm in diameter consist of multiple smaller vesicles (d = 1–2 mm) composed of even smaller spheres. Moreover, certain cases allow the formation of a tubular structure of large sphere walls (Figure A5E,F). In some cases, it is possible to identify a thin carbon film both on the internal and external walls of a sphere or its fragment (Figure A5F). Findings of spheroid microstructures in the Archean deposits are subject to a fierce debate.
Figure A1. Massive pyrite ore, polished section, reflected light microscopy, sample H1-38,9: (A) Massive sulfide ore image; (BD) relicts of primary pyrite crystals and their relationship with silicate minerals and fine-grained pyrite. Please note that fine-grained mass bears traces of metamorphic cleavage.
Figure A1. Massive pyrite ore, polished section, reflected light microscopy, sample H1-38,9: (A) Massive sulfide ore image; (BD) relicts of primary pyrite crystals and their relationship with silicate minerals and fine-grained pyrite. Please note that fine-grained mass bears traces of metamorphic cleavage.
Minerals 12 01143 g0a1
Figure A2. BSE image of finely dispersed precipitations of globular and microcrystalline pyrite in quartz–carbonate matrix of the host-rock (polished section, sample H3-56): (A) general view of finely dispersed pyrite accumulation; (BD) typical shapes of microcrystals. They are represented by both separate euhedral crystals (B,C) and their isometric (C) and subspherical (BD) intergrowths consisting of small (0.5–1.0 μm) vesicles.
Figure A2. BSE image of finely dispersed precipitations of globular and microcrystalline pyrite in quartz–carbonate matrix of the host-rock (polished section, sample H3-56): (A) general view of finely dispersed pyrite accumulation; (BD) typical shapes of microcrystals. They are represented by both separate euhedral crystals (B,C) and their isometric (C) and subspherical (BD) intergrowths consisting of small (0.5–1.0 μm) vesicles.
Minerals 12 01143 g0a2
Figure A3. Collomorphous pyrite, BSE image (AE). Laminated aggregate bordered with carbonaceous shale. It should be noted that parts (F,G) are outside of part (A).
Figure A3. Collomorphous pyrite, BSE image (AE). Laminated aggregate bordered with carbonaceous shale. It should be noted that parts (F,G) are outside of part (A).
Minerals 12 01143 g0a3
Figure A4. Reflected light microscopy image of the concentrically laminated pyrite grains surrounded by euhedral pyrite crystals. Note that the sample is the same as in Figure A3A. Also shown are ablation pits for in situ sulfur isotope analysis of δ34S and Δ33S.
Figure A4. Reflected light microscopy image of the concentrically laminated pyrite grains surrounded by euhedral pyrite crystals. Note that the sample is the same as in Figure A3A. Also shown are ablation pits for in situ sulfur isotope analysis of δ34S and Δ33S.
Minerals 12 01143 g0a4
Figure A5. BSE image of pyrite microspheres (vesicular microstructures) and their debris in a carbonaceous shale, polished section, sample H1-75, the Leksa deposit: (A,B) general view of a shale composed of thin layers of quartz, chlorite (chamosite), muscovite, carbonaceous matter and pyrite microsphere accumulations in it. Microspheres and their debris have a complex structure, consisting of individual equant microcrystals (C,D), and in some cases it can be assumed that their walls are tubular (E,F).
Figure A5. BSE image of pyrite microspheres (vesicular microstructures) and their debris in a carbonaceous shale, polished section, sample H1-75, the Leksa deposit: (A,B) general view of a shale composed of thin layers of quartz, chlorite (chamosite), muscovite, carbonaceous matter and pyrite microsphere accumulations in it. Microspheres and their debris have a complex structure, consisting of individual equant microcrystals (C,D), and in some cases it can be assumed that their walls are tubular (E,F).
Minerals 12 01143 g0a5
Some scientists consider them to be traces of microorganism activity (see, for example, ref. [110] and references therein); others interpret them as pyritized and silicified fragments of vesicular volcanic glass ([111]). In the present case it is difficult to suggest that these characteristics are typical of abiotic.
Table A1. Isotopic data for pyrite from sedimentary rocks of the Leksa deposit.
Table A1. Isotopic data for pyrite from sedimentary rocks of the Leksa deposit.
Sample
(Drill Core/Depth)
δ34S (‰)Δ33S (‰)Δ36S (‰)Pyrite Types
H1/50.7−1.30.03 Euhedral crystal disseminated in rocks
H1/50.7−0.1−0.04 Euhedral crystal disseminated in rocks
H1/50.7−0.9−0.02 Euhedral crystal disseminated in rocks
H1/50.7−1.20.03 Euhedral crystal disseminated in rocks
H1/50.75.40.04 Euhedral crystal disseminated in rocks
H1/54.613.40.02 Euhedral crystal disseminated in rocks
H1/54.613.90.34 Euhedral crystal disseminated in rocks
H1/54.613.5−0.02 Euhedral crystal disseminated in rocks
H1/54.6−3.20.00 Euhedral crystal disseminated in rocks
H1/54.6−2.9−0.05 Euhedral crystal disseminated in rocks
H1/63−1.00.00 Euhedral crystal disseminated in rocks
H1/75−4.80.84 Euhedral crystal disseminated in rocks
H1/75−4.2−0.38 Euhedral crystal disseminated in rocks
H1/75−7.00.23 Euhedral crystal disseminated in rocks
H1/75−1.0−0.51 Euhedral crystal disseminated in rocks
H2/24.031.9−0.18 Euhedral crystal disseminated in rocks
H2/24.032.7−0.16 Euhedral crystal disseminated in rocks
H2/61.52.6−0.01 Euhedral crystal disseminated in rocks
