Hostname: page-component-cb9f654ff-fg9bn Total loading time: 0 Render date: 2025-08-27T22:56:04.111Z Has data issue: false hasContentIssue false

The aerobic biosphere as an O2 sink before the Great Oxygenation Event: geobiological feedback to solid Earth and surface oxidation

Published online by Cambridge University Press:  01 August 2025

Eric Runge*
Affiliation:
Geoscience Center, https://ror.org/01y9bpm73 University of Göttingen , Göttingen, Germany Department of Biology and Environmental Science, Linnaeus University, Kalmar, Sweden
Sara Vulpius
Affiliation:
Institute of Geological Sciences, https://ror.org/046ak2485 Freie Universität Berlin , Berlin, Germany
Daniel Herwartz
Affiliation:
Department of Geosciences, https://ror.org/01y9bpm73 Ruhr University Bochum , Bochum, Germany
Andreas Pack
Affiliation:
Geoscience Center, https://ror.org/01y9bpm73 University of Göttingen , Göttingen, Germany
Caroline Brachmann
Affiliation:
Institute of Geological Sciences, https://ror.org/046ak2485 Freie Universität Berlin , Berlin, Germany German Aerospace Center, Berlin, Germany
Lena Noack
Affiliation:
Institute of Geological Sciences, https://ror.org/046ak2485 Freie Universität Berlin , Berlin, Germany
*
Corresponding author: Eric Runge; Email: eric.runge@lnu.se
Rights & Permissions [Opens in a new window]

Abstract

Microbial O2 production via oxygenic photosynthesis was vital in oxygenating the Earth’s surface environment during the Great Oxygenation Event (GOE) ca. 2.5 to 2.3 billion years ago. However, geochemical, paleontological and genomic data suggest the emergence of oxygenic photosynthesis precedes the GOE by at least 500 million years. This demonstrates that the first appearance of microbial O2 in the environment cannot explain the timing of atmospheric oxygenation. Instead, the GOE was facilitated by Earth’s geodynamic evolution, expanding cyanobacterial habitats and the changing redox state of the mantle, decreasing the abundance of reduced surface rocks, volcanic gases and aqueous solutes. These trends ultimately resulted in magnified O2 production rates and diminished O2 consumption rates. Thus, the GOE can be understood as a misbalance between O2 sources and sinks. One of the most critical O2 sinks on modern Earth is microbial O2 consumption via aerobic respiration, and accumulating evidence suggests its emergence well before the GOE. However, the role of aerobic microorganisms as an O2 sink delaying the GOE remains poorly explored. Here, we review the redox evolution of Earth’s mantle and surface environments, as well as the Archean evolution of aerobic microbial metabolisms. Oxygenic photosynthesis released O2 to the environment, but the secular oxidation of the solid Earth was critical in allowing O2 accumulation. Aerobic respiration expanded in response to the GOE, but our survey suggests it could have been a critical O2 sink even earlier. Hence, aerobic respiration can be seen as geobiological feedback to changes in the Earth system from deep in the mantle up to the surface. However, the timing and rate of O2 consumption by aerobic respiration before the GOE remain poorly constrained. We conclude by highlighting open questions and future research directions to understand the role of the aerobic O2 sink in delaying the GOE.

Information

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of The Mineralogical Society of the United Kingdom and Ireland

Introduction

The Great Oxygenation Event (GOE) occurred 2.5 to 2.3 billion years ago (Ga) and was one of the most significant revolutions in the Earth system. It was marked by an increase in atmospheric O2 by several orders of magnitude sourced from oxygenic photosynthesis (Lyons et al., Reference Lyons, Diamond, Planavsky, Reinhard and Li2021) (Table 1). The accumulation of atmospheric O2 changed the redox state of Earth’s surface environment (Lyons et al., Reference Lyons, Reinhard and Planavsky2014; Ostrander et al., Reference Ostrander, Johnson and Anbar2021), the mineralogical composition of the Earth (Hazen et al., Reference Hazen, Papineau, Bleeker, Downs, Ferry, McCoy, Sverjensky and Yang2008) and allowed for major biological innovations, including the much later evolution of eukaryotic organisms and the Cambrian explosion (David and Alm, Reference David and Alm2011; Zhang et al., Reference Zhang, Shu, Han, Zhang, Liu and Fu2014; Mills et al., Reference Mills, Boyle, Daines, Sperling, Pisani, Donoghue and Lenton2022). This also has astrobiological implications because high O2 concentrations in an exoplanet’s atmosphere could point to the existence of an oxygenic biosphere, although alternative abiotic explanations for O2 production exist (Meadows et al., Reference Meadows2018; Schwieterman et al., Reference Schwieterman2018). For these reasons, understanding the driving mechanisms of the GOE is important across various disciplines within the Earth and life sciences.

Box 1: Definition of the redox state and related terms

The term ‘redox state’ describes the oxidation potential of a system. In the context of this study, the system is the Earth mantle, crust, hydrosphere and atmosphere. ‘Oxidising conditions’ mean that elements, which occur in different oxidation states (e.g. Fe, Mn, Cr, S, C) predominantly occur in the oxidised state, e.g. Fe3+ or Mn4+, whereas under ‘reducing conditions’, Fe predominantly occurs as Fe2+ and Mn as Mn2+. The oxidation state can quantitatively be expressed in form of the oxygen fugacity ƒO2, which is approximately equal to the equilibrium oxygen partial pressure pO2. In reducing systems, ƒO2 is low; in oxidising systems, ƒO2 is high. Because the absolute numbers for ƒO2 are very low (e.g. 10–10), one expresses ƒO2 as log10(ƒO2). An important buffer for the O2 fugacity within planets is the Fe–FeO (‘iron-wüstite’, IW) buffer with:

$$ \mathrm{Fe}+\frac{1}{2}{\mathrm{O}}_2=\mathrm{FeO} $$

Metallic iron (Fe) occurs in Fe–Ni alloys and FeO as a component in many silicate minerals. As long as Fe and FeO are present, ƒO2 is fixed. For the reaction shown, the ƒO2 would be fixed by the IW buffer. The absolute ƒO2 varies with temperature and pressure. Oxygen fugacities that deviate from the ƒO2 buffered by the IW buffer are conventionally expressed as ƒO2 deviating in log10 units from ƒO2 buffered by the IW buffer. An ƒO2 that is two orders of magnitude lower (factor of 0.01) than buffered by the IW buffer at a given temperature would be termed log10(ƒO2) = IW – 2. Other O2 buffers exist, such as the quartz-fayalite-magnetite (QFM) buffer. However, the redox state is conventionally expressed in log10 units relative to ƒO2 buffered by the IW buffer. The most reduced system known is the H2-dominated solar nebula, where the ƒO2 was buffered by the H2–H2O equilibrium to IW – 7. Modern rocks have a redox state in the range buffered by the QFM buffer, which is at about IW + 3.7.

The ƒO2 determines the speciation of volatiles. If the ƒO2 is low, the system is reducing, meaning that reduced species such as H2, CO, CH4, H2S and NH3 prevail. In contrast, if the ƒO2 is high, the system is oxidising and oxidised species like H2O, CO2, SO2 and N2 are dominant (Kasting, Reference Kasting1993; Kasting et al., Reference Kasting1993; Ballhaus and Frost, Reference Ballhaus and Frost1994; Holloway and Blank, Reference Holloway, Blank, Carroll and Holloway1994; Delano, Reference Delano2001; Burgisser and Scaillet, Reference Burgisser and Scaillet2007; Trail et al., Reference Trail, Watson and Tailby2011; Gaillard et al., Reference Gaillard, Scaillet, Pichavant and Iacono-Marziano2015, Reference Gaillard, Bouhifd, Füri, Malavergne, Marrocchi, Noack, Ortenzi, Roskosz and Vulpius2021; Ortenzi et al., Reference Ortenzi2020; Yang et al., Reference Yang, Liu and Zhang2022). This demonstrates that the redox state is a fundamentally important parameter for the evolution of the Earth interior and surface system.

Table 1. Net reactions of microbial metabolism discussed in this paper (Konhauser, Reference Konhauser2007).

Much attention has been given to dating the emergence of oxygenic photosynthesis. Attempts at doing so included various putative biosignatures, like stromatolites in photic paleoenvironments, microfossils of cyanobacteria, carbon isotope signatures of photoautotrophic carbon fixation, lipid biomarkers in Archean rocks, as well as biogeochemical models (e.g. Buick, Reference Buick1992; Schopf, Reference Schopf1993; Mojzsis et al., Reference Mojzsis, Arrhenius, McKeegan, Harrison, Nutman and Friend1996; Hofmann et al., Reference Hofmann, Grey, Hickman and Thorpe1999; Brocks, Reference Brocks1999; Schidlowski, Reference Schidlowski2001; Kopp et al., Reference Kopp, Kirschvink, Hilburn and Nash2005). Many of these approaches are regarded as controversial (Brasier et al., Reference Brasier, Green, Lindsay, Mcloughlin, Steele and Stoakes2005; Rasmussen et al., Reference Rasmussen, Fletcher, Brocks and Kilburn2008; French et al., Reference French2015). These controversies have yielded a broad timespan for the possible emergence of oxygenic photosynthesis (ca. 3.5 to ca. 2.4 Ga). Molecular clock studies add a relatively recent approach to the problem. Using calibration points from the rock record, these studies point to an emergence of oxygenic photosynthesis at ca. 3.0 Ga, i.e. several hundred million years (m.y.) before the GOE (Schirrmeister et al., Reference Schirrmeister, Gugger and Donoghue2015; Sánchez-Baracaldo, Reference Sánchez-Baracaldo2015; Cardona et al., Reference Cardona, Sánchez-Baracaldo, Rutherford and Larkum2019; Garcia-Pichel et al., Reference Garcia-Pichel, Lombard, Soule, Dunaj, Wu and Wojciechowski2019; Jabłońska and Tawfik, Reference Jabłońska and Tawfik2021; Fournier et al., Reference Fournier, Moore, Rangel, Payette, Momper and Bosak2021; Boden et al., Reference Boden, Zhong, Anderson and Stüeken2024; but see Soo et al., Reference Soo, Hemp, Parks, Fischer and Hugenholtz2017). This timing is compatible with geochemical proxies indicating oxidative weathering in situ benthic microbial mats, local O2 levels of few to few tens of μM, or transient “whiffs” of O2 (reviewed in Ostrander et al., Reference Ostrander, Johnson and Anbar2021). Apparently, Earth’s atmosphere remained anoxic for at least 500 m.y. while oxygenic photosynthesis was already occurring. This delay of the GOE is one of its central conundrums.

Suggested ideas to solve this problem involve either an increasing rate of biological O2 production or a decreasing O2 consumption rate during the 500 m.y. preceding the GOE (e.g. Konhauser et al., Reference Konhauser2017; Catling and Zahnle, Reference Catling and Zahnle2020; Lyons et al., Reference Lyons, Tino, Fournier, Anderson, Leavitt, Konhauser and Stüeken2024). Thus, the GOE can be understood as the tipping point reached when O2 production rates by oxygenic photosynthesis exceeded the O2 consumption rates of all sinks. These sinks include reduced species such as volcanic gases in the atmosphere (e.g. H2, CH4), aqueous solutes (e.g. Fe2+, Mn2+), minerals in surface rocks (e.g. pyrite, uraninite) or sedimentary organic matter. The capacities of these sinks are, to a large degree, constrained by the redox evolution of the solid Earth. However, microorganisms have modulated the balance of Earth’s redox buffers by catalysing otherwise inhibited chemical reactions since life emerged more than 3.5 Ga (e.g. Falkowski et al., Reference Falkowski, Fenchel and Delong2008; Knoll et al., Reference Knoll, Bergmann and Strauss2016; Ostrander et al., Reference Ostrander, Johnson and Anbar2021; Runge et al., Reference Runge, Mansor, Kappler and Duda2023). Thus, the Earth’s buffering capacity against oxygenation can only be understood by integrating abiotic and biotic processes from deep in the mantle to the surface. Notably, advances in experimental microbiology and microbial ecology showed that microorganisms in diverse environments consume O2 below the canonical lower limit for aerobic respiration (i.e. the ‘Pasteur point’, 2.2 μM O2 at 25°C in seawater, e.g. Stolper et al., Reference Stolper, Revsbech and Canfield2010; Berg et al., Reference Berg2019; Ruff et al., Reference Ruff2023) (Table 1). Given the emergence of oxygenic photosynthesis at ca. 3.0 Ga, it seems plausible that the early production of biological O2 created aerobic niches since the mid-Archean. Nevertheless, the role of aerobic microorganisms as an O2 sink delaying the GOE remains poorly understood.

Here, we review the redox evolution of the solid Earth and its surface environments from planetary accretion to the GOE, aiming to explore the role of microbial O2 sinks in the Archean Earth. First, we reconstruct the solid Earth’s redox evolution, which constrains the capacity of its abiotic redox buffers and sets the stage on which microbial life proliferates. Then, we address the oxygenation of Earth’s surface environments, including the atmosphere and hydrosphere. Finally, we review evidence for the role of microbial O2 sources and sinks in the Archean. We highlight the complex interplay of abiotic and biotic processes in the substantial delay from the first biological O2 production to the onset of atmospheric oxygenation. Our survey suggests that the Earth’s aerobic biosphere is a crucial yet poorly understood Archean O2 sink that must be better quantified to unravel the delay of the GOE.

The delayed GOE: asynchronous solid Earth and surface oxidation

The Earth’s atmosphere and hydrosphere evolved from outgassing and condensation of volatiles from the mantle, therefore the solid Earth sets the stage for the evolution of the surface reservoir. It represents the reservoir from which the lithosphere, atmosphere, hydrosphere and biosphere evolved, thereby defining the Earth’s overall buffering capacity against oxygenation. However, the subsequent evolution of Earth’s redox state is also closely coupled to the evolution of life in its surface environments. This section reviews the deep-time redox evolution of Earth’s interior and surface reservoirs.

The evolving redox state of the solid Earth in deep time

The ƒO2 of Earth’s present-day upper mantle is QFM ± 2 (Fig. 1, see Box 1), but it decreases with depth (e.g. Haggerty, Reference Haggerty1978; Christie et al., Reference Christie, Carmichael and Langmuir1986; O’Neill and Wall, Reference St.C and Wall1987; Wood and Virgo, Reference Wood and Virgo1989; Wood et al., Reference Wood, Bryndzia and Johnson1990; Ballhaus et al., Reference Ballhaus, Berry and Green1991; O’Neill, Reference St.C1991; Holloway et al., Reference Holloway, Pan and Gudmundsson1992; Kasting, Reference Kasting1993; McCammon, Reference McCammon2005; Frost and McCammon, Reference Frost and McCammon2008; Cottrell and Kelley, Reference Cottrell and Kelley2011; Trail et al., Reference Trail, Watson and Tailby2011; Ardia et al., Reference Ardia, Hirschmann, Withers and Stanley2013; Gaillard et al., Reference Gaillard, Bouhifd, Füri, Malavergne, Marrocchi, Noack, Ortenzi, Roskosz and Vulpius2021; Yang et al., Reference Yang, Liu and Zhang2022) (Fig. 2). The transition zone is assumed to have an ƒO2 of about QFM – 4 (McCammon, Reference McCammon2005; Frost and McCammon, Reference Frost and McCammon2008; Ardia et al., Reference Ardia, Hirschmann, Withers and Stanley2013; Yang et al., Reference Yang, Liu and Zhang2022) and the lower mantle is supposed to have an ƒO2 below QFM – 5 (Frost et al., Reference Frost, Liebske, Langenhorst, McCammon, Trønnes and Rubie2004; McCammon, Reference McCammon2005; Ardia et al., Reference Ardia, Hirschmann, Withers and Stanley2013; Yang et al., Reference Yang, Liu and Zhang2022).

Figure 1. Commonly used mineral redox buffers and their relationship to ƒO2 plotted over temperature. Quartz-fayalite-magnetite QFM) and nickel-nickel-oxide (NiNiO) depict oxidised conditions, while iron-wustite (IW) and quartz-iron-fayalite (QIF) represent reduced conditions.

Figure 2. Evolution of the Earth´s redox state for different formation stages. The changing ƒO2 is indicated by the deviation in log units from the quartz-fayalite-magnetite (QFM) buffer and is explained in the text. The colours range from blue (reduced) to red (oxidised). The Earth is assumed to become more oxidised with time, with the most reduced values during the accretion period before core formation. It is thought that during the magma ocean period, ƒO2 evolved towards more oxidised values. The modern Earth is comparatively oxidised, with a decreasing redox state with depth (after McCammon, Reference McCammon2005). See the text for references on the redox state of the early Earth.

In contrast to the present day, it is commonly thought that the Earth was initially reduced, with an ƒO2 of IW – 2 (ca. QFM – 5.7) as a maximum value during core formation (before that, the ƒO2 could have been even as low as IW – 5 or about QFM – 8.7; Wade and Wood, Reference Wade and Wood2001, Reference Wade and Wood2005; Rubie et al., Reference Rubie, Frost, Mann, Asahara, Nimmo, Tsuno, Kegler, Holzheid and Palme2011, Reference Rubie, Jacobson, Morbidelli, O’Brien, Young, De Vries, Nimmo, Palme and Frost2015; Scaillet and Gaillard, Reference Scaillet and Gaillard2011; Cartier et al., Reference Cartier, Hammouda, Boyet, Bouhifd and Devidal2014; Fischer et al., Reference Fischer, Nakajima, Campbell, Frost, Harries, Langenhorst, Miyajima, Pollok and Rubie2015; Schaefer and Elkins-Tanton, Reference Schaefer and Elkins-Tanton2018; Gaillard et al., Reference Gaillard, Bouhifd, Füri, Malavergne, Marrocchi, Noack, Ortenzi, Roskosz and Vulpius2021) (Fig. 2). One argument for a reducing start of the Earth is the assumption that it initially accreted from highly reduced, volatile-depleted material (like enstatite chondrites). Another argument is that the metal–silicate equilibrium required for core formation suggests a considerably low ƒO2. A low ƒO2 enhances the siderophile behaviour of certain elements like nickel, cobalt, manganese, chromium, vanadium and silicon (Gessmann et al., Reference Gessmann, Rubie and McCammon1999; Wade and Wood, Reference Wade and Wood2001, Reference Wade and Wood2005; Rubie et al., Reference Rubie, Frost, Mann, Asahara, Nimmo, Tsuno, Kegler, Holzheid and Palme2011, Reference Rubie, Jacobson, Morbidelli, O’Brien, Young, De Vries, Nimmo, Palme and Frost2015; Scaillet and Gaillard, Reference Scaillet and Gaillard2011; Siebert et al., Reference Siebert, Badro, Antonangeli and Ryerson2013; Cartier et al., Reference Cartier, Hammouda, Boyet, Bouhifd and Devidal2014; Fischer et al., Reference Fischer, Nakajima, Campbell, Frost, Harries, Langenhorst, Miyajima, Pollok and Rubie2015; Schaefer and Elkins-Tanton, Reference Schaefer and Elkins-Tanton2018; Gaillard et al., Reference Gaillard, Bouhifd, Füri, Malavergne, Marrocchi, Noack, Ortenzi, Roskosz and Vulpius2021; but see Badro et al. (Reference Badro, Brodholt, Piet, Siebert and Ryerson2015) for an alternative viewpoint).