H2/61.52.30.04 Euhedral crystal disseminated in rocks
H2/61.5−5.10.05 Euhedral crystal disseminated in rocks
H2/61.52.4−0.02 Euhedral crystal disseminated in rocks
H2/61.52.20.02 Euhedral crystal disseminated in rocks
H3/56.013.81.48 Euhedral crystal disseminated in rocks
H3/56.012.61.28 Euhedral crystal disseminated in rocks
H3/56.07.60.25 Euhedral crystal disseminated in rocks
H3/56.011.50.40 Euhedral crystal disseminated in rocks
H3/56.013.80.12 Euhedral crystal disseminated in rocks
H3/56.011.60.28 Euhedral crystal disseminated in rocks
H3/56.012.10.32 Euhedral crystal disseminated in rocks
H3/56.010.20.28 Euhedral crystal disseminated in rocks
H3/56.012.00.35 Euhedral crystal disseminated in rocks
H3/56.011.90.28 Euhedral crystal disseminated in rocks
H3/56.010.30.41 Euhedral crystal disseminated in rocks
H3/56.05.40.10 Euhedral crystal disseminated in rocks
H3/56.016.90.67 Euhedral crystal disseminated in rocks
H3/56.012.20.42 Euhedral crystal disseminated in rocks
H1/759.91.65 Euhedral, observed as ring-shaped aggregates
H1/7511.02.65 Euhedral, observed as ring-shaped aggregates
H1/753.71.76−1.3Euhedral, observed as ring-shaped aggregates
H1/752.81.57−2.0Euhedral, observed as ring-shaped aggregates
H1/753.81.93−2.4Euhedral, observed as ring-shaped aggregates
H1/753.81.85−2.5Euhedral, observed as ring-shaped aggregates
H1/756.11.95 Euhedral crystal, overgrowth around colloform grains
H1/756.32.20 Euhedral crystal, overgrowth around colloform grains
H1/755.21.55 Euhedral crystal, overgrowth around colloform grains
H1/754.61.55 Euhedral crystal, overgrowth around colloform grains
H1/755.31.93 Euhedral crystal, overgrowth around colloform grains
H1/755.21.96 Euhedral crystal, overgrowth around colloform grains
H1/755.11.85 Euhedral crystal, overgrowth around colloform grains
H1/757.02.64 Euhedral crystal, overgrowth around colloform grains
H1/755.72.06 Euhedral crystal, overgrowth around colloform grains
H1/755.92.02 Euhedral crystal, overgrowth around colloform grains
H1/757.81.40 Colloform grains_1, core
H1/753.70.31 Colloform grains_1, concentric layer
H1/750.50.00 Colloform grains_1, concentric layer
H1/75−1.80.00 Colloform grains_1, concentric layer
H1/753.0−0.01 Colloform grains_1, concentric layer
H1/7516.2−0.01 Colloform grains_1, concentric layer
H1/7527.50.44 Colloform grains_1, concentric layer
H1/7524.50.33 Colloform grains_1, concentric layer
H1/7515.4−0.01 Colloform grains_1, concentric layer
H1/75−6.70.05 Colloform grains_1, concentric layer
H1/75−9.50.13 Colloform grains_2, non-laminated
H1/75−9.80.22 Colloform grains_2, non-laminated
H1/75−10.20.11 Colloform grains_2, non-laminated
H1/757.00.08 Colloform grains_3, non-laminated
H1/751.6−0.30 Colloform grains_3, non-laminated
H1/75−6.6−0.27 Colloform grains_3, non-laminated
H3/56.05.70.17 Colloform grains_4, core
H3/56.08.50.23 Colloform grains_4, concentric layer
H3/56.010.60.48 Colloform grains_4, concentric layer
H3/56.010.10.49 Colloform grains_4, concentric layer
H3/56.011.90.45 Colloform grains_4, concentric layer
Figure A6. Plot of Δ33S versus Δ36S for the euhedral pyrite crystals from the sample, showing in Figure A4. The dashed line on the plot is a linear fit to data with a slope of −1.15. Also shown is the solid line with a slope of −6.85 for mass-dependent sulfur isotope fractionation in Phanerozoic samples, reported in [31]. Note that our data cannot be associated with microbial mass-dependent sulfur isotope fractionation.
Figure A6. Plot of Δ33S versus Δ36S for the euhedral pyrite crystals from the sample, showing in Figure A4. The dashed line on the plot is a linear fit to data with a slope of −1.15. Also shown is the solid line with a slope of −6.85 for mass-dependent sulfur isotope fractionation in Phanerozoic samples, reported in [31]. Note that our data cannot be associated with microbial mass-dependent sulfur isotope fractionation.
Minerals 12 01143 g0a6

References

  1. Jørgensen, B.B.; Isaksen, M.F.; Jannasch, H.W. Bacterial sulfate reduction above 100° C in deep-sea hydrothermal vent sediments. Science 1992, 258, 1756–1757. [Google Scholar] [CrossRef]
  2. Jørgensen, B.B.; Findlay, A.J.; Pellerin, A. The Biogeochemical sulfur cycle of marine sediments. Front. Microbiol. 2019, 10, 849. [Google Scholar] [CrossRef]
  3. Hofmann, H.J.; Grey, K.; Hickman, A.H.; Thorpe, R.I. Origin of 3.45 Ga coniform stromatolites in Warrawoona Group, Western Australia. Geol. Soc. Am. Bull. 1999, 111, 1256–1262. [Google Scholar] [CrossRef]
  4. Philippot, P.; Van Zuilen, M.; Lepot, K.; Thomazo, C.; Farquhar, J.; Van Kranendonk, M.J. Early Archaean microorganisms preferred elemental sulfur, not sulfate. Science 2007, 317, 1534–1537. [Google Scholar] [CrossRef] [PubMed]
  5. Philippot, P.; Van Zuilen, M.; Rollion-Bard, C. Variations in atmospheric sulphur chemistry on early Earth linked to volcanic activity. Nat. Geosci. 2012, 5, 668–674. [Google Scholar] [CrossRef]
  6. Ueno, Y.; Ono, S.; Rumble, D.; Maruyama, S. Quadruple sulfur isotope analysis of ca. 3.5 Ga Dresser Formation: New evidence for microbial sulfate reduction in the early Archean. Geochim. Cosmochim. Acta 2008, 72, 5675–5691. [Google Scholar] [CrossRef]
  7. Shen, Y.; Farquhar, J.; Masterson, A.; Kaufman, A.J.; Buick, R. Evaluating the role of microbial sulfate reduction in the early Archean using quadruple isotope systematics. Earth Planet. Sci. Lett. 2009, 279, 383–391. [Google Scholar] [CrossRef]
  8. Wacey, D.; Noffke, N.; Cliff, J.; Barley, M.E.; Farquhar, J. Micro-scale quadruple sulfur isotope analysis of pyrite from the ∼3480 Ma Dresser Formation: New insights into sulfur cycling on the early Earth. Precambrian Res. 2015, 258, 24–35. [Google Scholar] [CrossRef]
  9. Baumgartner, R.J.; Caruso, S.; Fiorentini, M.L.; van Kranendonk, M.J.; Martin, L.; Jeon, H.; Pagès, A.; Wacey, D. Sulfidization of 3.48 billion-year-old stromatolites of the Dresser Formation, Pilbara Craton: Constraints from in-situ sulfur isotope analysis of pyrite. Chem. Geol. 2020, 538, 119488. [Google Scholar] [CrossRef]
  10. Roerdink, D.L.; Mason, P.R.D.; Whitehouse, M.J.; Reimer, T. High-resolution quadruple sulfur isotope analyses of 3.2 Ga pyrite from the Barberton Greenstone Belt in South Africa reveal distinct environmental controls on sulfide isotopic arrays. Geochim. Cosmochim. Acta 2013, 117, 203–215. [Google Scholar] [CrossRef]
  11. Montinaro, A.; Strauss, H.; Mason, P.R.D.; Roerdink, D.; Münker, C.; Schwarz-Schampera, U.; Arndt, N.T.; Farquhar, J.; Beukes, N.J.; Gutzmer, J.; et al. Paleoarchean sulfur cycling: Multiple sulfur isotope constraints from the Barberton Greenstone Belt, South Africa. Precambrian Res. 2015, 267, 311–322. [Google Scholar] [CrossRef]
  12. Nabhan, S.; Marin-Carbonne, J.; Mason, P.R.D.; Heubeck, C. In situ S-isotope compositions of sulfate and sulfide from the 3.2 Ga Moodies Group, South Africa: A record of oxidative sulfur cycling. Geobiology 2020, 18, 426–444. [Google Scholar] [CrossRef]
  13. Canfield, D.E. Biogeochemistry of sulfur isotopes. Rev. Miner. Geochem. 2001, 43, 607–636. [Google Scholar] [CrossRef]
  14. Shen, Y.; Buick, R.; Canfield, D.E. Isotopic evidence for microbial sulphate reduction in the early Archaean era. Nature 2001, 410, 77–81. [Google Scholar] [CrossRef]
  15. Liu, L.; Ireland, T.R.; Holden, P. SHRIMP 4-S isotope systematics of two pyrite generations in the 3.49 Ga Dresser Formation. Geochem. Perspect. Lett. 2021, 17, 45–49. [Google Scholar] [CrossRef]
  16. Kaufman, A.J.; Johnston, D.T.; Farquhar, J.; Masterson, A.L.; Lyons, T.W.; Bates, S.; Anbar, A.D.; Arnold, G.L.; Garvin, J.; Buick, R. Late Archean biospheric oxygenation and atmospheric evolution. Science 2007, 317, 1900–1903. Available online: https://0-www-jstor-org.brum.beds.ac.uk/stable/20048474 (accessed on 28 September 2007). [CrossRef] [PubMed]
  17. Ono, S.; Beukes, N.; Rumble, D. Origin of two distinct multiple-sulfur isotope compositions of pyrite in the 2.5 Ga Klein Naute Formation, Griqualand West Basin, South Africa. Precambrian Res. 2009, 169, 48–57. [Google Scholar] [CrossRef]
  18. Ono, S.; Kaufman, A.J.; Farquhar, J. Lithofacies control on multiple-sulfur isotope records and Neoarchean sulfur cycles. Precambrian Res. 2009, 169, 58–67. [Google Scholar] [CrossRef]
  19. Zerkle, A.L.; Claire, M.W.; Domagal-Goldman, S.D.; Farquhar, J.; Poulton, S.W. A bistable organic-rich atmosphere on the Neoarchaean Earth. Nat. Geosci. 2012, 5, 359–363. [Google Scholar] [CrossRef]
  20. Farquhar, J.; Cliff, J.; Zerkle, A.L.; Kamyshny, A.; Poulton, S.W.; Claire, M.; Adams, D.; Harms, B. Pathways for Neoarchean pyrite formation constrained by mass-independent sulfur isotopes. Proc. Natl. Acad. Sci. USA 2013, 110, 17638–17643. [Google Scholar] [CrossRef] [Green Version]
  21. Zhelezinskaia, I.; Kaufman, A.J.; Farquhar, J.; Cliff, J. Large sulfur isotope fractionations associated with Neoarchean microbial sulfate reduction. Science 2014, 346, 742–744. [Google Scholar] [CrossRef] [PubMed]
  22. Farquhar, J.; Peters, M.; Johnston, D.T.; Strauss, H.; Masterson, A.; Wiechert, U.; Kaufman, A.J. Isotopic evidence for Mesoarchaean anoxia and changing atmospheric sulphur chemistry. Nature 2007, 449, 706–709. [Google Scholar] [CrossRef] [PubMed]
  23. Guy, B.M.; Ono, S.; Gutzmer, J.; Kaufman, A.J.; Lin, Y.; Fogel, M.L.; Beukes, N.J. A multiple sulfur and organic carbon isotope record from non-conglomeratic sedimentary rocks of the Mesoarchean Witwatersrand Supergroup, South Africa. Precambrian Res. 2012, 216–219, 208–231. [Google Scholar] [CrossRef]
  24. Ono, S.; Beukes, N.J.; Rumble, D.; Fogel, M.L. Early evolution of atmospheric oxygen from multiple-sulfur and carbon isotope records of the 2.9 Ga Mozaan Group of the Pongola Supergroup, Southern Africa. S. Afr. J. Geol. 2006, 109, 97–108. [Google Scholar] [CrossRef]
  25. Strauss, H. Sulphur isotopes and the early Archaean sulphur cycle. Precambrian Res. 2003, 126, 349–361. [Google Scholar] [CrossRef]
  26. Farquhar, J.; Bao, H.; Thiemens, M. Atmospheric influence of Earth’s earliest sulphur cycle. Science 2000, 289, 756–758. [Google Scholar] [CrossRef]
  27. Johnston, D. Multiple sulfur isotopes and the evolution of Earth’s surface sulfurcycle. Earth-Sci. Rev. 2011, 106, 161–183. [Google Scholar] [CrossRef]
  28. Farquhar, J.; Savarino, J.; Airieau, S.; Thiemens, M.H. Observation of wavelength-sensitive mass-independent sulfur isotope effects during SO2 photolysis: Implications for the early atmosphere. J. Geophys. Res. 2001, 106, 32829–32839. [Google Scholar] [CrossRef]
  29. Pavlov, A.A.; Kasting, J.F. Mass-independent fractionation of sulphur isotopesin Archaean sediments: Strong evidence for an anoxic Archaean atmosphere. Astrobiology 2002, 2, 27–41. [Google Scholar] [CrossRef]
  30. Ono, S.; Eigenbrode, J.L.; Pavlov, A.A.; Kharecha, P.; Ruble, D.; Kasting, J.F.; Freeman, K.H. New insights into Archean sulfur cycle from mass-independent sulfurisotope records from the Hamersley Basin, Australia. Earth Planet. Sci. Lett. 2003, 213, 15–30. [Google Scholar] [CrossRef]