Assuming reduced conditions early in Earth’s history, its relatively oxidised state today requires oxidation over time (e.g. Wade and Wood, Reference Wade and Wood2005; Cartier et al., Reference Cartier, Hammouda, Boyet, Bouhifd and Devidal2014; Schaefer and Elkins-Tanton, Reference Schaefer and Elkins-Tanton2018; Gaillard et al., Reference Gaillard, Bouhifd, Füri, Malavergne, Marrocchi, Noack, Ortenzi, Roskosz and Vulpius2021). Pahlevan et al. (Reference Pahlevan, Schaefer and Hirschmann2019) found that if the magma ocean was initially reduced, it must have evolved to a higher ƒO2 (>IW + 1 or ~QFM – 2.7) during its final stages based on the D/H ratio. Moreover, Deng et al. (Reference Deng, Du, Karki, Ghosh and Lee2020) suggested that the magma ocean had a vertical gradient in ƒO2, with the upper layer reaching IW + 2 (~QFM – 1.7). However, it was also proposed that more oxidised, volatile-rich material (e.g. CI chondrites) was delivered during the last stages of accretion and core formation (e.g. Wänke et al., Reference Wänke, Dreibus, Jagoutz, Kröner, Hanson and Goodwin1984; Javoy, Reference Javoy1995; Wade and Wood, Reference Wade and Wood2005; Schönbächler et al., Reference Schönbächler, Carlson, Horan, Mock and Hauri2010; Rubie et al., Reference Rubie, Frost, Mann, Asahara, Nimmo, Tsuno, Kegler, Holzheid and Palme2011; Scaillet and Gaillard, Reference Scaillet and Gaillard2011; Marty, Reference Marty2012; Cartier et al., Reference Cartier, Hammouda, Boyet, Bouhifd and Devidal2014; Fischer et al., Reference Fischer, Nakajima, Campbell, Frost, Harries, Langenhorst, Miyajima, Pollok and Rubie2015; Dauphas, Reference Dauphas2017; Fischer-Gödde and Kleine, Reference Fischer-Gödde and Kleine2017; Lammer et al., Reference Lammer2018; Grewal et al., Reference Grewal, Dasgupta, Sun, Tsuno and Costin2019; Budde et al., Reference Budde, Burkhardt and Kleine2019; Fischer-Gödde et al., Reference Fischer-Gödde, Elfers, Münker, Szilas, Maier, Messling, Morishita, Van Kranendonk and Smithies2020; Gaillard et al., Reference Gaillard, Bouhifd, Füri, Malavergne, Marrocchi, Noack, Ortenzi, Roskosz and Vulpius2021). Rubie et al. (Reference Rubie, Frost, Mann, Asahara, Nimmo, Tsuno, Kegler, Holzheid and Palme2011) concluded that 30–40% of the final mass accreted was rather oxidised, therefore the evolution towards a more oxidised planet probably occurred already during the formation of the Earth.

Besides the variation in the delivered material, the change in the Earth´s redox state during accretion can also be explained by the increasing size of the Earth (Wade and Wood, Reference Wade and Wood2005). It was proposed that due to the higher pressures associated with the growth of the Earth, perovskite (Mg,Fe,Al)(Al,Si)O3) becomes the dominant phase of the lower mantle (stable below 660 km depth in present-day Earth). Perovskite formation drives Fe(II) disproportionation to Fe(III) and Fe(0) via Eq. (1):

(1) $$ 3{\mathrm{Fe}}^{2+}\mathrm{O}+{\mathrm{Al}}_2{\mathrm{O}}_3=2{\mathrm{Fe}}^{3+}{\mathrm{Al}\mathrm{O}}_3+{\mathrm{Fe}}^0 $$

Because Fe(0) has been sequestered into the core, the lower mantle became relatively enriched in Fe(III). It was suggested that the upper mantle became enriched over time due to convection (Mao and Bell, Reference Mao, Bell, Saxena, Bhattacharji, Annersten and Stephansson1977; Frost et al., Reference Frost, Liebske, Langenhorst, McCammon, Trønnes and Rubie2004, Reference Frost, Mann, Asahara and Rubie2008; Wade and Wood, Reference Wade and Wood2005). This transfer of Fe(III) from the lower to the upper mantle is also known as the ‘oxygen pump’ (Frost et al., Reference Frost, Liebske, Langenhorst, McCammon, Trønnes and Rubie2004, Reference Frost, Mann, Asahara and Rubie2008; Wade and Wood, Reference Wade and Wood2005). This process would also explain why the Martian mantle is more reduced than Earth (Righter and Drake, Reference Righter and Drake1996; Herd et al., Reference Herd, Papike and Brearley2001, Reference Herd, Borg, Jones and Papike2002; Wadhwa, Reference Wadhwa2001, Reference Wadhwa2008; Wade and Wood, Reference Wade and Wood2005; Righter et al., Reference Righter, Yang, Costin and Downs2008), even though Mars is enriched in volatiles and FeO. Perovskite is unstable in the Martian mantle due to its smaller size, resulting in lower lithostatic pressures, thus the self-oxidation of the mantle via perovskite formation could not occur on Mars (Wade and Wood, Reference Wade and Wood2005). It was proposed that the preferential partitioning of Fe(III) into the liquid phase enhances the equilibration of the redox state between the lower and upper mantle by mixing processes (Carmichael, Reference Carmichael1991; Scaillet and Gaillard, Reference Scaillet and Gaillard2011). In addition, FeO and FeO1.5 have different molar volumes and densities, further favouring a more oxidised upper mantle and a more reduced lower mantle (Deng et al., Reference Deng, Du, Karki, Ghosh and Lee2020).

Alternatively, it was suggested that the oxidation of the mantle occurred directly during the magma ocean state (Schaefer and Elkins-Tanton, Reference Schaefer and Elkins-Tanton2018; Pahlevan et al., Reference Pahlevan, Schaefer and Hirschmann2019). Accordingly, the sink and sequestration of iron metal into the core would leave behind an oxidised mantle without requiring crystallisation and whole-scale mantle mixing (Schaefer and Elkins-Tanton, Reference Schaefer and Elkins-Tanton2018; Pahlevan et al., Reference Pahlevan, Schaefer and Hirschmann2019). In particular, the crystallising magma ocean would become progressively oxidised over time (Scaillet and Gaillard, Reference Scaillet and Gaillard2011). Similar arguments have been brought forward for a carbon pump leading to the formation of diamonds in the lower mantle (causing oxidation) in case of a deep (potentially giant-impact-induced) magma ocean, which may explain the thick CO2 atmosphere of Venus in the absence of a late giant impact (Armstrong et al., Reference Armstrong, Frost, McCammon, Rubie and Boffa Ballaran2019). Moreover, H2 loss from the mantle by outgassing is also discussed as a mechanism for oxidising the upper mantle (Sharp et al., Reference Sharp, McCubbin and Shearer2013).

It has also been suggested that recycling of surface material could have oxidised the upper part of the mantle (Arculus, Reference Arculus1985; Kasting, Reference Kasting1993; Kasting et al., Reference Kasting1993; Kump et al., Reference Kump, Kasting and Barley2001; Smart et al., Reference Smart, Tappe, Stern, Webb and Ashwal2016; Nicklas et al., Reference Nicklas, Puchtel, Ash, Piccoli, Hanski, Nisbet, Waterton, Pearson and Anbar2019; Stagno and Aulbach, Reference Stagno, Aulbach, Moretti and Neuville2021). At least today, the material transported with the subducting slab is more oxidised than the surrounding mantle (e.g. Wood et al., Reference Wood, Bryndzia and Johnson1990; Ballhaus et al., Reference Ballhaus, Berry and Green1991; Blundy et al., Reference Blundy, Brodholt and Wood1991). Mikhail and Sverjensky (Reference Mikhail and Sverjensky2014) found that under oxidising conditions, N2 is the dominant nitrogen species over NH4+. They argue that, during subduction, the increased ƒO2 of the mantle wedges, compared to the surrounding upper mantle, results in N2-rich fluids. The ascent and outgassing of such fluids allow an enhanced N2 outgassing. Plate tectonics would, therefore, not only favour oxidised mantles and atmospheres but would also be needed for nitrogen-rich atmospheres like the Earth’s.

Duncan and Dasgupta (Reference Duncan and Dasgupta2017) turned the argument around: if reduced material (like organic carbon) was subducted, then this may have led to a transient increase of biological O2 in the atmosphere by removing reducing power from the surface reservoir. At the same time, it would lead to a reducing effect on the mantle (unless permanently sequestered into a hidden reservoir) and result in releasing reducing gases into the atmosphere on melting.

The estimates on when the upper mantle was oxidised to near-modern values range from 4.4 to 2.7 Ga (e.g. Canil, Reference Canil1997; Delano, Reference Delano2001; Lee et al., Reference Lee, Brandon and Norman2003; Li and Lee, Reference Li and Lee2004; Foley, Reference Foley2011; Scaillet and Gaillard, Reference Scaillet and Gaillard2011; Trail et al., Reference Trail, Watson and Tailby2011; Aulbach and Stagno, Reference Aulbach and Stagno2016; Rollinson et al., Reference Rollinson, Adetunji, Lenaz and Szilas2017; Nicklas et al., Reference Nicklas, Puchtel and Ash2018, Reference Nicklas, Puchtel, Ash, Piccoli, Hanski, Nisbet, Waterton, Pearson and Anbar2019) (Figs 3, 4a). An important archive for understanding the redox state of the early Earth’s mantle is the cerium concentration in zircons (ZrSiO4) (Loucks et al., Reference Loucks, Fiorentini and Henríquez2020). Cerium exists in both tri- and quadrivalent states in silicate melts. Zircons crystallising from these melts preferentially incorporate Ce4+ over Ce3+, substituting for Zr4+ in the zircon crystal structure. The ratio of Ce4+ to Ce3+ is influenced by the ƒO2 of the melt. As a result, the cerium concentration in magmatic zircons can indicate the oxygen content in the magma (Trail et al., Reference Trail, Watson and Tailby2011). Trail et al. (Reference Trail, Watson and Tailby2011) calibrated the relationship between the zircon/melt partitioning coefficient of cerium and the ƒO2 of the melt. Using their different oxygen isotopic compositions, Trail et al. (Reference Trail, Watson and Tailby2011) distinguished zircons derived from the mantle (δ18O = +5.3‰) and those from the crust. The cerium concentration data of both populations indicated that the host magmas had similar ƒO2 to the modern mantle, which has QFM ± 2 (Yang et al., Reference Yang, Liu and Zhang2022). It was observed that the primary mantle melts were not saturated in zircon, but rather, the ‘mantle’ zircons crystallised in melt residues. The δ18O values of these residues would still closely resemble the composition of the host mantle. At such ƒO2 values, the outgassing of CO2, N2, H2O, and SO2 would dominate over more reduced species like CO, H2, NH3 and H2O (Frost and McCammon, Reference Frost and McCammon2008).

Figure 3. Literature estimates regarding the ƒO2 of the upper mantle on early Earth (after Aulbach and Stagno, Reference Aulbach and Stagno2016; Schaefer and Elkins-Tanton, Reference Schaefer and Elkins-Tanton2018; Stagno and Aulbach, Reference Stagno, Aulbach, Moretti and Neuville2021). The shaded rectangles represent the results of the ƒO2 estimated from individual studies (normalised to the QFM buffer) over the respective ages of the samples examined. The references for the horizontal rectangles are Canil (Reference Canil1997), Delano (Reference Delano2001), Li and Lee (Reference Li and Lee2004), Nicklas et al. (Reference Nicklas, Puchtel and Ash2018, Reference Nicklas, Puchtel, Ash, Piccoli, Hanski, Nisbet, Waterton, Pearson and Anbar2019), Aulbach and Stagno (Reference Aulbach and Stagno2016) (AS), Aulbach et al. (Reference Aulbach, Woodland, Vasilyev, Galvez and Viljoen2017). The crosses are data points from Nicklas et al. (Reference Nicklas, Puchtel and Ash2018) and Nicklas et al. (Reference Nicklas, Puchtel, Ash, Piccoli, Hanski, Nisbet, Waterton, Pearson and Anbar2019) that represent an estimated ƒO2 based on the redox-dependent partitioning of vanadium between liquidus olivine and melt. The squares are orogenic eclogites, the circles are mantle eclogites and the diamond is a mid-ocean ridge ophiolite from Aulbach and Stagno (Reference Aulbach and Stagno2016). The symbols display the ƒO2 (corrected to 1 GPa) calculated from V/Sc ratios. The vertical error bars are predicted 1σ errors of the V/Sc ratios (representing 1σ of the mean per sample suite) and the horizontal error bars show age ranges or 1σ errors for isochron ages from the literature. The red star shows the calculated ƒO2 of the modern MORB and the arrow points toward the estimated ƒO2 of the uppermost mantle according to Trail et al. (Reference Trail, Watson and Tailby2011) of samples from 4.4 Ga. QFM, quartz-fayalite-magnetite.

The oxygen isotope ratios in these up to 4.4 Ga zircons have provided valuable insights into the history of early Earth’s water cycle. Studies by Peck et al. (Reference Peck, Valley, Wilde and Graham2001), Valley et al. (Reference Valley, Peck, King and Wilde2002) and Cavosie et al. (Reference Cavosie, Valley and Wilde2005) also examined oxygen isotope ratios in Hadean zircon, some of which yield elevated δ18O values consistent with the assimilation of sediments or crustal material derived from low-temperature water–rock interactions. The isotopic signatures found in these zircons thus suggest the presence of liquid water on the Earth’s surface during the Hadean as early as 4.4 Ga. These findings support the presence of a hydrosphere on Earth at that time. If correct, the δ18O of the hydrosphere needs to be considered because the δ18O of the early oceans may have been lower than the present oceans (Wallmann, Reference Wallmann2001; Sengupta and Pack, Reference Sengupta and Pack2018; Herwartz et al., Reference Herwartz, Pack and Nagel2021; Tatzel et al., Reference Tatzel, Frings, Oelze, Herwartz, Lünsdorf and Wiedenbeck2022; Isson and Rauzi, Reference Isson and Rauzi2024) and meteoric water generally comprises lower δ18O than seawater. Due to the variability of δ18O in the hydrosphere, water–rock interaction at high and low temperatures can result in a large range of silicate δ18O, and assimilation of such altered material is also known to generate low δ18O magmas (Bindeman et al., Reference Bindeman, Schmitt and Evans2010; Herwartz et al., Reference Herwartz, Pack, Krylov, Xiao, Muehlenbachs, Sengupta and Di Rocco2015; Zakharov et al., Reference Zakharov, Bindeman, Tanaka, Friðleifsson, Reed and Hampton2019). In general, assimilation of the altered mafic crust may not lead to elevated δ18O of magmas from which the zircons crystallised, therefore a mantle-like δ18O of Hadean zircons may be taken cautiously as an argument for the mantle origin of the zircons. An origin from the crust, however, would not allow any conclusions to be drawn about the redox state of the mantle.

Indeed, Hopkins et al. (Reference Hopkins, Harrison and Manning2008) suggested that the Hadean zircons they studied (which contained mineral inclusions) formed in a crustal setting. Based on the derived formation pressure and temperature (700°C, 7 kbar), they concluded that the surface heat flow from 4.2 to 4.0 Ga was only 75 mW/m2. Additionally, they proposed that the crustal zircon host melts may have formed above a subduction-like setting, where the subducting slab cools the underlying lithosphere. A similar conclusion was reached by Harrison et al. (Reference Harrison, Schmitt, McCulloch and Lovera2008), who suggested that Hadean zircons formed through crystallisation from crustal magmas. The negative ε(Hf,T) values observed in the set of zircons studied by Harrison et al. (Reference Harrison, Schmitt, McCulloch and Lovera2008) imply formation in a reservoir with sub-chondritic Lu/Hf (i.e. felsic crust), which may have formed as early as 4.5 Ga. Zircons that crystallise from a residual mantle melt should have positive ε(Hf,T) values.

Others used redox-sensitive elements like vanadium and chromium (e.g. Canil, Reference Canil1997; Delano, Reference Delano2001; Lee et al., Reference Lee, Brandon and Norman2003; Li and Lee, Reference Li and Lee2004; Aulbach and Viljoen, Reference Aulbach and Viljoen2015; Nicklas et al., Reference Nicklas, Puchtel and Ash2016, Reference Nicklas, Puchtel and Ash2018, Reference Nicklas, Puchtel, Ash, Piccoli, Hanski, Nisbet, Waterton, Pearson and Anbar2019; Aulbach and Stagno, Reference Aulbach and Stagno2016) or the Fe3+/(Fe3+ + Fe2+) ratio (Rollinson et al., Reference Rollinson, Adetunji, Lenaz and Szilas2017; Aulbach et al., Reference Aulbach, Woodland, Vasilyev, Galvez and Viljoen2017) to determine the redox state of early Earth’s mantle. Applying these methods, it was claimed that oxidation occurred early because since 3.9 to 3.5 Ga samples exhibit ƒO2 similar to modern mid-ocean ridge basalts (MORBs; Canil, Reference Canil1997; Delano, Reference Delano2001; Li and Lee, Reference Li and Lee2004; Rollinson et al., Reference Rollinson, Adetunji, Lenaz and Szilas2017). In contrast, more recent studies observed that Archean samples are still relatively reduced (QFM – 1.19 ± 0.33) compared to the post-Archean samples (including MORB: QFM – 0.26 ± 0.44). This observation hints at a transition from a relatively reduced towards an oxidised upper mantle during the mid or late Archean (Lee et al., Reference Lee, Brandon and Norman2003; Aulbach and Viljoen, Reference Aulbach and Viljoen2015; Aulbach and Stagno, Reference Aulbach and Stagno2016; Aulbach et al., Reference Aulbach, Woodland, Vasilyev, Galvez and Viljoen2017; Stagno and Fei, Reference Stagno and Fei2020) (Figs 3, 4a), which would have direct consequences on the volcanic outgassing efficiency and atmospheric evolution (Guimond et al., Reference Guimond, Noack, Ortenzi and Sohl2021).