  31. Ono, S.; Wing, B.; Johnston, D.; Farquhar, J.; Rumble, D. Mass-dependent fractionation of quadruple stable sulfur isotope system as a new tracer of sulfur biogeochemical cycles. Geochim. Cosmochim. Acta 2006, 70, 2238–2252. [Google Scholar] [CrossRef]
  32. Glebovitskii, V.A. The Early Precambrian of the Baltic Shield; Nauka: St. Petersburg, Russia, 2005. (In Russain) [Google Scholar]
  33. Hölttä, P.; Heilimo, E.; Huhma, H.; Kontinen, A.; Mertanen, S.; Mikkola, P.; Paavola, J.; Peltonen, P.; Semprich, J.; Slabunov, A.; et al. The Archaean of the Karelia Province in Finland. Geol. Surv. Finl. Spec. Pap. 2012, 54, 21–73. [Google Scholar]
  34. Kulikov, V.S.; Svetov, S.A.; Slabunov, A.I.; Kulikova, V.V.; Polin, A.K.; Golubev, A.I.; Gorkovets, V.Y.; Ivashchenko, V.I.; Gogolev, M.A. Geological map of Southeastern Fennoscandia (scale 1:750 000): A new approach to map compilation. Trans. KarRC RAS. 2017, 2, 3–41. [Google Scholar] [CrossRef]
  35. Slabunov, A.I.; Lobach-Zhuchenko, S.B.; Bibikova, E.V.; Sorjonen-Ward, P.; Balagansky, V.V.; Volodichev, O.I.; Shchipansky, A.A.; Svetov, S.A.; Chekulaev, V.P.; Arestova, N.A.; et al. The Archaean Nucleus of the Baltic/Fennoscandian Shield. In European Lithosphere Dynamics: Geological Society of London, Memoir; Gee, D.G., Stephenson, R.A., Eds.; European Lithosphere Dynamics Geological Society: London, UK, 2006; Volume 32, pp. 627–644. [Google Scholar]
  36. Slabunov, A.I.; Nesterova, N.S.; Egorov, A.V.; Kuleshevich, L.V.; Kevlich, V.I. Age of the Archean strata with banded iron formation in the Kostomuksha Greenstone Belt, Karelian Craton, Fennoscandian Shield: Constraints on the geochemistry and geochronology of zircons. Geochem. Int. 2021, 59, 341–356. [Google Scholar] [CrossRef]
  37. Hölttä, P.; Heilimo, E.; Huhma, H.; Kontinen, A.; Mertanen, S.; Mikkola, P.; Paavola, J.; Peltonen, P.; Semprich, J.; Slabunov, A.; et al. The Archaean Karelia and Belomorian Provinces, Fennoscandian Shield. In Evolution of Archean Crust and Early Life. Modern Approaches in Solid Earth Sciences; Dilek, Y., Furnes, H., Eds.; Springer: Berlin/Heidelberg, Germany, 2014; pp. 55–102. [Google Scholar] [CrossRef]
  38. Lobach-Zhuchenko, S.B.; Chekulaev, V.P.; Arestova, N.A.; Levskii, L.K.; Kovalenko, A.V. Archaean terranes in Karelia: Geological and isotopic–geochemical evidence. Geotectonics 2000, 34, 452–466. [Google Scholar]
  39. Puchtel, I.S.; Hofmann, A.W.; Amelin, Y.u.V.; Garbe-Schönberg, C.-D.; Samsonov, A.V.; Shchipansky, A.A. Combined mantle plume-island arc model for the formation of the 2.9 ga Sumozero-Kenozero greenstone belt, se baltic shield: Isotope and trace element constraints. Geochim. Cosmochim. Acta 1999, 63, 3579–3595. [Google Scholar] [CrossRef]
  40. Kuleshevich, L.V. Metamorphism and Ore Potential of Archean Greenstone Belts from the Southeastern Margin of the Baltic Shield; Karelian Scientific Center: Petrozavodsk, Russia, 1992. (In Russian) [Google Scholar]
  41. Kuleshevich, L.V.; Furman, V.N.; Fedyuk, Z.N. Promises for gold mineralization in the Kamenoozero structure of the Sumozero-Kenozero greenstone belt. Geol. Useful Miner. Karelia. 2005, 8, 50–67. (In Russian) [Google Scholar]
  42. Hölttä, P.; Heilimo, E.; Huhma, H.; Kontinen, A.; Lauri, L.; Slabunov, A. Paleoarchean rocks in the Fennoscandian Shield (Chapter 32). In Earth’s Oldest Rocks, 2nd ed.; Van Kranendonk, M.J., Bennett, V.C., Hoffmann, J.E., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 819–836. [Google Scholar]
  43. Vysotskii, S.V.; Khanchuk, A.I.; Kuleshevich, L.V.; Ignatiev, A.V.; Slabunov, A.I.; Velivetskaya, T.A. The Multi-isotope composition of sulfur in sulfides and microfossils of the Mesoarchean Leksa pyrite ore occurrence of the Karelian Craton: New data on abiogenic and biogenic effects on the formation of ancient ores. Dokl. Earth Sci. 2019, 485, 409–412. [Google Scholar] [CrossRef]
  44. Vysotskiy, S.V.; Slabunov, A.I.; Ignatiev, A.V.; Kuleshevich, L.V.; Velivetskaya, T.A. Multiple sulfur isotopes from Mesoarchean volcano-sedimentary massive sulfide deposits of the Karelian Craton: Significance for determining sulfur sources, biogeochemical processes and the genesis of deposits. Russ. Geol. Geophys. 2022. [Google Scholar] [CrossRef]
  45. Kuleshevich, L.V.; Belashev, B.Z. Pyrite mineralization in East Karelia (experience in studying the composition and electrophysical properties of pyrites). Geol. Miner. Karelia 1998, 1, 57–72. (In Russian) [Google Scholar]
  46. Galley, A.G.; Hannington, M.; Jonasson, I. Volcanogenic massive sulphide deposits. In Mineral Deposits of Canada—A Synthesis of Major Deposit-Types, District Metallogeny, the Evolution of Geological Provinces, and Exploration Methods; Goodfellow, W.D., Ed.; Special Publication 5; Geological Association of Canada, Mineral Deposits Division: Saint John, NB, Canada, 2007; pp. 141–161. [Google Scholar]
  47. Hannington, M.D.; Bleeker, W.; Kjarsgaard, I. Sulfide mineralogy, geochemistry, and ore genesis of the Kidd Creek deposit—Part I. North, central, and south orebodies. In The Giant Kidd Creek Volcanogenic Massive Sulfide Deposit, Western Abitibi Subprovince, Canada; Hannington, M.D., Barrie, C.T., Eds.; Economic Geology Monograph Series; Geo Science World: McLean, VA, USA, 1999; Volume 10, pp. 163–224. [Google Scholar] [CrossRef]
  48. Goodfellow, W.D.; McCutcheon, S.R. Geologic and genetic attributes of volcanic sediment-hosted massive sulfide deposits of the Bathurst mining camp, New Brunswick—A synthesis. In Massive Sulfide Deposits of the Bathurst Mining Camp, New Brunswick, and Northern Maine; Goodfellow, W.D., McCutcheon, S.R., Peter, J.M., Eds.; Economic Geology Monograph Series; Geo Science World: McLean, VA, USA, 2003; Volume 11, pp. 245–301. [Google Scholar]
  49. Barrie, C.T.; Hannington, M.D. Classification of volcanic-associated massive sulfide deposits based on host-rock composition. In Volcanic-Associated Massive Sulfide Deposits: Progresses and Examples in Modern and Ancient Settings; Barrie, C.T., Hannington, M.D., Eds.; Reviews in Economic Geology; Geo Science World: McLean, VA, USA, 1999; Volume 8, pp. 1–11. [Google Scholar] [CrossRef]
  50. Franklin, J.M.; Gibson, H.L.; Jonasson, I.R.; Galley, A.G. Volcanogenic massive sulfide deposits. In One Hundredth Anniversary Volume; Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P., Eds.; Geo Science World: McLean, VA, USA, 2005; pp. 523–560. [Google Scholar] [CrossRef]
  51. Hannington, M.D.; de Ronde, C.E.; Petersen, S. Sea-Floor Tectonics and Submarine Hydrothermal Systems. Economic Geology 100th Anniversary Volume, 1905–2005: Littleton, Colo; Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P., Eds.; Geo Science World: McLean, VA, USA, 2005; pp. 111–141. [Google Scholar] [CrossRef]
  52. Ignatiev, A.V.; Velivetskaya, T.A.; Budnitskiy, S.Y.; Yakovenko, V.V.; Vysotskiy, S.V.; Levitskii, V.I. Precision analysis of multisulfur isotopes in sulfides by femtosecond laser ablation GC-IRMS at high spatial resolution. Chem. Geol. 2018, 493, 316–326. [Google Scholar] [CrossRef]
  53. Velivetskaya, T.A.; Ignatiev, A.V.; Yakovenko, V.V.; Vysotskiy, S.V. An improved femtosecond laser-ablation fluorination method for measurements of sulfur isotopic anomalies (Δ33S and Δ36S) in sulfides with high precision. Rapid Commun. Mass Spectrom. 2019, 33, 1722–1729. [Google Scholar] [CrossRef] [PubMed]
  54. Farquhar, J.; Savarino, J.; Jackson, T.L.; Thiemens, M.H. Evidence of atmospheric sulphur in the martian regolith from sulphur isotopes in meteorites. Nature 2000, 404, 50–52. [Google Scholar] [CrossRef] [PubMed]
  55. Berner, R.A. Sedimentary pyrite formation. Am. J. Sci. 1970, 268, 1–23. [Google Scholar] [CrossRef]
  56. Rickard, D.T. Kinetics of pyrite formation by the H2S oxidation of iron (II) monosulfide in aqueous solutions between 25 and 125°C: The rate equation. Geochim. Cosmochim. Acta 1997, 61, 115–134. [Google Scholar] [CrossRef]
  57. Rickard, D.; Luther, G.W., III. Kinetics of pyrite formation by the H2S oxidation of iron (II) monosulfide in aqueous solutions between 25 and 125°C: The mechanism. Geochim. Cosmochim. Acta 1997, 61, 135–147. [Google Scholar] [CrossRef]
  58. Masterson, A.L.; Farquhar, J.; Wing, B.A. Sulfur mass-independent fractionation patterns in the broadband UV photolysis of sulfur dioxide: Pressure and third body effects. Earth Planet Sci. Lett. 2011, 306, 253–260. [Google Scholar] [CrossRef]
  59. Whitehill, A.R.; Ono, S. Excitation band dependence of sulfur isotope mass-independent fractionation during photochemistry of sulfur dioxide using broadband light sources. Geochim. Cosmochim. Acta 2012, 94, 238–253. [Google Scholar] [CrossRef]
  60. Ono, S.; Whitehill, A.R.; Lyons, J.R. Contribution of isotopologue self-shielding to sulfur mass-independent fractionation during sulfur dioxide photolysis. J. Geophys. Res. Atmospheres 2013, 118, 2444–2454. [Google Scholar] [CrossRef]
  61. Whitehill, A.R.; Xie, C.; Hu, X.; Xie, D.; Guo, H.; Ono, S. Vibronic origin of sulfur mass-independent isotope effect in photoexcitation of SO2 and the implications to the early earth’s atmosphere. Proc. Natl. Acad. Sci. USA 2013, 110, 17697–17702. [Google Scholar] [CrossRef]
  62. Endo, Y.; Ueno, Y.; Aoyama, S.; Danielache, S.O. Sulfur isotope fractionation by broadband UV radiation to optically thin SO2 under reducing atmosphere. Earth Planet Sci. Lett. 2016, 453, 9–22. [Google Scholar]
  63. Endo, Y.; Danielache, S.O.; Ueno, Y. Total pressure dependence of sulfur mass-independent fractionation by SO2 photolysis. Geophys. Res. Lett. 2019, 46, 483–491. [Google Scholar] [CrossRef]
  64. Ignatiev, A.V.; Velivetskaya, T.A.; Yakovenko, V.V. Effect of mass-independent isotope fractionation of sulfur (Δ33S and Δ36S) during SO2 photolysis in experiments with a broadband light source. Geochem Int. 2019, 57, 751–760. [Google Scholar] [CrossRef]
  65. Velivetskaya, T.A.; Ignatiev, A.V.; Yakovenko, V.V. Mass-independent sulfur isotope fractionation in the photochemical SO2 processes under the UV radiation of different wave length. Geochem Int. 2020, 58, 1228–1238. [Google Scholar] [CrossRef]
  66. Thode, H.; Kleerekoper, H.; McElcheran, D. Isotope fractionation in the bacterial reduction of sulphate. Research 1951, 4, 581–582. [Google Scholar]
  67. Kaplan, I.R.; Rafter, T.A. Fractionation of stable isotopes of sulfur by Thiobacilli. Science 1958, 127, 517–518. [Google Scholar] [CrossRef]
  68. Kaplan, I.R.; Rittenberg, S.C. Microbiological fractionation of sulphur isotopes. Microbiology 1964, 34, 195–212. [Google Scholar] [CrossRef]
  69. Kemp, A.L.W.; Thode, H.G. The mechanism of the bacterial reduction of sulphate and of sulphite from isotope fractionation studies. Geochim. Cosmochim. Acta 1968, 32, 71–91. [Google Scholar] [CrossRef]