The data shown in Fig. 3 suggest a gradual increase in the ƒO2 of the upper mantle observed from 3.0 to 2.0 Ga, which contrasts with a sudden increase proposed by previous studies (Canil, Reference Canil1997; Delano, Reference Delano2001; Li and Lee, Reference Li and Lee2004; Scaillet and Gaillard, Reference Scaillet and Gaillard2011; Rollinson et al., Reference Rollinson, Adetunji, Lenaz and Szilas2017). This discrepancy between a gradual and sudden increase in the redox sate is explained by heterogeneity of the early upper mantle due to incomplete mixing with the lower mantle, the addition of reduced meteoritic material or inherited from magma ocean processes (Ringwood, Reference Ringwood1979; Arculus, Reference Arculus1985; Nicklas et al., Reference Nicklas, Puchtel, Ash, Piccoli, Hanski, Nisbet, Waterton, Pearson and Anbar2019; Stagno and Fei, Reference Stagno and Fei2020; Stagno and Aulbach, Reference Stagno, Aulbach, Moretti and Neuville2021). Gu et al. (Reference Gu, Li, McCammon and Lee2016) experimentally demonstrated that oxidised lower mantle material is less dense than reduced lower mantle material. This enhances the ascent probability, leading to an efficient mixing between the lower and upper mantle. According to Gu et al. (Reference Gu, Li, McCammon and Lee2016), the upper mantle could have been oxidised within 800 m.y. via this mixing process. However, the process was probably prolonged due to the effect of the strength of bridgmanite (which is about three orders of magnitude higher compared to ferropericlase) on the mantle viscosity and, thus, on the mixing behaviour (Girard et al., Reference Girard, Amulele, Farla, Mohiuddin and Karato2016; Ballmer et al., Reference Ballmer, Houser, Hernlund, Wentzcovitch and Hirose2017; O’Neill and Aulbach, Reference O’Neill and Aulbach2022). Another reason for a delayed mantle mixing could have been a larger grain size resulting from hotter early Earth conditions. This larger grain size could have led to stronger plate boundaries, decreasing convective motion (Foley and Rizo, Reference Foley and Rizo2017). An inefficient mixing of the material from the lower mantle with the upper mantle would explain the preservation of primordial reservoirs suggested to explain observed isotope anomalies (e.g. Mukhopadhyay, Reference Mukhopadhyay2012; Debaille et al., Reference Debaille, O’Neill, Brandon, Haenecour, Yin, Mattielli and Treiman2013; Rizo et al., Reference Rizo, Boyet, Blichert-Toft and Rosing2013, Reference Rizo, Walker, Carlson, Touboul, Horan, Puchtel, Boyet and Rosing2016b, Reference Rizo, Walker, Carlson, Horan, Mukhopadhyay, Manthos, Francis and Jackson2016a; Girard et al., Reference Girard, Amulele, Farla, Mohiuddin and Karato2016; Ballmer et al., Reference Ballmer, Houser, Hernlund, Wentzcovitch and Hirose2017; Mundl et al., Reference Mundl, Touboul, Jackson, Day, Kurz, Lekic, Helz and Walker2017; Horan et al., Reference Horan, Carlson, Walker, Jackson, Garçon and Norman2018; Tusch et al., Reference Tusch, Münker, Hasenstab, Jansen, Marien, Kurzweil, Van Kranendonk, Smithies, Maier and Garbe-Schönberg2021, Reference Tusch, Hoffmann, Hasenstab, Fischer-Gödde, Marien, Wilson and Münker2022).

Furthermore, Aulbach and Stagno (Reference Aulbach and Stagno2016) propose that, in contrast to their suite of samples, the rocks measured by previous studies were not derived from the convective mantle. They argue that the latter intruded into a cratonic setting and thus experienced mixing with the sublithospheric mantle. An oxidised mantle at the end of the Archean has also been suggested due to an increase of mantle mixing gradually over time (O’Neill and Aulbach, Reference O’Neill and Aulbach2022), by a change in interior convection patterns from two-layered to one-layered mantle convection (Breuer and Spohn, Reference Breuer and Spohn1995) or by the onset of plate tectonics (Debaille et al., Reference Debaille, O’Neill, Brandon, Haenecour, Yin, Mattielli and Treiman2013; Andrault et al., Reference Andrault, Muñoz, Pesce, Cerantola, Chumakov, Kantor, Pascarelli, Rüffer and Hennet2018). The main argument for mantle mixing due to plate tectonics is to allow the more oxidised, bridgmanite-rich lower mantle (Mao and Bell, Reference Mao, Bell, Saxena, Bhattacharji, Annersten and Stephansson1977; Frost et al., Reference Frost, Liebske, Langenhorst, McCammon, Trønnes and Rubie2004, Reference Frost, Mann, Asahara and Rubie2008; Wade and Wood, Reference Wade and Wood2005) to efficiently mix with the more reducing upper mantle material due to slabs penetrating and stirring up the lower mantle. The mechanisms resulting in mantle mixing could explain the observed rise in upper mantle ƒO2 between 3.0 and 2.0 Ga (Figs 3, 4a; Aulbach and Viljoen, Reference Aulbach and Viljoen2015; Aulbach and Stagno, Reference Aulbach and Stagno2016; Aulbach et al., Reference Aulbach, Woodland, Vasilyev, Galvez and Viljoen2017; Stagno and Fei, Reference Stagno and Fei2020; O’Neill and Aulbach, Reference O’Neill and Aulbach2022).

First hints of locally oxidised surface environments around 3.0 Ga: implications from stable isotopes

The first geochemical evidence for locally oxidised conditions in marginal marine basins comes from measurements of stable isotopes (e.g. chromium, molybdenum, uranium) of marine black shales (e.g. Anbar et al., Reference Anbar2007; Scott et al., Reference Scott, Lyons, Bekker, Shen, Poulton, Chu and Anbar2008; Lyons et al., Reference Lyons, Reinhard and Planavsky2014; Planavsky et al., Reference Planavsky2014; Kendall et al., Reference Kendall, Creaser, Reinhard, Lyons and Anbar2015; Ossa Ossa et al., Reference Ossa Ossa, Hofmann, Vidal, Kramers, Belyanin and Cavalazzi2016, Reference Ossa Ossa, Hofmann, Wille, Spangenberg, Bekker, Poulton, Eickmann and Schoenberg2018; Wang et al., Reference Wang2018, Reference Wang, Ossa Ossa, Hofmann, Agangi, Paprika and Planavsky2020; Brüske et al., Reference Brüske2020; Kendall, Reference Kendall2021). Earth’s mantle and crustal rocks contain chromium in a trivalent state. In modern surface environments, Cr(III) is oxidised to soluble Cr(VI), which is preferentially enriched in heavy isotopes (δ53Cr > 0). The oxidation occurs by the reaction of Cr(III) with Mn(IV) oxides, which require free O2 exceeding 0.1–1% of the present atmospheric level (PAL) (Planavsky et al., Reference Planavsky2014); therefore heavy chromium isotopes are a proxy for the presence of free O2 in the surface environment. Molybdenum isotopes are another tracer for the presence of free O2. Molybdenum adsorbs on Mn(IV) oxide surfaces, a reaction with strong mass-dependent fractionation toward lighter isotopes, therefore low δ98/95Mo values hint towards the existence of Mn(IV) oxides, which require free O2 to form.

No chromium isotope fractionation has been observed in 3.8 Ga banded iron formations (BIFs) from Isua (Frei et al., Reference Frei, Gaucher, Poulton and Canfield2009), which is taken as evidence for atmospheric O2 pressures below 0.02–0.2 bar (i.e. 0.1–1% PAL) (Fig. 4b). This can be regarded as an indication that oxygenation of the surface reservoir had not yet initiated at 3.8 Ga. The earliest hints of locally oxidised conditions are currently recorded in the 3.0 Ga Sinqeni Formation of the Mozaan Group in South Africa (Planavsky et al., Reference Planavsky2014; Ossa Ossa et al., Reference Ossa Ossa, Hofmann, Vidal, Kramers, Belyanin and Cavalazzi2016, Reference Ossa Ossa, Hofmann, Wille, Spangenberg, Bekker, Poulton, Eickmann and Schoenberg2018; Smith and Beukes, Reference Smith and Beukes2023) (Fig. 4c). Contemporaneous oxidative weathering in soils was suggested based on the extensive mobilisation of redox-sensitive elements and fractionation of the redox-sensitive δ53Cr value. Crowe et al. (Reference Crowe, Døssing, Beukes, Bau, Kruger, Frei and Canfield2013) reported marked negative δ53Cr from the 3.0 Ga Nsuze paleosol and small positive δ53Cr from contemporaneous Ijzermyn iron formation (both from the Pongola Supergroup, South Africa). They concluded that free O2 exceeding 0.1% PAL existed in the Mesoarchean, some 600 m.y. before the GOE. However, modern weathering was identified at this site and may have altered the chromium isotope ratios (Albut et al., Reference Albut, Babechuk, Kleinhanns, Benger, Beukes, Steinhilber, Smith, Kruger and Schoenberg2018, Reference Albut, Kamber, Brüske, Beukes, Smith and Schoenberg2019). Post-depositional alteration as the cause for the measured chromium isotope fractionation was supported by Heard et al. (Reference Heard, Aarons, Hofmann, He, Ireland, Bekker, Qin and Dauphas2021). They could not confirm the fractionation of chromium isotopes in the Pongola Supergroup paleosol and concluded that the Mesoarchean was anoxic. Irrespective of these arguments, Smith and Beukes (Reference Smith and Beukes2023) combined evidence from detailed stratigraphy, mineralogy, petrography and carbonate mineral chemistry with isotopic evidence from δ13C to conclude that the local surface ocean within this basin was oxidised supporting previous δ56Fe and δ98Mo data. They suggest microaerophilic chemolithoautotrophs were responsible for iron and manganese oxidation, which would require the presence of free oxygen in the water column, but not the atmosphere. Thus, at least concerning O2, the chemical exchange between the hydrosphere and atmosphere can be suppressed. In the following, evidence for a persistently anoxic Archean atmosphere is summarised.

Atmospheric O2 content remains low between 3.25 and 2.75 Ga: implications from mineral archives

In addition to stable isotopes, indirect proxies such as certain mineral deposits can be used as oxygen barometers. Many minerals that are stable in the subsurface environment become oxidised when exposed to the O2-rich modern atmosphere. Notable among these minerals are sulfides like pyrite (FeS2), uraninite (UO2]) or siderite (FeCO3). In the presence of O2, pyrite is oxidised to Fe(III) (oxyhydr)oxides (rust), uraninite to soluble hexavalent species and siderite to Fe(III) (oxyhydr)oxides. Fluvial uraninite and pyrite detritus were described, e.g. by Ramdohr (Reference Ramdohr1958) and Schidlowski (Reference Schidlowski1981), in Archean sedimentary rocks from the Witwatersrand basin (South Africa). The rounded shape of the mineral grains and absence of oxidation rims suggest that they once occurred as river sand in an O2-free Archean environment. Detrital pyrite, gersdorffite [NiAsS], uraninite and siderite were described by Rasmussen and Buick (Reference Rasmussen and Buick1999) from Archean (3.25–2.75 Ga) fluvial sediments from Pilbara (Australia) and later by Hofmann et al., (Reference Hofmann, Bekker, Rouxel, Rumble and Master2009) from South Africa (3.2–2.7 Ga; Fig. 4c). These minerals can be used as oxygen barometers. For instance, the stability of uraninite in the surface environment is limited to atmospheric O2 levels below 10-2 times the PAL (Grandstaff, Reference Grandstaff1980). Detailed thermodynamic modelling resulted in an upper p(O2) limit of 3.2 × 10-5 bar (1.4 × 10-4 times the PAL) (Johnson et al., Reference Johnson, Gerpheide, Lamb and Fischer2014). The presence of detrital siderite puts an upper limit not only on free O2 but also on H2S. Abundant H2S would lead to the pyritisation of siderite, which is not observed in the Archean sediments studied by Rasmussen and Buick (Reference Rasmussen and Buick1999). They concluded that the Archean atmosphere was poor in H2S, with levels below 10–5 bar.

Hexavalent sulfur S(VI), as present in sulfate (SO42–), should not exist in the reduced Archean environment. Instead, S(IV), S(0), or S(–II) should be the dominating sulfur oxidation states in equilibrium with the lithosphere and atmosphere. However, sedimentary and hydrothermal barite (BaSO4) exists in Paleo- and Mesoarchean rocks from Australia (e.g. Dresser Formation) and South Africa (Barberton Greenstone Belt) (Heinrichs and Reimer, Reference Heinrichs and Reimer1977; Thorpe, Reference Thorpe1979; Walter et al., Reference Walter, Buick and Dunlop1980; Lowe et al., Reference Lowe, Drabon and Byerly2019) (Fig. 4c). The presence of oxidised sulfate within at least some surface waters is regarded as disequilibrium sulfate, i.e. it is produced by local processes but is not in thermodynamic equilibrium with the entire reduced environment (Olson et al., Reference Olson, Drabon and Johnston2022). Thus, the presence of sulfate minerals (such as barite) in the geological record is generally not regarded as representative of the redox state of the Archean ocean (Huston and Logan, Reference Huston and Logan2004).

One process to obtain the S(VI) to form barite is the UV-induced photodissociation and disproportionation of SO2 from volcanic degassing into reduced elemental sulfur S(0) and oxidised sulfate S(VI). Indeed, the Paleo- and Mesoarchean sulfate comprises sulfur isotope signatures revealing at least a partial origin from the atmosphere (Bao et al., Reference Bao, Rumble and Lowe2007; Ueno et al., Reference Ueno, Ono, Rumble and Maruyama2008). Triple oxygen isotope data reveal at least two distinct sources of oxygen in sulfate. Apart from an atmospheric endmember, photooxidation of dissolved Fe2+ to Fe3+ could have acted as a sulfur oxidiser and microbial sulfur cycling may also have been significant (Olson et al., Reference Olson, Drabon and Johnston2022). Further suggestions include sulfate formation through the reaction between reduced S(IV, 0, –II) components and water at high temperatures of igneous systems and the disproportionation of SO2 in hydrothermal systems (Halevy, Reference Halevy2013). Thus, sulfate does not require an oxidised environment but may result from very particular reactions involving SO2 from volcanic emissions. All these processes form sulfate, which is in thermodynamic disequilibrium with the atmo-, hydro- and lithosphere and hence contains little information about the redox state of the Archean Earth.

Large-scale oxidation begins around 2.7 Ga: insights from iron formations

Iron formations (IFs) are iron- and silica-rich marine chemical sediments that commonly display a distinct banding (i.e. banded iron formations, BIFs) (e.g. Bekker et al., Reference Bekker, Slack, Planavsky, Krapez, Hofmann, Konhauser and Rouxel2010; Konhauser et al., Reference Konhauser2017; Mänd et al., Reference Mänd, Robbins, Planavsky, Bekke and Konhauser2021; Aftabi et al., Reference Aftabi, Atapour, Mohseni and Babaki2021; Dreher et al., Reference Dreher, Schad, Robbins, Konhauser, Kappler and Joshi2021). Two main endmember types are distinguished. Algoma-type IFs are generally associated with volcanic provinces and comprise large positive europium anomalies inherited from anoxic vent fluids. These comparably small-scale deposits appear throughout the Archean and early Proterozoic (Barrett et al., Reference Barrett, Fralick and Jarvis1988; Bolhar et al., Reference Bolhar, Van Kranendonk and Kamber2005; Ohmoto et al., Reference Ohmoto, Watanabe, Yamaguchi, Naraoka, Haruna, Kakegawa, Hayashi and Kato2006b; Bekker et al., Reference Bekker, Slack, Planavsky, Krapez, Hofmann, Konhauser and Rouxel2010; Pirajno and Yu, Reference Pirajno and Yu2021). Superior-type IFs form on continental shelves covering extensive areas between 2.7 and 1.8 Ga (Fig. 4c), with a few occurrences already around 3.0 Ga (Huston and Logan, Reference Huston and Logan2004; Smith and Beukes, Reference Smith and Beukes2023). Especially after 2.4 Ga, some of these formed above the storm wave base, destroying the banding and generating granular iron formations (GIFs). The depositional depth seems to be related to the depth of the photic zone (Herwartz and Viehmann, Reference Herwartz and Viehmann2024). Superior-type IFs exhibit smaller europium anomalies, pointing to dominant contributions of rare earth elements derived from continental weathering or low-temperature alteration of oceanic crust rather than hydrothermal vents. The direct precipitation from open seawater makes superior-type IFs the prime target for reconstructing ambient seawater conditions (Bekker et al., Reference Bekker, Slack, Planavsky, Krapez, Hofmann, Konhauser and Rouxel2010; Konhauser et al., Reference Konhauser2017; Mänd et al., Reference Mänd, Robbins, Planavsky, Bekke and Konhauser2021). Today, IFs comprise iron-rich phases, including hematite, magnetite, siderite and iron silicates with variable redox states (mean oxidation state of ~Fe2.4+) and low (<<0.5 wt.%) organic carbon content (Klein and Beukes, Reference Klein and Beukes1992; Trendall, Reference Trendall, Altermann and Corcoran2002). However, the mineralogy observed today does not represent the primary precipitates from an ancient ocean (Konhauser et al., Reference Konhauser2017; Muhling and Rasmussen, Reference Muhling and Rasmussen2020). Most candidates for primary precipitates comprise Fe(III) (but see Muhling and Rasmussen, Reference Muhling and Rasmussen2020). Hence, large-scale oxidation of soluble Fe(II) to insoluble Fe(III) is required to form IFs. Several abiotic and biotic mechanisms have been suggested, most of which are proposed to occur within the photic zone of ocean water.

In the absence of an ozone layer, UV irradiation reaches the Earth’s surface, which induces photochemical oxidation of dissolved Fe2+ to Fe3+ (Cairns-Smith, Reference Cairns-Smith1978; Braterman et al., Reference Braterman, Cairns-Smith and Sloper1983; Anbar and Holland, Reference Anbar and Holland1992). It is suggested that this process occurs at a sufficient rate to form IF deposits (François, Reference François1986). In contrast, Konhauser et al. (Reference Konhauser, Amskold, Lalonde, Posth, Kappler and Anbar2007a) argue that the photochemical contribution to solid-phase precipitation is negligible, as most of the Fe2+ quickly forms poorly crystalline precursor phases to Fe(II) silicates and/or Fe(II) carbonates. The rate of indirect photochemical oxidation via atmospheric H2O2 is found to be too low to account for depositional rates of IF (Pecoits et al., Reference Pecoits, Smith, Catling, Philippot, Kappler and Konhauser2015). Another source of H2O2 is the decay of primordial radioactive isotopes dissolved in seawater. Ershov (Reference Ershov2021) estimates that the decay of highly soluble 40K alone may account for the oxidation of 1021 g of iron within a period between 4.3 and 2.5 Ga. The aqueous oxidation of Fe2+ to Fe3+ is favourable at high pH because this reaction generates protons (2Fe2+ + 4H2O → 2FeOOH + H2 + 4H+). Shibuya et al. (Reference Shibuya, Komiya, Nakamura, Takai and Maruyama2010) argue that high-temperature hydrothermal vent fluids, which are acidic today, had elevated pH in the Archean and comprised Fe3+. Experimental results by Dodd et al. (Reference Dodd2022) show that the decomposition of Fe(OH)2 in Archean seawater analogues produces Fe3+ species. The Fe(OH)2 compound is stable at elevated pH.

The spontaneous conversion of green rust (Fe42+Fe23+(OH)12SO4•8H2O) to magnetite (Fe2+Fe3+2O4) goes along with a net increase in Fe3+ (Tamaura et al., Reference Tamaura, Yoshida and Katsura1984; Li et al., Reference Li, Konhauser and Zhai2017). Green rust is commonly considered a primary iron precipitate in Archean oceans (e.g. Sun et al., Reference Sun, Lechte, Shi, Zhou, Zhou, Feng, Xie, Wu and Tang2022) which are, however, considered to be sulfate-poor, at least between 3.2 and 2.4 Ga (Huston and Logan, Reference Huston and Logan2004). Archean seawater chemistry (including pH and ion concentrations) considerably affects the efficiency of abiotic iron oxidation pathways (e.g. Konhauser et al., Reference Konhauser, Amskold, Lalonde, Posth, Kappler and Anbar2007a; Shibuya et al., Reference Shibuya, Komiya, Nakamura, Takai and Maruyama2010). Therefore, the respective net contribution to individual BIF deposits remains unclear and may vary spatially and over time for each abiotic oxidation mechanism.