  70. Sim, M.S.; Bosak, T.; Ono, S. Large sulfur isotope fractionation does not require disproportionation. Science 2011, 333, 74–77. [Google Scholar] [CrossRef]
  71. Drake, H.; Tullborg, E.-L.; Whitehouse, M.; Sandberg, B.; Blomfeldt, T.; Åström, M.E. Extreme fractionation and micro-scale variation of sulphur isotopes during bacterial sulphate reduction in deep groundwater systems. Geochim. Cosmochim. Acta 2015, 161, 1–18. [Google Scholar] [CrossRef]
  72. Drake, H.; Whitehouse, M.J.; Heim, C.; Reiners, P.W.; Tillberg, M.; Hogmalm, K.J.; Dopson, M.; Broman, C.; Åström, M.E. Unprecedented 34S-enrichment of pyrite formed following microbial sulfate reduction in fractured crystalline rocks. Geobiology 2018, 16, 556–574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Cavalazzi, B.; Agangi, A.; Barbieri, R.; Franchi, F.; Gasparotto, G. The formation of low-temperature sedimentary pyrite and its relationship with biologically-induced processes. Geol. Ore Depos. 2014, 56, 395–408. [Google Scholar] [CrossRef]
  74. Hallbauer, D.K.; Barton, J.M. The fossil gold placers of the Witwatersrand. Gold Bull. 1987, 20, 68–79. [Google Scholar] [CrossRef]
  75. Frizzo, P.; Rampazzo, G.; Molinaroli, E. Authigenic iron sulphides in Recent sediments of the Venice Lagoon (northern Italy). Eur. J. Miner. 1991, 3, 603–612. [Google Scholar] [CrossRef]
  76. Huang, F.; Gao, S.; Chen, L.; Su, L.; Li, Y.; Meng, L.; Liu, K.; Chai, C.; Qi, X. Micro-texture and in situ sulfur isotope of pyrite from the Baiyunpu Pb-Zn deposit in central Hunan, South China: Implications for the growth mechanism of colloform pyrite aggregates. J. Asian Earth Sci. 2020, 193, 104302. [Google Scholar] [CrossRef]
  77. Agangi, A.; Hofmann, A.; Rollion-Bard, C.; Marin-Carbonne, J.; Cavalazzi, B.; Large, R.; Meffre, S. Gold accumulation in the Archaean Witwatersrand Basin, South Africa—evidence from concentrically laminated pyrite. Earth Sci. Rev. 2015, 140, 27–53. [Google Scholar] [CrossRef]
  78. Canfield, D.E. Isotope fractionation by natural populations of sulfate-reducing bacteria. Geochim. Cosmochim. Acta 2001, 65, 1117–1124. [Google Scholar] [CrossRef]
  79. Habicht, K.S.; Gade, M.; Thamdrup, B.; Berg, P.; Canfield, D.E. Calibration of sulfate levels in the Archean Ocean. Science 2002, 298, 2372–2374. [Google Scholar] [CrossRef] [PubMed]
  80. Jamieson, J.W.; Wing, B.A.; Farquhar, J.; Hannington, M.D. Neoarchaean seawater sulphate concentrations from sulphur isotopes in massive sulphide ore. Nat. Geosci. 2013, 6, 61–64. [Google Scholar] [CrossRef]
  81. Crowe, S.A.; Paris, G.; Katsev, S.; Jones, C.; Kim, S.-T.; Zerkle, A.L.; Nomosatryo, S.; Fowle, D.A.; Adkins, J.F.; Sessions, A.L.; et al. Sulfate was a trace constituent of Archean seawater. Science 2014, 346, 735–739. Available online: https://0-www-jstor-org.brum.beds.ac.uk/stable/24917791 (accessed on 7 November 2014). [CrossRef]
  82. Canfield, D.E. The evolution of the Earth surface sulfur reservoir. Am. J. Sci. 2004, 304, 839–861. [Google Scholar] [CrossRef]
  83. Ingvorsen, K.; Zehnder, A.J.B.; Jørgensen, B.B. Kinetics of sulfate and acetate uptake by Desulfobacter postgatei. Appl. Environ. Microbiol. 1984, 47, 403–408. [Google Scholar] [CrossRef] [PubMed]
  84. Ono, S.; Eigenbrode, J.L.; Pavlov, A.A.; Kharecha, P.; Rumble, D.; Kasting, J.F.; Freeman, K.H. New insights into Archaean sulfur cycle from mass-independent sulfur isotope records from the Hamersley Basin, Australia. Earth Planet Sci. Lett. 2003, 213, 15–30. [Google Scholar] [CrossRef]
  85. Cypionka, H.; Smock, A.M.; Bottcher, M.E. A combined pathway of sulfur compound disproportionation in Desulfovivrio desulfuricans. FEMS Microbiol. Lett. 1998, 166, 181–186. [Google Scholar] [CrossRef]
  86. Canfield, D.E.; Thamdrup, B. The production of 34S depleted sulfide during bacterial disproportionation of elemental sulfur. Science 1994, 266, 1973–1975. [Google Scholar] [CrossRef] [PubMed]
  87. Thamdrup, B.; Finster, K.; Hansen, J.W.; Bak, F. Bacterial disproportionation of elemental sulfur coupled to chemical reduction of iron or manganese. Appl. Environ. Microbiol. 1993, 59, 101–108. [Google Scholar] [CrossRef]
  88. Canfield, D.E.; Thamdrup, B.; Fleischer, S. Isotope fractionation and metabolism by pure and enrichment cultures of elemental sulfur-disproportionating bacteria. Limnol. Oceanogr. 1998, 43, 253–264. [Google Scholar] [CrossRef]
  89. Böttcher, M.E.; Thamdrup, B.; Vennemann, T.W. Oxygen and sulfur isotope fractionation during anaerobic bacterial disproportionation of elemental sulfur. Geochim. Cosmochim. Acta 2001, 65, 1601–1609. [Google Scholar] [CrossRef]
  90. Böttcher, M.E.; Thamdrup, B. Anaerobic sulfide oxidation and stable isotope fractionation associated with bacterial sulfur disproportionation in the presence of MnO2. Geochim. Cosmochim. Acta 2001, 65, 1573–1581. [Google Scholar] [CrossRef]
  91. Böttcher, M.E.; Thamdrup, B.; Gehre, M.; Theune, A. 34S/32S and 18O/16O fractionation during sulfur disproportionation by Desulfobulbus propionicus. Geomicrobiol. J. 2005, 22, 219–226. [Google Scholar] [CrossRef]
  92. Johnston, D.T.; Farquhar, J.; Wing, B.A.; Kaufman, A.J.; Canfield, D.E.; Habicht, K.S. Multiple sulfur isotope fractionations in biological systems: A case study with sulfate reducers and sulfur disproportionators. Am. J. Sci. 2005, 305, 645–660. [Google Scholar] [CrossRef]
  93. Canfield, D.E.; Teske, A. Late Proterozoic rise in atmospheric oxygen concentration inferred from phylogenetic and sulphur isotope studies. Nature 1996, 382, 127–132. [Google Scholar] [CrossRef] [PubMed]
  94. Johnston, D.T.; Farquhar, J.; Canfield, D.E. Sulfur isotope insights into microbial sulfate reduction: When microbes meet models. Geochim. Cosmochim. Acta 2007, 71, 3929–3947. [Google Scholar] [CrossRef]
  95. Farquhar, J.; Johnston, D.T.; Wing, B.A.; Habicht, K.S.; Canfield, D.E.; Airieau, S.; Thiemens, M.H. Multiple sulphur isotopic interpretations of biosynthetic pathways: Implications for biological signatures in the sulphur isotope record. Geobiology 2003, 1, 27–36. [Google Scholar] [CrossRef]
  96. Farquhar, J.; Johnston, D.T.; Wing, B.A. Implications of conservation of mass effects on mass-dependent isotope fractionations: Influence of network structure on sulfur isotope phase space of dissimilatory sulfate reduction. Geochim. Cosmochim. Acta 2007, 71, 5862–5875. [Google Scholar] [CrossRef]
  97. Ye, Y.T.; Wu, C.D.; Zhai, L.N.; An, Z.Z. Pyrite morphology and episodic euxinia of the Ediacaran Doushantuo Formation in South China. Sci. China Inf. Sci. 2017, 60, 102–113. [Google Scholar] [CrossRef]
  98. Liu, Z.; Chen, D.; Zhang, J.; Lü, X.; Wang, Z.; Liao, W.; Shi, X.; Tang, J.; Xie, G. Pyrite morphology as an indicator of paleoredox conditions and shale gas content of the longmaxi and wufeng shales in the Middle Yangtze Area, South China. Minerals 2019, 7, 428. [Google Scholar] [CrossRef]
  99. Duverger, A.; Berg, J.S.; Busigny, V.; Guyo, F.; Bernard, S.; Miot, J. Mechanisms of pyrite formation promoted by sulfate-reducing bacteria in pure culture. Front. Earth Sci. 2020, 8, 588310. [Google Scholar] [CrossRef]
  100. Wacey, D.; Kilburn, M.R.; Saunders, M.; Cliff, J.; Brasier, M.D. Microfossils of sulphur-metabolizing cells in 3.4-billion-year-old rocks of Western Australia. Nat. Geosci. 2011, 4, 698–702. [Google Scholar] [CrossRef]
  101. Duck, L.J.; Glikson, M.; Golding, S.D.; Webb, R.E. Microbial remains and other carbonaceous forms from the 3.24 Ga Sulphur Springs black smoker deposit, Western Australia. Precambrian Res. 2007, 154, 205–220. [Google Scholar] [CrossRef]
  102. Schidlowski, M. Probable life-forms from the Precambrian of the Witwatersrand system (South Africa). Nature 1965, 205, 895–896. [Google Scholar] [CrossRef]
  103. Guan, C.; Wang, W.; Zhou, C.; Muscente, A.D.; Wan, B.; Chen, X.; Yuan, X.; Chen, Z.; Ouyang, Q. Controls on fossil pyritization: Redox conditions, sedimentary organic matter content, and Chuaria preservation in the Ediacaran Lantian Biota. Palaeo. Palaeo. Palaeo. 2017, 474, 26–35. [Google Scholar] [CrossRef]
  104. Lepot, K. Signatures of early microbial life from the Archean (4 to 2.5 Ga) eon. Earth-Sci. Rev. 2020, 209, 103296. [Google Scholar] [CrossRef]
  105. Donald, R.; Southam, G. Low temperature anaerobic bacterial diagenesis of ferrous monosulfide to pyrite. Geochim. Cosmochim. Acta 1999, 63, 2019–2023. [Google Scholar] [CrossRef]
  106. Bao, H.; Sun, T.; Kohl, I.; Peng, Y. Comment on “Early Archaean microorganisms preferred elemental sulfur, not sulfate”. Science 2008, 319, 1336. [Google Scholar] [CrossRef] [PubMed]
  107. Finster, K. Microbiological disproportionation of inorganic sulfur compounds. J. Sulphur Chem. 2008, 29, 281–292. [Google Scholar] [CrossRef]
  108. Finster, K.; Liesack, W.; Thamdrup, B. Elemental sulfur and thiosulfate disproportionation by Desulfocapsa sulfoexigens sp. nov., a new anaerobic bacterium isolated from marine surface sediment. Appl. Environ. Microbiol. 1998, 64, 119–125. [Google Scholar] [CrossRef]
  109. Holland, H.D. Systematics of isotopic composition of sulfur in oceans during the Phanerozoic and its implications for atmospheric oxygen. Geochim. Cosmochim. Acta 1973, 37, 2605–2616. [Google Scholar] [CrossRef]
  110. Sugitani, K. Early Archean (Pre-3.0 Ga) Cellularly Preserved Microfossils and Microfossil-like Structures from the Pilbara Craton, Western Australia—A Review. In Earth’s Oldest Rocks; van Kranendonk, M.J., Bennett, V.C., Hoffmann, J.E., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 1007–1028. [Google Scholar] [CrossRef]
  111. Wacey, D.; Noffke, N.; Saunders, M.; Guagliardo, P.; Pyle, D.M. Volcanogenic pseudo-fossils from the ~3.48 Ga Dresser Formation, Pilbara, Western Australia. Astrobiology 2018, 18, 539–555. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (A) Map showing the tectonic division of the Fennoscandian shield and the location of the Sumozero-Kenozero greenstone belt [35,36]. 1—Baikalids and Caledonides; 2—Paleoproterozoic juvenile crust; 3—Archean greenstone (A) and paragneiss (B) complexes; 4—Archean crust; 5—The Karelian Craton subprovince boundaries; 6—boundaries of Paleoproterozoic orogens. Letter abbreviation: KP—Kola Province, BP—Belomorian Province; NC—Norrbotten Craton. (B) Geological sketch map of the Kamennoozero greenstone belt and locations of ore deposits indicated by stars. Red star indicates the location of the studied Leksa deposit [34,39,40,41]. 1–3—Paleoproterozoic (Sumian): 1—basaltic andesite; 2—gabbro; 3—ultrabasite; 4–10—Archean: 4—Neoarchean granite; 5—Archean diorite, granodiorite, and tonalite; 6—Mesoarchean BARD- and Adakitic-series sedimentary-volcanic rocks interlaying with interbedded pyrite; 7—pyritic ore occurrences; 8—Mesoarchean (A) ultramafite and (B) gabbro; 9—Mesoarchean tholeiitic basalt; 10—Mesoarchean komatiite; 11—Paleo-Mesoarchean TTG-gneiss; 12—faults.