The proposed biotic iron oxidation mechanisms can be subdivided into indirect oxidation by free O2 from oxygenic photosynthesis (Cloud, Reference Cloud1973; Klein and Beukes, Reference Klein and Beukes1992) and direct oxidation either by chemolithoautotrophic or anoxygenic photoautotrophic iron-oxidising bacteria (Konhauser et al., Reference Konhauser, Hamade, Raiswell, Morris, Ferris, Southam and Canfield2002; Kappler and Newman, Reference Kappler and Newman2004; Kappler et al., Reference Kappler, Pasquero, Konhauser and Newman2005). The relative proportions of these pathways can be approximated from reactive transport modelling (Ozaki et al., Reference Ozaki, Thompson, Simister, Crowe and Reinhard2019). This approach shows how variations between individual settings with variable nutrient and Fe2+ supply and the available light intensity within a given water mass control the dominating oxidation pathway (Ozaki et al., Reference Ozaki, Thompson, Simister, Crowe and Reinhard2019; Herwartz and Viehmann, Reference Herwartz and Viehmann2024).

Manganese in IFs is a main tracer for the oxygenation of Earth’s hydrosphere (Robbins et al., Reference Robbins2023). Tetravalent Mn(IV) oxides form at redoxclines via consumption of dissolved molecular O2 and are thus direct evidence for oxygenic photosynthesis (see Robbins et al., Reference Robbins2023 for a review). Oxidised Mn4+, Fe3+ and organic matter form particles that sink towards the seafloor. This process is observed in anoxic basins today and is known as the Fe–Mn shuttle (Dellwig et al., Reference Dellwig, Leipe, März, Glockzin, Pollehne, Schnetger, Yakushev, Böttcher and Brumsack2010). While respective particles slowly sink below the chemocline into the anoxic water body, the Mn4+ is reduced again by dissolved Fe2+ (Dellwig et al., Reference Dellwig, Leipe, März, Glockzin, Pollehne, Schnetger, Yakushev, Böttcher and Brumsack2010; Kurzweil et al., Reference Kurzweil, Wille, Gantert, Beukes and Schoenberg2016; Ossa Ossa et al., Reference Ossa Ossa, Hofmann, Wille, Spangenberg, Bekker, Poulton, Eickmann and Schoenberg2018). Deposition of such Mn4+ particles in the sediment is only viable at low Fe2+, e.g. distal to the iron source (Smith and Beukes, Reference Smith and Beukes2023), or when oxygenic photosynthesis is so active that the flux of sinking Mn4+ particles outcompetes the upwelling flux of Fe2+. During and in the aftermath of the GOE, enormous amounts of Mn4+ particles have been deposited on the seafloor, forming the world’s largest manganese deposits (Gutzmer and Beukes, Reference Gutzmer and Beukes1996; Tsikos et al., Reference Tsikos, Beukes, Moore and Harris2003; Sekine et al., Reference Sekine2011), reflecting the high productivity around that time. Elevated manganese contents are a prime indicator for “whiffs of oxygen” (Planavsky et al., Reference Planavsky2014; Ossa Ossa et al., Reference Ossa Ossa, Hofmann, Vidal, Kramers, Belyanin and Cavalazzi2016; Smith and Beukes, Reference Smith and Beukes2023) and a general increase in manganese contents in IFs is observed at the onset of the GOE, e.g. in the Transvaal Supergroup of South Africa (Tsikos et al., Reference Tsikos, Beukes, Moore and Harris2003; Schröder et al., Reference Schröder, Bedorf, Beukes and Gutzmer2011; Kurzweil et al., Reference Kurzweil, Wille, Gantert, Beukes and Schoenberg2016; Smith, Reference Smith, Siegesmund, Basei, Oyhantçabal and Oriolo2018).

Subsequent oxidation of organic matter in the sediment partially reduced Fe3+ and Mn4+ back to soluble Fe2+ and Mn2+. Hence, diagenetic processes can be responsible for the variable mineralogy observed in IFs today. For instance, the Fe and Mn in siderite and rhodochrosite can be derived from the oxidation of organic matter by Fe3+ and Mn4+, which precludes the use of IF mineralogy to reconstruct paleo-atmospheric gas concentrations (Reinhard and Planavsky, Reference Reinhard and Planavsky2011). Identifying primary mineral phases and other features, such as the banding of BIFs, has been the main challenge in using these rocks as reliable archives (Mänd et al., Reference Mänd, Robbins, Planavsky, Bekke and Konhauser2021; Mundl-Petermeier et al., Reference Mundl-Petermeier, Viehmann, Tusch, Bau, Kurzweil and Münker2022; Bau et al., Reference Bau, Frei, Garbe-Schönberg and Viehmann2022).

Abundant whiffs of oxygen between 2.6 and 2.5 Ga: implications from stable isotopes and black shales

Whiffs of oxygen in marine sediments become more abundant in the Neoarchean (2.6–2.5 Ga) towards the GOE (Anbar et al., Reference Anbar2007; Scott et al., Reference Scott, Lyons, Bekker, Shen, Poulton, Chu and Anbar2008; Lyons et al., Reference Lyons, Reinhard and Planavsky2014; Kendall et al., Reference Kendall, Creaser, Reinhard, Lyons and Anbar2015; Ostrander et al., Reference Ostrander, Nielsen, Owens, Kendall, Gordon, Romaniello and Anbar2019; Brüske et al., Reference Brüske2020) (Fig. 4c). Frei et al. (Reference Frei, Gaucher, Poulton and Canfield2009) reported on sedimentary rocks with marked positive δ53Cr, suggesting that O2 rich oases existed before the GOE. These oases probably occurred near the shore, and rivers washed heavy Cr(VI) into the oceans, where chemical sediments preserved the isotope signature.

Significant volumes of black shales started forming at 2.7 Ga (Fig. 4c), indicating a substantial burial of organic carbon that was probably a response to increasing primary productivity via oxygenic photosynthesis (Condie, Reference Condie2001; Lyons et al., Reference Lyons, Reinhard and Planavsky2014). Oxygenic photosynthesis is assumed to be one of the primary mechanisms leading to the significant accumulation of O2 in the oceans and the atmosphere. Additionally, O2 accumulation is favoured by organic carbon burial (Lee et al., Reference Lee, Yeung, McKenzie, Yokoyama, Ozaki and Lenardic2016). Thus, black shales indirectly record the enhanced oxidation of the hydrosphere-atmosphere system.

Significant enrichment of free O2 in the surface reservoir recorded by sediments and sulfur isotopes starting at 2.5 Ga

Widespread release of O2 from an oxygenated surface ocean is evident after around 2.3 Ga when red beds emerge (Konhauser et al., Reference Konhauser2017) (Fig. 4c). Iron loss is observed during paleosol weathering (Rye and Holland, Reference Rye and Holland1998) and detrital pyrite and uraninite disappearance, just like the MIF-S (mass-independent sulfur isotope fractionation) anomaly disappears with the rise of O2 in the atmosphere (Bekker, Reference Bekker2001). The MIF-S signal is assumed to occur due to photolysis and/or photoexcitation of volcanogenic SO2 by ultraviolet light in a reduced, anoxic atmosphere. Reduced sulfur species displaying a positive ∆33S signature result from this photolytic process (e.g. Farquhar, Reference Farquhar2000; Farquhar et al., Reference Farquhar, Savarino, Airieau and Thiemens2001; Ono et al., Reference Ono, Eigenbrode, Pavlov, Kharecha, Rumble, Kasting and Freeman2003; Whitehill and Ono, Reference Whitehill and Ono2012; Whitehill et al., Reference Whitehill, Xie, Hu, Xie, Guo and Ono2013). The sudden disappearance of the MIF-S signal at ca. 2.3 Ga is one of the most well-known and solid evidence for the accumulation of free O2 above 10–5 PAL and thus marks the onset of the GOE (Fig. 4c) (e.g. Pavlov and Kasting, Reference Pavlov and Kasting2002; Ono et al., Reference Ono, Eigenbrode, Pavlov, Kharecha, Rumble, Kasting and Freeman2003; Bekker et al., Reference Bekker, Holland, Wang, Rumble, Stein, Hannah, Coetzee and Beukes2004; Zahnle et al., Reference Zahnle, Claire and Catling2006; Domagal-Goldman et al., Reference Domagal-Goldman, Kasting, Johnston and Farquhar2008; Guo et al., Reference Guo2009; Luo et al., Reference Luo, Ono, Beukes, Wang, Xie and Summons2016; Warke et al., Reference Warke, Di Rocco, Zerkle, Lepland, Prave, Martin, Ueno, Condon and Claire2020; Poulton et al., Reference Poulton, Bekker, Cumming, Zerkle, Canfield and Johnston2021). The positive ∆33S signal in Archaean sedimentary sulfides is recognisable but comparatively low between 3.9 and 2.7 Ga (except for a peak between ca. 3.2 and 3.2 Ga). Between about 2.7 and 2.5 Ga a pronounced MIF-S spike is observed coinciding with the GOE (Farquhar, Reference Farquhar2000; Mojzsis et al., Reference Mojzsis, Coath, Greenwood, McKeegan and Harrison2003; Ono et al., Reference Ono, Eigenbrode, Pavlov, Kharecha, Rumble, Kasting and Freeman2003; Hu et al., Reference Hu, Rumble and Wang2003; Bekker et al., Reference Bekker, Holland, Wang, Rumble, Stein, Hannah, Coetzee and Beukes2004; Whitehouse et al., Reference Whitehouse, Kamber, Fedo and Lepland2005; Papineau et al., Reference Papineau, Mojzsis, Coath, Karhu and McKeegan2005; Jamieson et al., Reference Jamieson, Wing, Hannington and Farquhar2006; Ohmoto et al., Reference Ohmoto, Watanabe, Ikemi, Poulson and Taylor2006a; Ono et al., Reference Ono, Beukes, Rumble and Fogel2006; Cates and Mojzsis, Reference Cates and Mojzsis2006; Johnston et al., Reference Johnston, Poulton, Fralick, Wing, Canfield and Farquhar2006; Kamber and Whitehouse, Reference Kamber and Whitehouse2007; Papineau et al., Reference Papineau, Mojzsis and Schmitt2007; Philippot et al., Reference Philippot, Van Zuilen, Lepot, Thomazo, Farquhar and Van Kranendonk2007; Kaufman et al., Reference Kaufman, Johnston, Farquhar, Masterson, Lyons, Bates, Anbar, Arnold, Garvin and Buick2007; Bao et al., Reference Bao, Rumble and Lowe2007; Farquhar et al., Reference Farquhar, Johnston and Wing2007; Domagal-Goldman et al., Reference Domagal-Goldman, Kasting, Johnston and Farquhar2008; Partridge et al., Reference Partridge, Golding, Baublys and Young2008; Johnston et al., Reference Johnston, Farquhar, Summons, Shen, Kaufman, Masterson and Canfield2008; Ueno et al., Reference Ueno, Ono, Rumble and Maruyama2008; Ono et al., Reference Ono, Beukes and Rumble2009; Thomazo et al., Reference Thomazo, Ader, Farquhar and Philippot2009a; Shen et al., Reference Shen, Farquhar, Masterson, Kaufman and Buick2009; Guo et al., Reference Guo2009; Gaillard et al., Reference Gaillard, Scaillet and Arndt2011; Lyons et al., Reference Lyons, Reinhard and Planavsky2014; Ono, Reference Ono2017; Kendall, Reference Kendall2021).

While the atmosphere, surface oceans and marginal basins are oxidised after the GOE (Lyons et al., Reference Lyons, Reinhard and Planavsky2014), the deep ocean remains anoxic until 1.8 Ga (Huston and Logan, Reference Huston and Logan2004). Deep ocean oxygenation requires Phanerozoic-like atmospheric O2 levels and deep ocean convection (Reinhard and Planavsky, Reference Reinhard and Planavsky2022). Therefore, fully oxidised oceans as we know them today do not appear until atmospheric O2 levels approach modern levels in the late Proterozoic oxidation event (Reinhard and Planavsky, Reference Reinhard and Planavsky2022; but see Xu et al., Reference Xu, Qin, Wang, Li, Shi, Tang and Liu2023), and even then bottom-water anoxia seem to have been the rule rather than the exception until the mid-Paleozoic era (Stockey et al. Reference Stockey, Cole, Farrell, Agić, Boag, Brocks, Canfield, Cheng, Crockford, Cui, Dahl, Del Mouro, Dewing, Dornbos, Emmings, Gaines, Gibson, Gill, Gilleaudeau and Sperling2024).

Figure 4. Oxidation and oxygenation of the upper mantle and the surface environment over time. (a) Calculated ƒO2 of samples derived from the upper mantle after Aulbach and Stagno (Reference Aulbach and Stagno2016) and Stagno and Aulbach (Reference Stagno, Aulbach, Moretti and Neuville2021) (see Fig. 3 for the legend). The rectangles at the top of the figure display some important geodynamic events: extensive formation of continental crust ca. 3.5–2.4 Ga (Collerson and Kamber, Reference Collerson and Kamber1999; Huston and Logan, Reference Huston and Logan2004), onset of modern style plate tectonics (PT) ca. 3.2–3.0 Ga (Smithies et al., Reference Smithies, Champion, Van Kranendonk, Howard and Hickman2005; Van Kranendonk et al., Reference Van Kranendonk, Smithies, Hickman and Champion2007; Van Kranendonk, Reference Van Kranendonk2011; Duncan and Dasgupta, Reference Duncan and Dasgupta2017; Kuang et al., Reference Kuang, Morra, Yuen, Kusky, Jiang, Yao and Qi2023), major crustal growth ca. 2.7 Ga (Gaillard et al., Reference Gaillard, Scaillet and Arndt2011), intrusion of large igneous provinces (LIPs) 2.5–2.4 Ga (Ernst and Bleeker, Reference Ernst and Bleeker2010; Gumsley et al., Reference Gumsley, Chamberlain, Bleeker, Söderlund, de Kock, Larsson and Bekker2017), first glaciations (Ice) ca. 2.4 Ga (Kirschvink et al., Reference Kirschvink, Gaidos, Bertani, Beukes, Gutzmer, Maepa and Steinberger2000; Gumsley et al., Reference Gumsley, Chamberlain, Bleeker, Söderlund, de Kock, Larsson and Bekker2017). (b) Partial pressure of specific gas species over time after Catling and Zahnle (Reference Catling and Zahnle2020). The two grey dashed lines indicate the partial pressure of CO2. The upper line (K) is after Kasting (Reference Kasting1987) and Herwartz et al. (Reference Herwartz, Pack and Nagel2021), and the lower line (CZ) is after Catling and Zahnle (Reference Catling and Zahnle2020). (c) Geological evidence for oxygenation of the surface environment is explained in detail in the text. The red star indicates early whiffs of oxygen at 3.0 Ga. WO, more abundant later whiffs of oxygen; BS, black shales; Red, red beds (see the main text for references). Mass-independent sulfur isotope fractionation (MIF-S) marks the positive and negative excursions of Δ33S in ‰ (after Ono, Reference Ono2017; see also the main text). The data for banded iron formations (BIFs) (S, = superior-type; A, algoma-type) and sulfate deposits (barite) is from Huston and Logan (Reference Huston and Logan2004). Reddish colours display evidence of oxygenation, while blueish colours indicate reduced conditions. The boxes without colours are deposits discussed in the literature as possible hints for redox conditions, even though the general opinion is that they cannot be used as redox proxies. (d) Timetable for the emergence of the microbial metabolic processes discussed in this review. Solid lines represent well-established timeframes. Dashed lines represent tentative timeframes. Question marks signify highly uncertain periods. Adapted from Lepot (Reference Lepot2020) and modified based on references in the text. The red bar across the whole figure indicates the timing of the GOE (2.5–2.3 Ga; see text for references). GOE, Great Oxygenation Event; QFM, quartz-fayalite-magnetite

O2 sources vs. sinks: balancing atmospheric oxygenation

Upper mantle oxidation to near modern ƒO2 probably occurred between 3.0 and 2.0 Ga (Fig. 4). The first geochemical evidence for localised O2 appears ca. 3.0 Ga, consistent with paleontological and phylogenetic evidence for the emergence of oxygenic photosynthesis (Anbar et al., Reference Anbar2007; Planavsky et al., Reference Planavsky2014; Schirrmeister et al., Reference Schirrmeister, Gugger and Donoghue2015; Sánchez-Baracaldo, Reference Sánchez-Baracaldo2015; Cardona et al., Reference Cardona, Sánchez-Baracaldo, Rutherford and Larkum2019; Garcia-Pichel et al., Reference Garcia-Pichel, Lombard, Soule, Dunaj, Wu and Wojciechowski2019; Jabłońska and Tawfik, Reference Jabłońska and Tawfik2021, Reference Jabłońska and Tawfik2021; Fournier et al., Reference Fournier, Moore, Rangel, Payette, Momper and Bosak2021; Boden et al., Reference Boden, Zhong, Anderson and Stüeken2024) (Fig. 4d). Despite the constant chemical exchange between atmosphere and hydrosphere, mineralogical and geochemical evidence in the Earth’s sedimentary rock records contrasting timelines for their respective oxygenation. Irrespective of this problem, atmospheric oxygenation only occurs at 2.5–2.3 Ga (e.g. Holland, Reference Holland2002, Reference Holland2006; Bekker et al., Reference Bekker, Holland, Wang, Rumble, Stein, Hannah, Coetzee and Beukes2004; Canfield, Reference Canfield2005; Kasting et al., Reference Kasting, Howard, Wallmann, Veizer, Shields and Jaffrés2006; Guo et al., Reference Guo2009; Luo et al., Reference Luo, Ono, Beukes, Wang, Xie and Summons2016; Gumsley et al., Reference Gumsley, Chamberlain, Bleeker, Söderlund, de Kock, Larsson and Bekker2017; Warke et al., Reference Warke, Di Rocco, Zerkle, Lepland, Prave, Martin, Ueno, Condon and Claire2020; Ossa Ossa et al., Reference Ossa Ossa2022). This indicates a delay between upper mantle oxidation, the emergence of oxygenic photosynthesis and the GOE by several hundred m.y. (Fig. 4). Thus, the onset of microbial O2 production alone cannot satisfactorily explain the timing of the GOE.

Constraining the O2 source-limited oxygenic photosynthesis?

One explanation for this delayed oxygenation of the Earth’s surface environments is that the productivity of oxygenic photosynthesis in early cyanobacteria was limited, decreasing the microbial O2 flux (see Dick et al., Reference Dick, Grim and Klatt2018 for a detailed review). Cyanobacteria depend on bioavailable nitrogen and are major agents for nitrogen fixation in today’s surface oceans (Field et al., Reference Field, Behrenfeld, Randerson and Falkowski1998; Zehr and Kudela, Reference Zehr and Kudela2011). Nitrogen fixation is catalysed by the enzyme nitrogenase, which contains molybdenum (Postgate, Reference Postgate1998). Molybdenum may have been scarce in the reducing environments of the early Archean Earth, where it was poorly soluble (Williams and Fraústo Da Silva, Reference Williams and Fraústo Da Silva2003). Thus, it was suggested that nitrogen fixation in cyanobacteria was inhibited (Zerkle et al., Reference Zerkle, House, Cox and Canfield2006). At the same time, O2 output by oxygenic photosynthesis could have inhibited other nitrogen-fixing microorganisms, ultimately starving cyanobacteria of bioavailable nitrogen (Shi and Falkowski, Reference Shi and Falkowski2008; Kasting and Canfield, Reference Kasting, Canfield, Knoll, Canfield and Konhauser2012). Nitrogen fixation could have occurred via lightning-driven atmospheric reactions (Navarro-González et al., Reference Navarro-González, Molina and Molina1998; Wong et al., Reference Wong, Charnay, Gao, Yung and Russell2017). Still, the isotopic composition of most Archean sedimentary nitrogen isotope records suggests this process was not quantitatively important for sustaining primary production (Barth et al., Reference Barth, Stüeken, Helling, Rossmanith, Peng, Walters and Claire2023). Instead, nitrogen isotope evidence is consistent with biological nitrogen fixation by at least 3.2 Ga (Stüeken et al., Reference Stüeken, Buick, Guy and Koehler2015a, Reference Stüeken, Kipp, Koehler and Buick2016). This is consistent with phylogenetic studies suggesting an early emergence of nitrogen fixation in cyanobacteria (Latysheva et al., Reference Latysheva, Junker, Palmer, Codd and Barker2012). Hydrothermal sources may have sufficiently compensated the low supply of molybdenum for nitrogenase from oxidative weathering in the Archean (Evans et al., Reference Evans, Coogan, Kaçar and Seyfried2023). Moreover, hydrothermal systems probably played a role in recycling sedimentary ammonium (Stüeken et al., Reference Stüeken, Boocock, Robinson, Mikhail and Johnson2021; Martin et al., Reference Martin, Stüeken, Michaud, Münker, Weyer, Van Hees and Gehringer2024). Therefore, nitrogen may not have been a limiting factor for cyanobacterial productivity in the late Archean.

Bioavailable phosphorous, in the form of phosphate, is widely considered another limiting factor for primary productivity in the Archean and early Proterozoic ocean (Derry, Reference Derry2015; Reinhard et al., Reference Reinhard, Planavsky, Gill, Ozaki, Robbins, Lyons, Fischer, Wang, Cole and Konhauser2017; Ossa Ossa et al., Reference Ossa Ossa, Hofmann, Spangenberg, Poulton, Stüeken, Schoenberg, Eickmann, Wille, Butler and Bekker2019; Walton et al., Reference Walton, Ewens, Coates, Blake, Planavsky, Reinhard, Ju, Hao and Pasek2023). This is despite its supply from continental weathering (Hao et al., Reference Hao, Knoll, Huang, Hazen and Daniel2020; Watanabe and Tajika, Reference Watanabe and Tajika2021) and possible hydrothermal sources (Rasmussen et al., Reference Rasmussen, Muhling, Suvorova and Fischer2021, Reference Rasmussen, Muhling, Tosca and Fischer2023). For example, Ozaki et al. (Reference Ozaki, Thompson, Simister, Crowe and Reinhard2019) provide a model for open ocean settings and investigate the competition between O2-producing cyanobacteria and photoferrotrophs, the latter being adapted to lower light levels, allowing them to thrive deeper in the water column. Accordingly, nutrients such as phosphate and Fe2+ from upwelling water masses are consumed by photoferrotrophs, leaving surface water starved in either phosphate or Fe2+ (Kappler et al., Reference Kappler, Pasquero, Konhauser and Newman2005; Ozaki et al., Reference Ozaki, Thompson, Simister, Crowe and Reinhard2019). If oceanic iron/phosphate ratios are high, oxygenic photosynthesis in the upper water column is efficiently suppressed (Ozaki et al., Reference Ozaki, Thompson, Simister, Crowe and Reinhard2019). In coastal settings, where the water column is shallow and more nutrients are supplied from the continent, benthic microbial mats are observed that probably produce O2 (Homann et al., Reference Homann, Heubeck, Airo and Tice2015; Homann, Reference Homann2019), therefore the productivity of cyanobacteria probably varied spatially and over time (Konhauser et al., Reference Konhauser2018). Moreover, the bioavailability of phosphorous may have been limited due to inefficient remineralisation of organic matter (Kipp and Stüeken, Reference Kipp and Stüeken2017) or phosphate scavenging by Fe2+ and adsorption on or co-precipitation with (biogenic) Fe(III) minerals in ferruginous oceans (Bjerrum and Canfield, Reference Bjerrum and Canfield2002; Laakso and Schrag, Reference Laakso and Schrag2014; Derry, Reference Derry2015). The efficiency of phosphate scavenging by Fe (III) minerals is debated (Konhauser et al., Reference Konhauser, Lalonde, Amskold and Holland2007b; Jones et al., Reference Jones, Nomosatryo, Crowe, Bjerrum and Canfield2015). Recent analyses of carbonate-associated phosphate in Archean rocks also challenge severe phosphate limitation in coeval waters (Ingalls et al., Reference Ingalls, Grotzinger, Presen, Rasmussen and Fischer2022; Crockford and Halevy, Reference Crockford and Halevy2022). Hence, the possible phosphate limitation of Archean cyanobacteria remains an open question.

Phylogenetic evidence suggests the earliest cyanobacteria were benthic freshwater strains that only diversified into brackish and marine habitats in the late Archean (Blank and Sánchez-Baracaldo, Reference Blank and Sánchez-Baracaldo2010; Schirrmeister et al., Reference Schirrmeister, Sanchez-Baracaldo and Wacey2016; Sánchez-Baracaldo et al., Reference Sánchez-Baracaldo, Raven, Pisani and Knoll2017; Grettenberger et al., Reference Grettenberger, Gold and Brown2025). Planktonic cyanobacteria may have only appeared in the Neoproterozoic, expanding their habitat to the open ocean (Sánchez-Baracaldo et al., Reference Sánchez-Baracaldo, Ridgwell and Raven2014, Reference Sánchez-Baracaldo, Bianchini, Di Cesare, Callieri and Chrismas2019; Sánchez-Baracaldo, Reference Sánchez-Baracaldo2015; Schirrmeister et al., Reference Schirrmeister, Sanchez-Baracaldo and Wacey2016). A benthic lifestyle, on the other hand, would have constrained the spatial extent of Archean cyanobacterial habitats to terrestrial or coastal areas, limiting their overall O2 production (Sánchez-Baracaldo et al., Reference Sánchez-Baracaldo, Ridgwell and Raven2014, Reference Sánchez-Baracaldo, Raven, Pisani and Knoll2017; Lalonde and Konhauser, Reference Lalonde and Konhauser2015; Sánchez-Baracaldo, Reference Sánchez-Baracaldo2015). Once cyanobacteria expanded to marine environments, they may have suffered from iron toxicity in the Archean oceans (Swanner et al., Reference Swanner, Mloszewska, Cirpka, Schoenberg, Konhauser and Kappler2015a, Reference Swanner, Wu, Hao, Wüstner, Obst, Moran, McIlvin, Saito and Kappler2015b; Dreher et al., Reference Dreher, Schad, Robbins, Konhauser, Kappler and Joshi2021). If true, this was most probably due to reactive oxygen species produced during Fe(II) oxidation by photosynthetic O2 (Rush and Bielski, Reference Rush and Bielski1985). However, more recent experiments did not observe such effects in open bottle cultures that allowed for gas exchange, limiting the accumulation of photosynthetic O2 to concentrations assumed for Archean oxygen oases (≤10 μM; Herrmann et al., Reference Herrmann, Sorwat, Byrne, Frankenberg-Dinkel and Gehringer2021). In these sunlit environments, early cyanobacteria would have also been exposed to high levels of UV radiation due to the absence of an ozone shield (Mloszewska et al., Reference Mloszewska, Cole, Planavsky, Kappler, Whitford, Owttrim and Konhauser2018). Recently, it has also been suggested that the net O2 production in Archean cyanobacterial mats was lower than previously thought due to shorter day lengths (Klatt et al., Reference Klatt, Chennu, Arbic, Biddanda and Dick2021) or an inefficient photosystem in early cyanobacteria (Grettenberger and Sumner, Reference Grettenberger and Sumner2024). One or more of these factors could have limited the productivity of cyanobacteria in the late Archean. The increasing abundance of continental crust through the Archean (Fig. 4a) (Kemp and Hawkesworth, Reference Kemp and Hawkesworth2014; Smit and Mezger, Reference Smit and Mezger2017; Korenaga, Reference Korenaga2018) may have helped overcome some of these limitations by supplying weathering-derived phosphate and creating shallow marine habitats, increasing the biological O2 source. However, reliable primary productivity estimates depend on the magnitudes of these effects, which remain to be determined.

Critical O2 sinks: reduced gases and solutes in Archean surface environments

Apart from limited biological O2 production, low ambient O2 levels could also be due to large fluxes into O2 sinks. The most prominent O2 sinks are reduced gases and aqueous solutes in the Archean atmosphere and oceans (e.g. Fe2+, Mn2+, H2S, Corg) (Holland, Reference Holland2002; Claire et al., Reference Claire, Catling and Zahnle2006; Gaillard et al., Reference Gaillard, Scaillet and Arndt2011; Lyons et al., Reference Lyons, Reinhard and Planavsky2014, Reference Lyons, Tino, Fournier, Anderson, Leavitt, Konhauser and Stüeken2024; Lee et al., Reference Lee, Yeung, McKenzie, Yokoyama, Ozaki and Lenardic2016; Catling and Zahnle, Reference Catling and Zahnle2020). The sizes and capacities of these sinks are partly controlled by fluid–rock interactions and volcanic outgassing in chemical equilibrium with the redox state of the Earth’s mantle and crust (Gaillard et al., Reference Gaillard, Bouhifd, Füri, Malavergne, Marrocchi, Noack, Ortenzi, Roskosz and Vulpius2021). As discussed above, Earth supposedly accreted from relatively reduced material. However, the mantle subsequently experienced oxidation between 3.0 and 2.0 Ga (Fig. 4), shifting the redox state of volcanic gases toward more oxidised species (Claire et al., Reference Claire, Catling and Zahnle2006; Aulbach and Stagno, Reference Aulbach and Stagno2016; O’Neill and Aulbach, Reference O’Neill and Aulbach2022). When the abundance of continental crust rose, magmatic outgassing was increasingly shallow and subaerial and, thus, more oxidised due to the pressure dependence of volatile speciation in magmatic systems (Holland, Reference Holland2002; Kump and Barley, Reference Kump and Barley2007; Gaillard et al., Reference Gaillard, Scaillet and Arndt2011). Less mafic and ultramafic rocks at the surface meant fewer reduced solutes (e.g. Fe2+, Mn2+) in seawater (Kump et al., Reference Kump, Kasting and Barley2001; Lee et al., Reference Lee, Yeung, McKenzie, Yokoyama, Ozaki and Lenardic2016) and fewer reduced species from serpentinisation (e.g. H2) (Hoffmann, Reference Hoffmann2017; Smit and Mezger, Reference Smit and Mezger2017). More continental landmass increased the accommodation space for the burial of reduced sediments, removing these critical O2 sinks from the surface environment (Canfield, Reference Canfield2005; Lee et al., Reference Lee, Yeung, McKenzie, Yokoyama, Ozaki and Lenardic2016; Zhao et al., Reference Zhao, Mills, Homoky and Peacock2023).

Biotic processes like microbial methanogenesis also control the abundance of gaseous sinks in surface environments. Methanogens are strictly anaerobic archaea that form CH4 either by reduction of a carbon substrate (e.g. hydrogenotrophic methanogenesis, CO2 reduction with H2 as the electron donor) or by disproportionation (i.e. fermentation) of organic compounds (e.g. acetoclastic methanogenesis, acetate disproportionation) (Head, Reference Head, Ehrlich, Newman and Kappler2016) (Table 1). Methane can contribute to O2 consumption via photochemically generated CH3 and OH radicals in the atmosphere (Pavlov et al., Reference Pavlov, Brown and Kasting2001; Kasting and Siefert, Reference Kasting and Siefert2002; Claire et al., Reference Claire, Catling and Zahnle2006; Daines and Lenton, Reference Daines and Lenton2016). A study on the ferruginous and sulfate-poor Lake Matano, an Archean ocean analogue site, showed a limitation of dissimilatory sulfate reduction in favour of methanogenesis (Crowe et al., Reference Crowe2011). Methanogens could, therefore, also have been important agents for the remineralisation of organic matter from Archean primary production (e.g. Thompson et al., Reference Thompson2019). Highly 13C depleted CH413C down to –56‰) from fluid inclusions in hydrothermal quartz of the 3.48 Ga Dresser Formation may represent the oldest direct evidence for methanogenesis (Ueno et al., Reference Ueno, Yamada, Yoshida, Maruyama and Isozaki2006). However, this CH4 may originate from abiotic organic synthesis (Sherwood Lollar and McCollom, Reference Sherwood Lollar and McCollom2006). Carbon isotopic evidence for methanogenesis is also found in ca. 3.0 Ga fluvio-lacustrine Lalla Rokh Sandstone (δ13Corg –30 to –38‰; Stüeken and Buick, Reference Stüeken and Buick2018), and in the shallow marine or lacustrine 2.72 Ga Tumbiana Formation in Western Australia (δ13Corg down to –56‰; Thomazo et al., Reference Thomazo, Ader and Philippot2011), which are probably unaffected by hydrothermal CH4. Indeed, recent molecular clock studies place the emergence of methanogenesis even ≥3.5 Ga (Wolfe and Fournier, Reference Wolfe and Fournier2018; but see Roger and Susko, Reference Roger and Susko2018 for an alternative viewpoint). Assuming biogenic CH4 fluxes similar to today, the Archean atmosphere could have maintained CH4 concentrations of thousands of ppm (Pavlov et al., Reference Pavlov, Brown and Kasting2001; Kasting and Siefert, Reference Kasting and Siefert2002; Kharecha et al., Reference Kharecha, Kasting and Siefert2005). The CH4 flux probably decreased when sufficient seawater sulfate was available and sulfate-reducing bacteria overtook methanogens in organic carbon mineralisation (Zahnle et al., Reference Zahnle, Claire and Catling2006). Declining oceanic nickel concentrations may have further inhibited methanogens (Konhauser et al., Reference Konhauser, Pecoits, Lalonde, Papineau, Nisbet, Barley, Arndt, Zahnle and Kamber2009). It seems plausible that the late Archean decline of the atmospheric CH4 pool was crucial for the subsequent rise of atmospheric O2, rendering biogenic CH4 an important O2 sink in the Archean. Together with an increasing flux of biogenic O2, these processes exhausted the capacity of Earth’s gaseous and aqueous O2 sinks throughout the Archean, paving the way for the GOE.

The Archean aerobic O2 sink: insights from recent environments, ancient rocks and modern genomes

Another potential O2 sink in the Archean is its reduction coupled to the oxidation of various electron donors (organic matter, CH4, NH4+, NO2, Mn(II), Fe(II), sulfide) for conserving energy (i.e. aerobic respiration) by microorganisms (Table 1). Aerobic respiration is associated with a greater energy yield compared to anaerobic (i.e. O2-free) respiration, making it highly competitive in environments where O2 is available. Indeed, it is the most competitive pathway for organic carbon remineralisation to CO2 on modern Earth and a critical buffer against further atmospheric O2 accumulation (Berner, Reference Berner1989). Microbial oxidation of CH4 to CO2 using O2 as the terminal electron acceptor (i.e. aerobic methanotrophy) is a critical CH4 sink (Table 1). Ammonium oxidation to nitrite and nitrate by O2 (i.e. nitrification; Table 1) is dominantly controlled by aerobic microorganisms like Nitrosomonas and Nitrobacter sp. in modern oceans (Falkowski, Reference Falkowski1997; Stüeken et al., Reference Stüeken, Pellerin, Thomazo, Johnson, Duncanson and Schoepfer2024). Microbial manganese oxidation by O2 (Table 1) is the main mechanism for Mn(IV) oxide production in modern oceans (Tebo et al., Reference Tebo, Johnson, McCarthy and Templeton2005). This is because the microbially mediated oxidation of Mn(II) by O2 is up to five orders of magnitude faster than its abiotic counterpart in seawater-like conditions (Nealson et al., Reference Nealson, Tebo and Rosson1988; Tebo, Reference Tebo1991; Hansel, Reference Hansel and Poole2017; Yu and Leadbetter, Reference Yu and Leadbetter2020). Therefore, Mn(II) oxidation in the presence of O2 is generally mediated by Mn-oxidising microorganisms, even at sub-micromolar O2 concentrations (Tebo et al., Reference Tebo, Bargar, Clement, Dick, Murray, Parker, Verity and Webb2004, Reference Tebo, Johnson, McCarthy and Templeton2005; Schippers et al., Reference Schippers, Neretin, Lavik, Leipe and Pollehne2005; Clement et al., Reference Clement, Luther and Tebo2009; Learman et al., Reference Learman, Voelker, Vazquez-Rodriguez and Hansel2011). Microaerophilic Fe(II)-oxidising bacteria (e.g. Gallionella, Leptothrix, Mariprofundus sp) couple the oxidation of Fe(II) to the reduction of O2 (Table 1). At neutral pH, the oxidation of reduced sulfur compounds (e.g. H2S) to S0 or sulfate can occur at micromolar O2 concentrations using O2 or nitrate as an electron acceptor (e.g. Beggiatoa; Hentschel and Felbeck, Reference Hentschel and Felbeck1993; Jørgensen and Gallardo, Reference Jørgensen and Gallardo1999; Girguis et al., Reference Girguis, Lee, Desaulniers, Childress, Pospesel, Felbeck and Zal2000; Dahl et al., Reference Dahl, Friedrich and Kletzin2008) (Table 1). In acidic environments (e.g. hydrothermal sulfide systems, acid rock drainage sites), microorganisms can exploit the aerobic oxidation of S0 or sulfide minerals, such as pyrite (e.g. Acidothiobacillus; Segerer et al., Reference Segerer, Neuner, Kristjansson and Stetter1986; Dahl et al., Reference Dahl, Friedrich and Kletzin2008). Microbial sulfide oxidation is orders of magnitude faster than abiotic sulfide oxidation in most sedimentary environments, particularly at low O2 concentrations, highlighting its role in consuming O2 in the environment (Luther et al., Reference Luther, Findlay, MacDonald, Owings, Hanson, Beinart and Girguis2011).

Despite the significance and diversity of microbial O2 sinks in recent environments and previous suggestions for the antiquity of aerobic respiration (‘respiration early hypothesis’; Castresana and Saraste, Reference Castresana and Saraste1995), the role of biological O2 consumption in the Archean is poorly explored. The proliferation of aerobic microorganisms is traditionally assumed to postdate the GOE due to the canonical lower limit for aerobic respiration (the ‘Pasteur Point’) of 2.2 μM O2 in seawater at 25°C (Devol, Reference Devol1978). However, the discovery of ‘nanaerobic’ life (Baughn and Malamy, Reference Baughn and Malamy2004), respiring aerobically at nanomolar O2 concentrations, challenges this paradigm. For instance, Escherichia coli, a well-studied model organism, grows aerobically at O2 concentrations as low as 3 nM (Stolper et al., Reference Stolper, Revsbech and Canfield2010). Moreover, it is increasingly recognised that many microorganisms typically considered strict anaerobes can also respire aerobically at low O2 concentrations (Cypionka, Reference Cypionka2000; Lee et al., Reference Lee, Youn, Kang and Lee2019; Berg et al., Reference Berg2019). It was cautioned that aerobic growth rates under nanomolar O2 concentrations are strongly muted, suggesting that anaerobic respiration (e.g. Fe(III) reducers) could outcompete aerobic microorganisms in the Archean (Ducluzeau et al., Reference Ducluzeau, Schoepp-Cothenet, Van Lis, Baymann, Russell and Nitschke2014). However, advances in O2 microsensing and metatranscriptomic analysis demonstrate that aerobic respiration is widespread in recent environments, even under apparent anoxia, where aerobic microorganisms may consume O2 faster than it can accumulate (Berg et al., Reference Berg, Ahmerkamp, Pjevac, Hausmann, Milucka and Kuypers2022 and references therein).

Recent oxygen minimum zones and stratified lakes parallel late Archean environments by their low O2 concentrations within or below the photic zone, shedding light on the possible role of microbial communities and biogeochemical processes before the GOE. Aerobic methanotrophy (Table 1) has been identified as an efficient CH4 sink coupled to cryptic O2 from oxygenic photosynthesis in the photic zone of lakes (Oswald et al., Reference Oswald, Milucka, Brand, Littmann, Wehrli, Kuypers and Schubert2015; Milucka et al., Reference Milucka, Kirf, Lu, Krupke, Lam, Littmann, Kuypers and Schubert2015). These conditions could have been widespread in Archean lakes and oceans, supporting the suggested role of aerobic methanotrophy in buffering atmospheric O2 accumulation (Daines and Lenton, Reference Daines and Lenton2016). Nitrification is efficient in oxygen minimum zones, even at nanomolar O2 concentrations (Kalvelage et al., Reference Kalvelage, Jensen, Contreras, Revsbech, Lam, Günter, LaRoche, Lavik and Kuypers2011, Reference Kalvelage2015; Füssel et al., Reference Füssel, Lam, Lavik, Jensen, Holtappels, Günter and Kuypers2012; Thamdrup et al., Reference Thamdrup, Dalsgaard and Revsbech2012; Beman et al., Reference Beman, Leilei Shih and Popp2013; Bristow et al., Reference Bristow2016). Optimal rates of microaerophilic Fe(II) oxidation by Sideroxydans were observed at 5–20 μM O2 (Maisch et al., Reference Maisch, Lueder, Laufer, Scholze, Kappler and Schmidt2019), but the marine strain Mariprofundus can still grow at sub-micromolar O2 concentrations (Chiu et al., Reference Chiu, Kato, McAllister, Field and Chan2017; McAllister et al., Reference McAllister, Moore, Gartman, Luther, Emerson and Chan2019). Most importantly, however, because the abiotic oxidation of Fe(II) is slow at neutral pH and low O2 (Søgaard et al., Reference Søgaard, Medenwaldt and Abraham-Peskir2000), microaerophilic Fe(II) oxidation outcompetes abiotic oxidation at or below 50 μM O2 (Druschel et al., Reference Druschel, Emerson, Sutka, Suchecki and Luther2008). Aerobic sulfide oxidation also occurs in apparently anoxic environments of modern lakes and oxygen minimum zones when the influx and microbial consumption of O2 are balanced, resulting in a cryptic O2 cycle (Sommer et al., Reference Sommer2017; Callbeck et al., Reference Callbeck2018; Berg et al., Reference Berg2019). The micromolar O2 concentrations inferred for late Archean oxygen oases satisfy even conservative lower limits for aerobic respiration, demonstrating that aerobic respiration was viable in Neoarchean and, perhaps transiently, in Mesoarchean surface waters. Cryptic O2 consumption in recent environments, resulting in O2 concentrations below the detection limits of modern microsensors, suggests the downwelling and downward diffusion of oxygenated surface waters may have even allowed for aerobic metabolism in apparently anoxic deeper Archean settings.

Early studies suggested that aerobic respiration is required to mass-balance the preserved organic carbon in Archean black shales (Towe, Reference Towe1990). It is difficult to verify this by investigating the rock record because isotope fractionation involved in heterotrophy is much less than during the initial autotrophic carbon fixation (Hayes, Reference Hayes2001). The resulting carbon species of aerobic methanotrophy are strongly depleted in 13C and can be bound in carbonates or assimilated in microbial biomass, enabling the reconstruction of CH4 oxidation in the geological record (Hayes, Reference Hayes2001; Eigenbrode and Freeman, Reference Eigenbrode and Freeman2006). The δ13C values in sedimentary carbonaceous matter of down to –60‰ were interpreted as evidence for aerobic methanotrophs in the late Archean (Hayes, Reference Hayes and Schopf1983, Reference Hayes and Bengtson1994; Hayes and Waldbauer, Reference Hayes and Waldbauer2006). Eigenbrode and Freeman (Reference Eigenbrode and Freeman2006) present indirect evidence for aerobic respiration based on δ13Corg analysis of ≤2.7 Ga sedimentary rocks. They conclude that the more consistently 13C-depleted deep versus shallow facies (δ13Corg = –40 to –45‰ and –57 to –28‰, respectively) demonstrate a more prominent role of CH4 cycling in anoxic deep water versus (aerobic) respiration of photosynthetic organic matter in oxic surface water. The presence of 3β-methylhopane biomarkers in the 2.7–2.5 Ga Transvaal Supergroup and Hamersley Group supports this (Brocks et al., Reference Brocks, Buick, Logan and Summons2003; Eigenbrode et al., Reference Eigenbrode, Freeman and Summons2008; Waldbauer et al., Reference Waldbauer, Sherman, Sumner and Summons2009), but the syngeneity of biomarkers in these localities to their host rocks was contested (Brocks, Reference Brocks2011). Highly depleted δ13C signatures in Archean carbonaceous matter could also be explained by anaerobic oxidation of methane (AOM) (Hinrichs, Reference Hinrichs2002; Thomazo et al., Reference Thomazo, Pinti, Busigny, Ader, Hashizume and Philippot2009b; Guy et al., Reference Guy, Ono, Gutzmer, Kaufman, Lin, Fogel and Beukes2012; Stüeken et al., Reference Stüeken, Buick, Anderson, Baross, Planavsky and Lyons2017; Flannery et al., Reference Flannery, Allwood, Summons, Williford, Abbey, Matys and Ferralis2018; Lepot et al., Reference Lepot, Williford, Philippo, Thomazo, Ushikubo, Kitajima, Mostefaoui and Valley2019), which is a major CH4 sink on modern Earth (Knittel and Boetius, Reference Knittel and Boetius2009). The quantitative importance of AOM may have been limited in the Archean due to low marine sulfate levels before the GOE (Catling et al., Reference Catling, Claire and Zahnle2007) but alternative electron acceptors like Fe(III) seem plausible (Knittel and Boetius, Reference Knittel and Boetius2009; Stüeken and Buick, Reference Stüeken and Buick2018).

Nitrification produces 15N-enriched residual ammonium (δ15N up to +35‰) that can be assimilated and recorded in sedimentary organic matter (Mariotti et al., Reference Mariotti, Germon, Hubert, Kaiser, Letolle, Tardieux and Tardieux1981; Casciotti, Reference Casciotti2009; Mandernack et al., Reference Mandernack, Mills, Johnson, Rahn and Kinney2009). Nitrite can be used to oxidise ammonium in the absence of excess O2 (anaerobic ammonium oxidation, ‘anammox’; Mulder et al., Reference Mulder, Graaf, Robertson and Kuenen1995; Van De Graaf et al., Reference Van De Graaf, Mulder, De Bruijn, Jetten, Robertson and Kuenen1995; Lam et al., Reference Lam, Lavik, Jensen, Van De Vossenberg, Schmid, Woebken, Gutiérrez, Amann, Jetten and Kuypers2009) (Table 1). However, the anammox reaction also requires O2 because nitrite cannot be produced anaerobically (Stüeken et al., Reference Stüeken, Kipp, Koehler and Buick2016). Ammonium oxidation may also be coupled to the reduction of sulfate (i.e. sulfammox) or Fe(III) (i.e. feammox) (Clement et al., Reference Clement, Shrestha, Ehrenfeld and Jaffe2005; Yang et al., Reference Yang, Weber and Silver2012; Rios-Del Toro et al., Reference Rios-Del Toro, Valenzuela, López-Lozano, Cortés-Martínez, Sánchez-Rodríguez, Calvario-Martínez, Sánchez-Carrillo and Cervantes2018). The latter was suggested as the possibly dominant ammonium oxidation pathway in the early Archean when O2 scarcity prevented nitrite and sulfate accumulation (Pellerin et al., Reference Pellerin, Thomazo, Ader, Marin-Carbonne, Alleon, Vennin and Hofmann2023). Both denitrification to N2O or N2 and anammox produce 15N-enriched residual nitrate (Stüeken et al., Reference Stüeken, Pellerin, Thomazo, Johnson, Duncanson and Schoepfer2024). Notably, the δ15N record in metasedimentary rocks can be further shifted to more positive values with increasing metamorphic grade due to the release of isotopically light ammonium or N2 (Ader et al., Reference Ader, Thomazo, Sansjofre, Busigny, Papineau, Laffont, Cartigny and Halverson2016). Ample nitrogen isotope evidence from low-grade metasedimentary rocks (greenschist facies and below) shows highly variable δ15N signatures (ca. –11 to +50‰) and a secular increase of δ15N of approximately 2‰ at ca. 2.8–2.6 Ga (Garvin et al., Reference Garvin, Buick, Anbar, Arnold and Kaufman2009; Godfrey and Falkowski, Reference Godfrey and Falkowski2009; Thomazo et al., Reference Thomazo, Ader and Philippot2011; Busigny et al., Reference Busigny, Lebeau, Ader, Krapež and Bekker2013; Stüeken et al., Reference Stüeken, Buick and Schauer2015b, Reference Stüeken, Kipp, Koehler and Buick2016; Koehler et al., Reference Koehler, Buick, Kipp, Stüeken and Zaloumis2018; Pellerin et al., Reference Pellerin, Thomazo, Ader, Rossignol, Rego, Busigny and Philippot2024). The overall increase in sedimentary δ15N values strongly suggests an increasing role of ammonium oxidation in marine environments. Sulfate was virtually unavailable as an electron acceptor during this time (e.g. Crowe et al., Reference Crowe2014) and δ15N values of up to +37.5‰ from marine sediments are inconsistent with feammox (Pellerin et al., Reference Pellerin, Thomazo, Ader, Rossignol, Rego, Busigny and Philippot2024). Therefore, this trend is best explained by the rise of nitrification, denitrification and/or anammox in the Meso- to Neoarchean (Garvin et al., Reference Garvin, Buick, Anbar, Arnold and Kaufman2009; Godfrey and Falkowski, Reference Godfrey and Falkowski2009; Thomazo et al., Reference Thomazo, Ader and Philippot2011; Busigny et al., Reference Busigny, Lebeau, Ader, Krapež and Bekker2013; Stüeken et al., Reference Stüeken, Buick and Schauer2015b, Reference Stüeken, Kipp, Koehler and Buick2016; Koehler et al., Reference Koehler, Buick, Kipp, Stüeken and Zaloumis2018; Pellerin et al., Reference Pellerin, Thomazo, Ader, Rossignol, Rego, Busigny and Philippot2024), consistent with phylogenetic reconstructions (Parsons et al., Reference Parsons, Stüeken, Rosen, Mateos and Anderson2021). These processes, including the presence of nitrate, were probably transient and limited to settings with oxygenated surface waters.

Microaerophilic Fe(II)-oxidising bacteria commonly form characteristic mineral-organic structures that consist of Fe(III) (oxydr)oxide-encrusted stalks (Chan et al., Reference Chan, Emerson and Luther2016), which show good potential for fossilisation (Picard et al., Reference Picard, Kappler, Schmid, Quaroni and Obst2015). The fossil record of these stalks in hydrothermal jaspers, where preservation of such delicate structures is most probable, extends back to at least 1.74 Ga (Little et al., Reference Little2021). The oldest putative findings of such fossils are reported in 3.77 Ga hydrothermal vent deposits (Dodd et al., Reference Dodd, Papineau, Grenne, Slack, Rittner, Pirajno, O’Neil and Little2017; Papineau et al., Reference Papineau, She, Dodd, Iacoviello, Slack, Hauri, Shearing and Little2022), although the biogenicity of these features would require free O2 in Eoarchean seafloor hydrothermal systems or a different metabolic affinity. On modern Earth, microaerophilic Fe(II) oxidisers inhabit a limited niche at opposing gradients of O2 and Fe2+ in marine and terrestrial environments like hydrothermal vents, sediment–water interfaces and soils (Kappler et al., Reference Kappler, Bryce, Mansor, Lueder, Byrne and Swanner2021). However, in Archean oxygen oases where O2 was produced in the photic zone and Fe2+ was supplied from below, this niche could have been much larger (Holm, Reference Holm1989; Konhauser et al., Reference Konhauser, Hamade, Raiswell, Morris, Ferris, Southam and Canfield2002; Dreher et al., Reference Dreher, Schad, Robbins, Konhauser, Kappler and Joshi2021).

The presence of abundant Mn oxides in late Archean sedimentary rocks is another signature of aerobic metabolism. In the absence of O2, Mn(II) may also be oxidised by a range of abiotic and other biotic mechanisms, including photooxidation by UV radiation, Mn-dependent anoxygenic photosynthesis and Mn oxidation coupled to alternative electron acceptors like nitrate (Luther et al., Reference Luther, Sundby, Lewis, Brendel and Silverberg1997; Johnson et al., Reference Johnson, Webb, Thomas, Ono, Kirschvink and Fischer2013; Daye et al., Reference Daye, Klepac-Ceraj, Pajusalu, Rowland, Farrell-Sherman, Beukes, Tamura, Fournier and Bosak2019; Liu et al., Reference Liu, Hao, Elzinga, Piotrowiak, Nanda, Yee and Falkowski2020). However, the transfer of Mn oxides from the water column to the sediments and their preservation required an oxygenated depositional environment devoid of Fe(II) and H2S, in which aerobic microorganisms dominate Mn(II) oxidation today (Jones et al., Reference Jones, Crowe, Sturm, Leslie, MacLean, Katsev, Henny, Fowle and Canfield2011; Smith and Beukes, Reference Smith and Beukes2023; Robbins et al., Reference Robbins2023; Mhlanga et al., Reference Mhlanga, Tsikos, Lee, Rouxel, Boyce, Harris and Lyons2023).

Fossil evidence for the antiquity of aerobic sulfur oxidation is scarce, probably due to the metastable nature of the product S0, the poor preservation potential of sulfur oxidiser cells upon silicification and a lack of distinct sulfur isotope signatures (Canfield, Reference Canfield, Valley and Cole2001; Cosmidis et al., Reference Cosmidis, Nims, Diercks and Templeton2019; Nims et al., Reference Nims, Lafond, Alleon, Templeton and Cosmidis2021). Nevertheless, morphological characteristics (size, cell wall structure) combined with paleoecological considerations were used to interpret carbonaceous microstructures in the 2.52 Ga Gamohaan Formation (South Africa) as sulfur-oxidising bacteria similar to Thiomargarita (Czaja et al., Reference Czaja, Beukes and Osterhout2016). Microbial sulfur oxidation is also supported by multiple sulfur isotope compositions of 3.22 Ga sulfates (Nabhan et al., Reference Nabhan, Marin-Carbonne, Mason and Heubeck2020) and possible microbially induced corrosion features on ca. 3.4 Ga detrital pyrite (Wacey et al., Reference Wacey, Saunders, Brasier and Kilburn2011). If true, this does not necessarily indicate an aerobic metabolism due to the potential coupling of sulfur oxidation with nitrate reduction. Chromium isotopes in Archean BIFs trace chromium mobilisation during the oxidative weathering of pyrite exposed on continents by 2.48 Ga (Konhauser et al., Reference Konhauser2011). It was suggested that this was due to the activity of acidophilic sulfide-oxidising bacteria (Konhauser et al., Reference Konhauser2011) or pyrite oxidation by photochemically generated Fe3+ (Hao et al., Reference Hao, Liu, Goff, Steadman, Large, Falkowski and Yee2022). In marine environments, aerobic sulfur oxidisers may have been limited to microbial mats, oxidising sulfide generated in underlying sediments (Konhauser, Reference Konhauser2007). In the open ocean, sulfide was dominantly sourced from hydrothermal systems and probably scavenged by Fe2+ before it reached oxygenated surface waters (Canfield et al., Reference Canfield, Rosing and Bjerrum2006), although euxinic conditions possibly prevailed locally in the late Archean (Reinhard et al., Reference Reinhard, Raiswell, Scott, Anbar and Lyons2009; Scott et al., Reference Scott, Bekker, Reinhard, Schnetger, Krapež, Rumble and Lyons2011). These examples highlight the fragmented nature of direct evidence for aerobic metabolism from the geological record.

Such biosignatures can be used as calibration points in phylogenetic studies exploring the emergence and diversification of aerobic microorganisms on early Earth. A Mesoarchean origin of aerobic respiration is supported by the emergence of enzymes involved in oxygen cycling to ≥2.9 Ga, i.e. at least ca. 500 m.y. before the GOE (David and Alm, Reference David and Alm2011; Wang et al., Reference Wang, Jiang, Kim, Qu, Ji, Mittenthal, Zhang and Caetano-Anollés2011; Kim et al., Reference Kim, Qin, Jiang, Chen, Xiong, Caetano-Anollés, Zhang and Caetano-Anollés2012; Jabłońska and Tawfik, Reference Jabłońska and Tawfik2021; Boden et al., Reference Boden, Konhauser, Robbins and Sánchez-Baracaldo2021). Possibly, early enzymes catalysing O2 reduction might be a detoxification mechanism for coping with oxidative stress (‘aerotolerance’) rather than aerobic respiration (Brochier-Armanet et al., Reference Brochier-Armanet, Talla and Gribaldo2009; Gribaldo et al., Reference Gribaldo, Talla and Brochier-Armanet2009; Jabłońska and Tawfik, Reference Jabłońska and Tawfik2021, Reference Jabłońska and Tawfik2022). Aerotolerance could be an adaptation to abiotic sources of reactive oxygen species and O2 on early Earth (Haqq-Misra et al., Reference Haqq-Misra, Kasting and Lee2011; He et al., Reference He, Wu, Xian, Zhu, Yang, Lv, Li and Konhauser2021, Reference He2023; Stone et al., Reference Stone, Edgar, Gould and Telling2022). Nevertheless, the early origin of these enzymes is consistent with geochemical proxies suggesting locally or transiently sufficient O2 for aerobic respiration (Anbar et al., Reference Anbar2007; Ostrander et al., Reference Ostrander, Johnson and Anbar2021), current reconstructions of the emergence of oxygenic photosynthesis at ca. 3.0 Ga (Schirrmeister et al., Reference Schirrmeister, Gugger and Donoghue2015; Sánchez-Baracaldo, Reference Sánchez-Baracaldo2015; Cardona et al., Reference Cardona, Sánchez-Baracaldo, Rutherford and Larkum2019; Garcia-Pichel et al., Reference Garcia-Pichel, Lombard, Soule, Dunaj, Wu and Wojciechowski2019; Jabłońska and Tawfik, Reference Jabłońska and Tawfik2021; Fournier et al., Reference Fournier, Moore, Rangel, Payette, Momper and Bosak2021; Boden et al., Reference Boden, Zhong, Anderson and Stüeken2024) and the recent discovery of ancestral high redox potential quinones predating the emergence of cyanobacteria (Elling et al., Reference Elling2025). Notably, for O2 to leave a proxy record, it must degas into the atmosphere (in the case of MIF-S) or affect the solubility of redox-sensitive elements (e.g. chromium, molybdenum). However, released O2 may not reach concentrations reflected in O2 proxies due to its consumption by various O2 sinks, including aerobic microorganisms. Conversely, oxidative weathering within benthic microbial mats can produce O2 proxy signals, even if O2 does not accumulate in the environment due to microbial consumption (Lalonde and Konhauser, Reference Lalonde and Konhauser2015). Therefore, aerobic microorganisms are a plausible O2 sink after the emergence of oxygenic photosynthesis, even when seawater or atmospheric O2 concentrations were too low to leave a proxy record. Aerobic microorganisms may thus have consumed O2 produced by cyanobacteria, even if the O2 flux was small due to the possible inhibition of oxygenic photosynthesis. Thus, as soon as oxygenic photosynthesis emerged ca. 3.0 Ga, an aerobic niche appeared for microorganisms to exploit. In concert, the current evidence renders it probable that the Archean aerobic biosphere represented an O2 sink that helped delay the GOE.

Synthesis and future research directions

Earth presumably accreted from reduced material but the mantle oxidised early in its history due to core formation, late accretion of relatively oxidised material, Fe(III) transfer from the lower to the upper mantle and H2 outgassing (e.g. Wade and Wood, Reference Wade and Wood2001; Frost et al., Reference Frost, Liebske, Langenhorst, McCammon, Trønnes and Rubie2004; Rubie et al., Reference Rubie, Frost, Mann, Asahara, Nimmo, Tsuno, Kegler, Holzheid and Palme2011; Sharp et al., Reference Sharp, McCubbin and Shearer2013; Pahlevan et al., Reference Pahlevan, Schaefer and Hirschmann2019). Between 3.0 and 2.0 Ga, the upper mantle probably evolved to near-modern ƒO2 (e.g. Aulbach and Stagno, Reference Aulbach and Stagno2016; Stagno and Fei, Reference Stagno and Fei2020; O’Neill and Aulbach, Reference O’Neill and Aulbach2022) (Figs 2, 3). The earliest putative evidence for localised and/or transient O2 in Earth’s surface environments appears at 3.0 Ga (Planavsky et al., Reference Planavsky2014; Ossa Ossa et al., Reference Ossa Ossa, Hofmann, Vidal, Kramers, Belyanin and Cavalazzi2016, Reference Ossa Ossa, Hofmann, Wille, Spangenberg, Bekker, Poulton, Eickmann and Schoenberg2018; Smith and Beukes, Reference Smith and Beukes2023), coinciding with phylogenetic studies on the emergence of oxygenic photosynthesis (e.g. Schirrmeister et al., Reference Schirrmeister, Gugger and Donoghue2015; Sánchez-Baracaldo, Reference Sánchez-Baracaldo2015; Fournier et al., Reference Fournier, Moore, Rangel, Payette, Momper and Bosak2021) (Fig. 4). ‘Whiffs of oxygen’ inferred from manganese enrichments and stable isotopes become more abundant at 2.6–2.5 Ga, suggesting an expansion of oxygenated surface waters (e.g. Anbar et al., Reference Anbar2007; Kendall et al., Reference Kendall, Creaser, Reinhard, Lyons and Anbar2015; Smith and Beukes, Reference Smith and Beukes2023). Atmospheric oxygenation is recorded by the disappearance of the MIF-S signal at ca. 2.3 Ga (e.g. Farquhar, Reference Farquhar2000; Pavlov and Kasting, Reference Pavlov and Kasting2002; Bekker et al., Reference Bekker, Holland, Wang, Rumble, Stein, Hannah, Coetzee and Beukes2004; Poulton et al., Reference Poulton, Bekker, Cumming, Zerkle, Canfield and Johnston2021) (Fig. 4). This timeline demonstrates a delay of several hundred million years between the first appearance of O2 in Earth’s surface environments and its atmospheric accumulation. Thus, the emergence of oxygenic photosynthesis alone cannot satisfactorily explain the timing of the GOE.

Various models were put forward to explain this delay. One set of ideas centres around a suppressed biological O2 source due to phosphorous limitation, iron toxicity or ecological factors affecting the productivity of oxygenic photosynthesis in the Archean (e.g. Swanner et al., Reference Swanner, Mloszewska, Cirpka, Schoenberg, Konhauser and Kappler2015a; Sánchez-Baracaldo, Reference Sánchez-Baracaldo2015; Reinhard et al., Reference Reinhard, Planavsky, Gill, Ozaki, Robbins, Lyons, Fischer, Wang, Cole and Konhauser2017). Other models focus on O2 sinks, like reduced gases and aqueous solutes in surface environments, which acted as effective buffers against atmospheric oxygenation before their capacity diminished over time (e.g. Holland, Reference Holland2002; Gaillard et al., Reference Gaillard, Scaillet and Arndt2011; Lee et al., Reference Lee, Yeung, McKenzie, Yokoyama, Ozaki and Lenardic2016; O’Neill and Aulbach, Reference O’Neill and Aulbach2022). However, the role of aerobic respiration, a critical O2 sink today, seems less constrained for the Archean. Studies on recent environments and microbial incubation experiments increasingly show that aerobic growth occurs at O2 concentrations inferred for late Archean oases or even under apparently anoxic conditions (e.g. Stolper et al., Reference Stolper, Revsbech and Canfield2010; Milucka et al., Reference Milucka, Kirf, Lu, Krupke, Lam, Littmann, Kuypers and Schubert2015; Berg et al., Reference Berg2019). Combined evidence from biosignatures and phylogenetic reconstructions suggests the presence of an aerobic biosphere at least since the emergence of oxygenic photosynthesis (e.g. Godfrey and Falkowski, Reference Godfrey and Falkowski2009; David and Alm, Reference David and Alm2011; Jabłońska and Tawfik, Reference Jabłońska and Tawfik2021). These aerobic microorganisms could have lived closely associated with early cyanobacteria, helping to prevent environmental oxygenation since 3.0 Ga (Berg et al., Reference Berg, Ahmerkamp, Pjevac, Hausmann, Milucka and Kuypers2022). Microbial O2 consumption was probably coupled to the oxidation of organic matter, CH4, iron, manganese and sulfur before the end of the Archean (Fig. 5). The importance of this sink must have increased over time as a direct response to progressive oxygenation of Earth’s surface environments. This, in turn, was facilitated by the solid Earth’s redox evolution, shifting volcanic gases and aqueous solutes to more oxidised species. Viewed this way, the expansion of the aerobic biosphere represents geobiological feedback to solid Earth and surface oxidation, helping to delay the GOE for several hundred M.y. after the emergence of oxygenic photosynthesis.

Figure 5. Schematic of microbial O2 sources and sinks on the late Archean Earth. Oxygenic photosynthesis is the major O2 source. Biomass from primary productivity, anaerobic respiration (i.e. dissimilatory reduction of NO3-, Mn(IV), Fe(III), SO42-), methanogenesis and abiotic sources (not indicated) yield diverse O2 sinks (i.e. Corg, NH4+, Mn(II), Fe(II), H2S, CH4). Aerobic microorganisms couple the oxidation of these sinks to the reduction of O2, forming the microbial O2 sink. The geochemical zonation on the left was redrawn from Canfield and Thamdrup (Reference Canfield and Thamdrup2009). Note that this zonation reflects the decreasing energy yield of the corresponding respiration process and may strongly overlap in natural environments, therefore it does not necessarily match the depth profile of the indicated chemical species.

Still, several questions remain concerning the efficacy of the aerobic biosphere as an O2 sink. It was previously noted that the advent of aerobic metabolisms under Archean Earth conditions does not necessarily demonstrate their environmental impact (Lyons et al., Reference Lyons, Tino, Fournier, Anderson, Leavitt, Konhauser and Stüeken2024). Indeed, the degree to which aerobic respiration could buffer O2 production by oxygenic photosynthesis is currently unknown. This highlights the need for an improved mechanistic and quantitative assessment of the aerobic biosphere as an O2 sink in the Archean. When and in what sequence did the various aerobic metabolisms emerge? What was the environmental distribution of aerobic microorganisms and how much O2 could they consume under conditions in Archean aquatic settings? How did competition between aerobic microorganisms for resources impact their activity and what are the relative roles of microbial versus abiotic O2 consumption in the environment?

A critical prerequisite for quantifying the role of the aerobic O2 sink is constraining the O2 flux. Despite significant advances, the productivity of cyanobacteria before the GOE and the spatial distribution of oxygenated environments remain open questions. Reconstructing whether the earliest molecular mechanisms for O2 reduction were coupled to energy conservation rather than just detoxification would help identify when the biosphere became a more efficient O2 sink. Protocols for biosignature detection must improve to pinpoint the earliest evidence of the various aerobic metabolisms in the geological record and serve as calibration points for phylogenetic studies on the genomes of modern (nan)aerobic microorganisms. This will help refine the evolutionary history of Earth’s aerobic biosphere. Moreover, the environmental prevalence of aerobic growth at nanomolar O2 concentrations is poorly constrained on modern Earth but may have been larger in the Archean (Berg et al., Reference Berg, Ahmerkamp, Pjevac, Hausmann, Milucka and Kuypers2022). Broad surveys of modern apparently anoxic environments, integrating geochemical and genomic evidence, will be crucial in constraining the prevalence of nanaerobic life. Co-culturing experiments of cyanobacteria and different aerobic microorganisms under Archean ocean conditions, including reduced species like Fe2+, may better constrain the past activity of aerobic respiration. These data are critical for quantitative models on Archean microbial O2 consumption, which previously did not account for the metabolic diversity of the aerobic biosphere (Goldblatt et al., Reference Goldblatt, Lenton and Watson2006; Claire et al., Reference Claire, Catling and Zahnle2006; Catling et al., Reference Catling, Claire and Zahnle2007; Daines and Lenton, Reference Daines and Lenton2016), possibly underestimating the capacity of the microbial O2 sink. Addressing these issues will help answer which aerobic microorganisms consumed how much O2, when and where in Archean environments and improve our understanding of why the GOE happened when it did.

Acknowledgements

The authors thank the associate editor and two anonymous reviewers for their insightful comments, which improved the manuscript. Jan-Peter Duda is thanked for his input on an early draft of this manuscript. Eric Runge was supported by the Emmy Noether program of the Deutsche Forschungsgemeinschaft (DFG; grant to Jan-Peter Duda; DU 1450/7-1). Sara Vulpius was funded by the DFG within the SPP 1833 “Building a Habitable Earth” project number NO 1324/2-1 and partly funded by the STRUCTURE program of Freie Universität Berlin (FUB). Daniel Herwartz was supported by the Heisenberg Program of the DFG (HE 6357/4-1). This study was also funded by the Project-ID 263649064–TRR 170 as well as by the European Union (ERC, DIVERSE, 101087755). Views and opinions expressed are, however, those of the author(s) only and do not necessarily reflect those of the European Union or the European Research Council Executive Agency. Neither the European Union nor the granting authority can be held responsible.

Competing interests

The authors declare none.

References

Ader, M. Thomazo, C., Sansjofre, P., Busigny, V., Papineau, D., Laffont, R., Cartigny, P. and Halverson, G.P. (2016) Interpretation of the nitrogen isotopic composition of Precambrian sedimentary rocks: Assumptions and perspectives. Chemical Geology, 429, 93110, doi:10.1016/j.chemgeo.2016.02.010.CrossRefGoogle Scholar
Aftabi, A., Atapour, H., Mohseni, S. and Babaki, A. (2021) Geochemical discrimination among different types of banded iron formations (BIFs): A comparative review. Ore Geology Reviews, 136, 104244, doi:10.1016/j.oregeore2021.104244.CrossRefGoogle Scholar
Albut, G., Babechuk, M.G., Kleinhanns, I.C., Benger, M., Beukes, N.J., Steinhilber, B., Smith, A.J.B., Kruger, S.J. and Schoenberg, R. (2018) Modern rather than Mesoarchaean oxidative weathering responsible for the heavy stable Cr isotopic signatures of the 2.95 Ga old Ijzermijn iron formation (South Africa). Geochimica et Cosmochimica Acta, 228, 157189, doi:10.1016/j.gca.2018.02.034.CrossRefGoogle Scholar
Albut, G., Kamber, B.S., Brüske, A., Beukes, N.J., Smith, A.J.B. and Schoenberg, R. (2019) Modern weathering in outcrop samples versus ancient paleoredox information in drill core samples from a Mesoarchaean marine oxygen oasis in Pongola Supergroup, South Africa. Geochimica et Cosmochimica Acta, 265, 330353, doi:10.1016/j.gca.2019.09.001.CrossRefGoogle Scholar
Anbar, A.D. et al. (2007) A Whiff of Oxygen Before the Great Oxidation Event? Science, 317, 19031906, doi:10.1126/science.1140325.CrossRefGoogle ScholarPubMed
Anbar, A.D. and Holland, H.D. (1992) The photochemistry of manganese and the origin of banded iron formations. Geochimica et Cosmochimica Acta, 56, 25952603, doi:10.1016/0016-7037(92)90346-K.CrossRefGoogle ScholarPubMed
Andrault, D., Muñoz, M., Pesce, G., Cerantola, V., Chumakov, A., Kantor, I., Pascarelli, S., Rüffer, R. and Hennet, L. (2018) Large oxygen excess in the primitive mantle could be the source of the Great Oxygenation Event. Geochemical Perspectives Letters, 6, 510, doi:10.7185/geochemlet.1801.CrossRefGoogle Scholar
Arculus, R.J. (1985) Oxidation status of the mantle: past and present. Annual Review of Earth and Planetary Sciences, 13, 7595.10.1146/annurev.ea.13.050185.000451CrossRefGoogle Scholar
Ardia, P., Hirschmann, M.M., Withers, A.C. and Stanley, B.D. (2013) Solubility of CH4 in a synthetic basaltic melt, with applications to atmosphere–magma ocean–core partitioning of volatiles and to the evolution of the Martian atmosphere. Geochimica et Cosmochimica Acta, 114, 5271, doi:10.1016/j.gca.2013.03.028.CrossRefGoogle Scholar
Armstrong, K., Frost, D.J., McCammon, C.A., Rubie, D.C. and Boffa Ballaran, T. (2019) Deep magma ocean formation set the oxidation state of Earth’s mantle. Science, 365, 903906, doi:10.1126/science.aax8376.CrossRefGoogle ScholarPubMed
Aulbach, S. and Stagno, V. (2016) Evidence for a reducing Archean ambient mantle and its effects on the carbon cycle. Geology, 44, 751754, doi:10.1130/G38070.1.CrossRefGoogle Scholar
Aulbach, S. and Viljoen, K.S. (2015) Eclogite xenoliths from the Lace kimberlite, Kaapvaal craton: From convecting mantle source to palaeo-ocean floor and back. Earth and Planetary Science Letters, 431, 274286, doi:10.1016/j.epsl.2015.08.039.CrossRefGoogle Scholar
Aulbach, S., Woodland, A.B., Vasilyev, P., Galvez, M.E. and Viljoen, K.S. (2017) Effects of low-pressure igneous processes and subduction on Fe3+/ΣFe and redox state of mantle eclogites from Lace (Kaapvaal craton). Earth and Planetary Science Letters, 474, 283295, doi:10.1016/j.epsl.2017.06.030.CrossRefGoogle Scholar
Badro, J., Brodholt, J.P., Piet, H., Siebert, J. and Ryerson, F.J. (2015) Core formation and core composition from coupled geochemical and geophysical constraints. Proceedings of the National Academy of Sciences, 112, 1231012314, doi:10.1073/pnas.1505672112.CrossRefGoogle ScholarPubMed
Ballhaus, C., Berry, R.F. and Green, D.H. (1991) High pressure experimental calibration of the olivine-orthopyroxene-spinel oxygen geobarometer: implications for the oxidation state of the upper mantle. Contributions to Mineralogy and Petrology, 107, 2740.10.1007/BF00311183CrossRefGoogle Scholar
Ballhaus, C. and Frost, B.R. (1994) The generation of oxidized CO2-bearing basaltic melts from reduced CH4-bearing upper mantle sources. Geochimica et Cosmochimica Acta, 58, 49314940, doi:10.1016/0016-7037(94)90222-4.CrossRefGoogle Scholar
Ballmer, M.D., Houser, C., Hernlund, J.W., Wentzcovitch, R.M. and Hirose, K. (2017) Persistence of strong silica-enriched domains in the Earth’s lower mantle. Nature Geoscience, 10, 236240, doi:10.1038/ngeo2898.CrossRefGoogle Scholar
Bao, H., Rumble, D. and Lowe, D.R. (2007) The five stable isotope compositions of Fig Tree barites: Implications on sulfur cycle in ca. 3.2Ga oceans. Geochimica et Cosmochimica Acta, 71, 48684879, doi:10.1016/j.gca.2007.05.032.CrossRefGoogle Scholar
Barrett, T.J., Fralick, P.W. and Jarvis, I. (1988) Rare-earth-element geochemistry of some Archean iron formations north of Lake Superior, Ontario. Canadian Journal of Earth Sciences, 25, 570580, doi:10.1139/e88-055.CrossRefGoogle Scholar
Barth, P., Stüeken, E.E., Helling, C., Rossmanith, L., Peng, Y., Walters, W. and Claire, M. (2023) Isotopic constraints on lightning as a source of fixed nitrogen in Earth’s early biosphere: Nature Geoscience, 16, 478484, doi:10.1038/s41561-023-01187-2.CrossRefGoogle Scholar
Bau, M., Frei, R., Garbe-Schönberg, D. and Viehmann, S. (2022) High-resolution Ge-Si-Fe, Cr isotope and Th-U data for the Neoarchean Temagami BIF, Canada, suggest primary origin of BIF bands and oxidative terrestrial weathering 2.7 Ga ago. Earth and Planetary Science Letters, 589, 117579, doi:10.1016/j.epsl.2022.117579.CrossRefGoogle Scholar
Baughn, A.D. and Malamy, M.H. (2004) The strict anaerobe Bacteroides fragilis grows in and benefits from nanomolar concentrations of oxygen. Nature, 427, 441444, doi:10.1038/nature02285.CrossRefGoogle ScholarPubMed
Bekker, A. (2001) Chemostratigraphy of the Paleoproterozoic Duitschland Formation, South Africa: implications for coupled climate change and carbon cycling. American Journal of Science, 301, 261285, doi:10.2475/ajs.301.3.261.CrossRefGoogle Scholar
Bekker, A., Holland, H.D., Wang, P.-L., Rumble, D., Stein, H.J., Hannah, J.L., Coetzee, L.L. and Beukes, N.J. (2004) Dating the rise of atmospheric oxygen. Nature, 427, 117120, doi:10.1038/nature02260.CrossRefGoogle ScholarPubMed
Bekker, A., Slack, J.F., Planavsky, N., Krapez, B., Hofmann, A., Konhauser, K.O. and Rouxel, O.J. (2010) Iron formation: the sedimentary product of a complex interplay among mantle, tectonic, oceanic, and biospheric processes. Economic Geology, 105, 467508, doi:10.2113/gsecongeo.105.3.467.CrossRefGoogle Scholar
Beman, J.M., Leilei Shih, J. and Popp, B.N. (2013) Nitrite oxidation in the upper water column and oxygen minimum zone of the eastern tropical North Pacific Ocean. The ISME Journal, 7, 21922205, doi:10.1038/ismej.2013.96.CrossRefGoogle ScholarPubMed
Berg, J.S. et al. (2019) Dark aerobic sulfide oxidation by anoxygenic phototrophs in anoxic waters. Environmental Microbiology, 21, 16111626, doi:10.1111/1462-2920.14543.CrossRefGoogle ScholarPubMed
Berg, J.S., Ahmerkamp, S., Pjevac, P., Hausmann, B., Milucka, J. and Kuypers, M.M.M. (2022) How low can they go? Aerobic respiration by microorganisms under apparent anoxia. FEMS Microbiology Reviews, 46, fuac006, doi:10.1093/femsre/fuac006.CrossRefGoogle Scholar
Berner, R.A. (1989) Biogeochemical cycles of carbon and sulfur and their effect on atmospheric oxygen over Phanerozoic time. Global and Planetary Change, 75, 97122.10.1016/0921-8181(89)90018-0CrossRefGoogle Scholar
Bindeman, I.N., Schmitt, A.K. and Evans, D.A.D. (2010) Limits of hydrosphere-lithosphere interaction: Origin of the lowest-known δ18O silicate rock on Earth in the Paleoproterozoic Karelian rift. Geology, 38, 631634, doi:10.1130/G30968.1.CrossRefGoogle Scholar
Bjerrum, C.J. and Canfield, D.E. (2002) Ocean productivity before about 1.9 Gyr ago limited by phosphorus adsorption onto iron oxides. Nature, 417, 159162, doi:10.1038/417159a.CrossRefGoogle ScholarPubMed
Blank, C.E. and Sánchez-Baracaldo, P. (2010) Timing of morphological and ecological innovations in the cyanobacteria – a key to understanding the rise in atmospheric oxygen. Geobiology, 8, 123, doi:10.1111/j.1472-4669.2009.00220.x.CrossRefGoogle ScholarPubMed
Blundy, J.D., Brodholt, J.P. and Wood, B.J. (1991) Carbon–fluid equilibria and the oxidation state of the upper mantle. Nature, 349, 321324, doi:10.1038/349321a0.CrossRefGoogle Scholar
Boden, J.S., Konhauser, K.O., Robbins, L.J. and Sánchez-Baracaldo, P. (2021) Timing the evolution of antioxidant enzymes in cyanobacteria. Nature Communications, 12, 4742, doi:10.1038/s41467-021-24396-y.CrossRefGoogle ScholarPubMed
Boden, J.S., Zhong, J., Anderson, R.E. and Stüeken, E.E. (2024) Timing the evolution of phosphorus-cycling enzymes through geological time using phylogenomics. Nature Communications, 15, 3703, doi:10.1038/s41467-024-47914-0.CrossRefGoogle ScholarPubMed
Bolhar, R., Van Kranendonk, M.J. and Kamber, B.S. (2005) A trace element study of siderite–jasper banded iron formation in the 3.45Ga Warrawoona Group, Pilbara Craton—Formation from hydrothermal fluids and shallow seawater. Precambrian Research, 137, 93114, doi:10.1016/j.precamres.2005.02.001.CrossRefGoogle Scholar
Brasier, M., Green, O., Lindsay, J., Mcloughlin, N., Steele, A. and Stoakes, C. (2005) Critical testing of Earth’s oldest putative fossil assemblage from the ∼3.5Ga Apex chert, Chinaman Creek, Western Australia. Precambrian Research, 140, 55102, doi:10.1016/j.precamres.2005.06.008.CrossRefGoogle Scholar
Braterman, P.S., Cairns-Smith, A.G. and Sloper, R.W. (1983) Photo-oxidation of hydrated Fe2+—significance for banded iron formations. Nature, 303, 163164, doi:10.1038/303163a0.CrossRefGoogle Scholar
Breuer, D. and Spohn, T. (1995) Possible flush instability in mantle convection at the Archaean–Proterozoic transition. Nature, 378, 608610, doi:10.1038/378608a0.CrossRefGoogle Scholar
Bristow, L.A. et al. (2016) Ammonium and nitrite oxidation at nanomolar oxygen concentrations in oxygen minimum zone waters. Proceedings of the National Academy of Sciences, 113, 1060110606, doi:10.1073/pnas.1600359113.CrossRefGoogle ScholarPubMed
Brochier-Armanet, C., Talla, E. and Gribaldo, S. (2009) The multiple evolutionary histories of dioxygen reductases: implications for the origin and evolution of aerobic respiration. Molecular Biology and Evolution, 26, 285297, doi:10.1093/molbev/msn246.CrossRefGoogle ScholarPubMed
Brocks, J.J. (1999) Archean molecular fossils and the early rise of eukaryotes. Science, 285, 10331036, doi:10.1126/science.285.5430.1033.CrossRefGoogle ScholarPubMed
Brocks, J.J. (2011) Millimeter-scale concentration gradients of hydrocarbons in Archean shales: Live-oil escape or fingerprint of contamination? Geochimica et Cosmochimica Acta, 75, 31963213, doi:10.1016/j.gca.2011.03.014.CrossRefGoogle Scholar
Brocks, J.J. Buick, R. Logan, G.A. and Summons, R.E. (2003) Composition and syngeneity of molecular fossils from the 2.78 to 2.45 billion-year-old Mount Bruce Supergroup, Pilbara Craton, Western Australia. Geochimica et Cosmochimica Acta, 67(21), 42894319.10.1016/S0016-7037(03)00208-4CrossRefGoogle Scholar
Brüske, A. et al. (2020) The onset of oxidative weathering traced by uranium isotopes. Precambrian Research, 338, 105583, doi:10.1016/j.precamres.2019.105583.CrossRefGoogle Scholar
Budde, G., Burkhardt, C. and Kleine, T. (2019) Molybdenum isotopic evidence for the late accretion of outer Solar System material to Earth. Nature Astronomy, 3, 736741, doi:10.1038/s41550-019-0779-y.CrossRefGoogle Scholar
Buick, R. (1992) The antiquity of oxygenic photosynthesis: evidence from stromatolites in sulphate-deficient Archaean lakes. Science, 255, 7477, doi:10.1126/science.11536492.CrossRefGoogle ScholarPubMed
Burgisser, A. and Scaillet, B. (2007) Redox evolution of a degassing magma rising to the surface. Nature, 445, 194197, doi:10.1038/nature05509.CrossRefGoogle ScholarPubMed
Busigny, V., Lebeau, O., Ader, M., Krapež, B. and Bekker, A. (2013) Nitrogen cycle in the Late Archean ferruginous ocean. Chemical Geology, 362, 115130, doi:10.1016/j.chemgeo.2013.06.023.CrossRefGoogle Scholar
Cairns-Smith, A.G. (1978) Precambrian solution photochemistry, inverse segregation, and banded iron formations. Nature, 276, 807808, doi:10.1038/276807a0.CrossRefGoogle Scholar
Callbeck, C.M. et al. (2018) Oxygen minimum zone cryptic sulfur cycling sustained by offshore transport of key sulfur oxidizing bacteria. Nature Communications, 9, 1729, doi:10.1038/s41467-018-04041-x.CrossRefGoogle ScholarPubMed
Canfield, D.E. (2001) Biogeochemistry of sulfur isotopes. Pp. 607636 in: Reviews in Mineralogy and Geochemistry, Vol. 43(1) (Valley, J.W. and Cole, D.R., editors). Mineralogical Society of America.Google Scholar
Canfield, D.E. (2005) The early history of atmospheric oxygen: Homage to Robert M. Garrels. Annual Review of Earth and Planetary Sciences, 33, 136, doi:10.1146/annureearth.33.092203.122711.CrossRefGoogle Scholar
Canfield, D.E., Rosing, M.T. and Bjerrum, C. (2006) Early anaerobic metabolisms. Philosophical Transactions of the Royal Society B: Biological Sciences, 361, 18191836, doi:10.1098/rstb.2006.1906.CrossRefGoogle ScholarPubMed
Canfield, D.E. and Thamdrup, B. (2009) Towards a consistent classification scheme for geochemical environments, or, why we wish the term ‘suboxic’ would go away. Geobiology, 7, 385392, doi:10.1111/j.1472-4669.2009.00214.x.CrossRefGoogle ScholarPubMed
Canil, D. (1997) Vanadium partitioning and the oxidation state of Archaean komatiite magmas. Nature, 389, 842845, doi:10.1038/39860.CrossRefGoogle Scholar
Cardona, T., Sánchez-Baracaldo, P., Rutherford, A.W. and Larkum, A.W. (2019) Early Archean origin of Photosystem II. Geobiology, 17, 127150, doi:10.1111/gbi.12322.CrossRefGoogle ScholarPubMed
Carmichael, I.S.E. (1991) The redox states of basic and silicic magmas: a reflection of their source regions? Contributions to Mineralogy and Petrology, 106, 129141, doi:10.1007/BF00306429.CrossRefGoogle Scholar
Cartier, C., Hammouda, T., Boyet, M., Bouhifd, M.A. and Devidal, J.-L. (2014) Redox control of the fractionation of niobium and tantalum during planetary accretion and core formation. Nature Geoscience, 7, 573576, doi:10.1038/ngeo2195.CrossRefGoogle Scholar
Casciotti, K.L. (2009) Inverse kinetic isotope fractionation during bacterial nitrite oxidation. Geochimica et Cosmochimica Acta, 73, 20612076, doi:10.1016/j.gca.2008.12.022.CrossRefGoogle Scholar
Castresana, J. and Saraste, M. (1995) Evolution of energetic metabolism: the respiration-early hypothesis. Trends in Biochemical Sciences, 20, 443448.10.1016/S0968-0004(00)89098-2CrossRefGoogle Scholar
Cates, N.L. and Mojzsis, S.J. (2006) Chemical and isotopic evidence for widespread Eoarchean metasedimentary enclaves in southern West Greenland. Geochimica et Cosmochimica Acta, 70, 42294257, doi:10.1016/j.gca.2006.05.014.CrossRefGoogle Scholar
Catling, D.C., Claire, M.W. and Zahnle, K.J. (2007) Anaerobic methanotrophy and the rise of atmospheric oxygen. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 365, 18671888, doi:10.1098/rsta.2007.2047.CrossRefGoogle ScholarPubMed
Catling, D.C. and Zahnle, K.J. (2020) The Archean atmosphere. Science Advances, 6, eaax1420, doi:10.1126/sciadaax1420.CrossRefGoogle ScholarPubMed
Cavosie, A.J., Valley, J.W., Wilde, S.A. and E.I.M.F. (2005) Magmatic δ18O in 4400–3900 Ma detrital zircons: A record of the alteration and recycling of crust in the Early Archean. Earth and Planetary Science Letters, 235, 663681, doi:10.1016/j.epsl.2005.04.028.CrossRefGoogle Scholar
Chan, C.S., Emerson, D. and Luther, G.W. (2016) The role of microaerophilic Fe-oxidizing micro-organisms in producing banded iron formations. Geobiology, 14, 509528, doi:10.1111/gbi.12192.CrossRefGoogle ScholarPubMed
Chiu, B.K., Kato, S., McAllister, S.M., Field, E.K. and Chan, C.S. (2017) Novel pelagic iron-oxidizing zetaproteobacteria from the Chesapeake Bay oxic–anoxic transition zone. Frontiers in Microbiology, 8, 1280, doi:10.3389/fmicb.2017.01280.CrossRefGoogle ScholarPubMed
Christie, D.M., Carmichael, I.S.E. and Langmuir, C.H. (1986) Oxidation states of mid-ocean ridge basalt glasses. Earth and Planetary Science Letters, 79, 397411, doi:10.1016/0012-821X(86)90195-0.CrossRefGoogle Scholar
Claire, M.W., Catling, D.C. and Zahnle, K.J. (2006) Biogeochemical modelling of the rise in atmospheric oxygen. Geobiology, 4, 239269, doi:10.1111/j.1472-4669.2006.00084.x.CrossRefGoogle Scholar
Clement, B.G., Luther, G.W. and Tebo, B.M. (2009) Rapid, oxygen-dependent microbial Mn(II) oxidation kinetics at sub-micromolar oxygen concentrations in the Black Sea suboxic zone. Geochimica et Cosmochimica Acta, 73, 18781889, doi:10.1016/j.gca.2008.12.023.CrossRefGoogle Scholar
Clement, J., Shrestha, J., Ehrenfeld, J. and Jaffe, P. (2005) Ammonium oxidation coupled to dissimilatory reduction of iron under anaerobic conditions in wetland soils. Soil Biology and Biochemistry, 37, 23232328, doi:10.1016/j.soilbio.2005.03.027.CrossRefGoogle Scholar
Cloud, P. (1973) Paleoecological significance of the banded iron-formation. Economic Geology, 68, 11351143, doi:10.2113/gsecongeo.68.7.1135.CrossRefGoogle Scholar
Collerson, K.D. and Kamber, B.S. (1999) Evolution of the continents and the atmosphere inferred from Th-U-Nb systematics of the depleted mantle. Science, 283, 15191522, doi:10.1126/science.283.5407.1519.CrossRefGoogle ScholarPubMed
Condie, K. (2001) Precambrian superplumes and supercontinents: a record in black shales, carbon isotopes, and paleoclimates? Precambrian Research, 106, 239260, doi:10.1016/S0301-9268(00)00097-8.CrossRefGoogle Scholar
Cosmidis, J., Nims, C.W., Diercks, D. and Templeton, A.S. (2019) Formation and stabilization of elemental sulfur through organomineralization. Geochimica et Cosmochimica Acta, 247, 5982, doi:10.1016/j.gca.2018.12.025.CrossRefGoogle Scholar
Cottrell, E. and Kelley, K.A. (2011) The oxidation state of Fe in MORB glasses and the oxygen fugacity of the upper mantle. Earth and Planetary Science Letters, 305, 270282, doi:10.1016/j.epsl.2011.03.014.CrossRefGoogle Scholar
Crockford, P. and Halevy, I. (2022) Questioning the paradigm of a phosphate-limited Archean biosphere. Geophysical Research Letters, 49, e2022GL099818, doi:10.1029/2022GL099818.CrossRefGoogle Scholar
Crowe, S.A. et al. (2014) Sulfate was a trace constituent of Archean seawater. Science, 346, 735739.10.1126/science.1258966CrossRefGoogle ScholarPubMed
Crowe, S.A. et al. (2011) The methane cycle in ferruginous Lake Matano: methane cycle in ferruginous Lake Matano. Geobiology, 9, 6178, doi:10.1111/j.1472-4669.2010.00257.x.CrossRefGoogle ScholarPubMed
Crowe, S.A., Døssing, L.N., Beukes, N.J., Bau, M., Kruger, S.J., Frei, R. and Canfield, D.E. (2013) Atmospheric oxygenation three billion years ago. Nature, 501, 535538, doi:10.1038/nature12426.CrossRefGoogle ScholarPubMed
Cypionka, H. (2000) Oxygen respiration by Desulfovibrio species. Annual Review of Microbiology, 54, 827848, doi:10.1146/annuremicro.54.1.827.CrossRefGoogle ScholarPubMed
Czaja, A.D., Beukes, N.J. and Osterhout, J.T, (2016) Sulfur-oxidizing bacteria prior to the Great Oxidation Event from the 2.52 Ga Gamohaan Formation of South Africa. Geology, 44, 983986, doi:10.1130/G38150.1.CrossRefGoogle Scholar
Dahl, C., Friedrich, C. and Kletzin, A. (2008) Sulfur oxidation in Prokaryotes. In: Encyclopedia of Life Sciences (ELS). John Wiley & Sons, Ltd, doi:10.1002/9780470015902.a0021155.Google Scholar
Daines, S.J. and Lenton, T.M. (2016) The effect of widespread early aerobic marine ecosystems on methane cycling and the Great Oxidation. Earth and Planetary Science Letters, 434, 4251, doi:10.1016/j.epsl.2015.11.021.CrossRefGoogle Scholar
Dauphas, N. (2017) The isotopic nature of the Earth’s accreting material through time. Nature, 541, 521524, doi:10.1038/nature20830.CrossRefGoogle ScholarPubMed
David, L.A. and Alm, E.J. (2011) Rapid evolutionary innovation during an Archaean genetic expansion. Nature, 469, 9396, doi:10.1038/nature09649.CrossRefGoogle ScholarPubMed
Daye, M., Klepac-Ceraj, V., Pajusalu, M., Rowland, S., Farrell-Sherman, A., Beukes, N., Tamura, N., Fournier, G. and Bosak, T. (2019) Light-driven anaerobic microbial oxidation of manganese. Nature, 576, 311314, doi:10.1038/s41586-019-1804-0.CrossRefGoogle ScholarPubMed
Debaille, V., O’Neill, C., Brandon, A.D., Haenecour, P., Yin, Q.-Z., Mattielli, N. and Treiman, A.H. (2013) Stagnant-lid tectonics in early Earth revealed by 142Nd variations in late Archean rocks. Earth and Planetary Science Letters, 373, 8392, doi:10.1016/j.epsl.2013.04.016.CrossRefGoogle Scholar
Delano, J.W. (2001) Redox History of the Earth’s Interior Since ∼3900 Ma: Implications for Prebiotic Molecules. Origins of Life and Evolution of Biospheres, 31, 311341.10.1023/A:1011895600380CrossRefGoogle ScholarPubMed
Dellwig, O., Leipe, T., März, C., Glockzin, M., Pollehne, F., Schnetger, B., Yakushev, E.V., Böttcher, M.E. and Brumsack, H.-J. (2010) A new particulate Mn–Fe–P-shuttle at the redoxcline of anoxic basins. Geochimica et Cosmochimica Acta, 74, 71007115, doi:10.1016/j.gca.2010.09.017.CrossRefGoogle Scholar
Deng, J., Du, Z., Karki, B.B., Ghosh, D.B. and Lee, K.K.M. (2020) A magma ocean origin to divergent redox evolutions of rocky planetary bodies and early atmospheres. Nature Communications, 11, doi:10.1038/s41467-020-15757-0.CrossRefGoogle ScholarPubMed
Derry, L.A. (2015) Causes and consequences of mid-Proterozoic anoxia. Geophysical Research Letters, 42, 85388546, doi:10.1002/2015GL065333.CrossRefGoogle Scholar
Devol, A.H. (1978) Bacterial oxygen uptake kinetics as related to biological processes in oxygen deficient zones of the oceans. Deep Sea Research, 25, 137146, doi:10.1016/0146-6291(78)90001-2.CrossRefGoogle Scholar
Dick, G.J., Grim, S.L. and Klatt, J.M. (2018) Controls on O2 production in cyanobacterial mats and implications for Earth’s oxygenation. Annual Review of Earth and Planetary Sciences, 46, 123147, doi:10.1146/annurev-earth-082517-010035.CrossRefGoogle Scholar
Dodd, M.S. et al. (2022) Abiotic anoxic iron oxidation, formation of Archean banded iron formations, and the oxidation of early Earth. Earth and Planetary Science Letters, 584, 117469, doi:10.1016/j.epsl.2022.117469.CrossRefGoogle Scholar
Dodd, M.