Figure 1. (A) Map showing the tectonic division of the Fennoscandian shield and the location of the Sumozero-Kenozero greenstone belt [35,36]. 1—Baikalids and Caledonides; 2—Paleoproterozoic juvenile crust; 3—Archean greenstone (A) and paragneiss (B) complexes; 4—Archean crust; 5—The Karelian Craton subprovince boundaries; 6—boundaries of Paleoproterozoic orogens. Letter abbreviation: KP—Kola Province, BP—Belomorian Province; NC—Norrbotten Craton. (B) Geological sketch map of the Kamennoozero greenstone belt and locations of ore deposits indicated by stars. Red star indicates the location of the studied Leksa deposit [34,39,40,41]. 1–3—Paleoproterozoic (Sumian): 1—basaltic andesite; 2—gabbro; 3—ultrabasite; 4–10—Archean: 4—Neoarchean granite; 5—Archean diorite, granodiorite, and tonalite; 6—Mesoarchean BARD- and Adakitic-series sedimentary-volcanic rocks interlaying with interbedded pyrite; 7—pyritic ore occurrences; 8—Mesoarchean (A) ultramafite and (B) gabbro; 9—Mesoarchean tholeiitic basalt; 10—Mesoarchean komatiite; 11—Paleo-Mesoarchean TTG-gneiss; 12—faults.
Minerals 12 01143 g001
Figure 2. Geological sketch map of the Leksa deposit (compiled using the data of Kuleshevich (1992) and Kuleshevich et al. (2005) [40,41] with some modifications and additions). 1—Quaternary sediments, 2—quartz–carbonate–albite–sericite schists, 3—carbon-bearing schists, 4—sulfide ores, 5—silicification zone, with chloritoid blastesis (Chl), with fuchsite, 6—attitude, 7—metasomatic alteration zones (chloritization). Yellow dots on drillhole logs indicate sampling locations and depth below the surface.
Figure 2. Geological sketch map of the Leksa deposit (compiled using the data of Kuleshevich (1992) and Kuleshevich et al. (2005) [40,41] with some modifications and additions). 1—Quaternary sediments, 2—quartz–carbonate–albite–sericite schists, 3—carbon-bearing schists, 4—sulfide ores, 5—silicification zone, with chloritoid blastesis (Chl), with fuchsite, 6—attitude, 7—metasomatic alteration zones (chloritization). Yellow dots on drillhole logs indicate sampling locations and depth below the surface.
Minerals 12 01143 g002
Figure 3. BSE images of pyrite grains from drill cores of the Leksa deposit: (A) concentrically laminated pyrite. Note that euhedral pyrite crystals (approximately 0.5 mm in diameter and less) occur around the laminated grain; (B) Disseminated euhedral pyrite crystals; (C) Micro-sized euhedral pyrite crystals observed as ring-shaped aggregates; Image (D) shows magnified view of marked area in (C).
Figure 3. BSE images of pyrite grains from drill cores of the Leksa deposit: (A) concentrically laminated pyrite. Note that euhedral pyrite crystals (approximately 0.5 mm in diameter and less) occur around the laminated grain; (B) Disseminated euhedral pyrite crystals; (C) Micro-sized euhedral pyrite crystals observed as ring-shaped aggregates; Image (D) shows magnified view of marked area in (C).
Minerals 12 01143 g003
Figure 4. Δ33S vs. δ34S values for colloform and euhedral pyrite studied here. Data from Table A1 in Appendix A. Uncertainties for the data shown are smaller than the symbol sizes for δ34S and ∆33S.
Figure 4. Δ33S vs. δ34S values for colloform and euhedral pyrite studied here. Data from Table A1 in Appendix A. Uncertainties for the data shown are smaller than the symbol sizes for δ34S and ∆33S.
Minerals 12 01143 g004
Figure 5. Δ36S vs. Δ33S values for a subset of the euhedral pyrite samples. Data from Table A1 in Appendix A. The dashed line is linear fit to data with a slope of −1.15.
Figure 5. Δ36S vs. Δ33S values for a subset of the euhedral pyrite samples. Data from Table A1 in Appendix A. The dashed line is linear fit to data with a slope of −1.15.
Minerals 12 01143 g005
Figure 6. The Mesoarchean sedimentary sulfide δ34S and ∆33S data. Grey unfilled circles are literature data for Mesoarchean samples from Southern Africa and Western Australia [23,26,31]. Filled black circles are data representing in this study in Table A1. Also shown is the dashed line represented the Archean reference array (∆33S = 0.9 δ34S).
Figure 6. The Mesoarchean sedimentary sulfide δ34S and ∆33S data. Grey unfilled circles are literature data for Mesoarchean samples from Southern Africa and Western Australia [23,26,31]. Filled black circles are data representing in this study in Table A1. Also shown is the dashed line represented the Archean reference array (∆33S = 0.9 δ34S).
Minerals 12 01143 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vysotskiy, S.V.; Velivetskaya, T.A.; Ignatiev, A.V.; Slabunov, A.I.; Aseeva, A.V. Multiple Sulfur Isotope Evidence for Bacterial Sulfate Reduction and Sulfate Disproportionation Operated in Mesoarchaean Rocks of the Karelian Craton. Minerals 2022, 12, 1143. https://0-doi-org.brum.beds.ac.uk/10.3390/min12091143

AMA Style

Vysotskiy SV, Velivetskaya TA, Ignatiev AV, Slabunov AI, Aseeva AV. Multiple Sulfur Isotope Evidence for Bacterial Sulfate Reduction and Sulfate Disproportionation Operated in Mesoarchaean Rocks of the Karelian Craton. Minerals. 2022; 12(9):1143. https://0-doi-org.brum.beds.ac.uk/10.3390/min12091143

Chicago/Turabian Style

Vysotskiy, Sergey V., Tatyana A. Velivetskaya, Aleksandr V. Ignatiev, Aleksandr I. Slabunov, and Anna V. Aseeva. 2022. "Multiple Sulfur Isotope Evidence for Bacterial Sulfate Reduction and Sulfate Disproportionation Operated in Mesoarchaean Rocks of the Karelian Craton" Minerals 12, no. 9: 1143. https://0-doi-org.brum.beds.ac.uk/10.3390/min12091143

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop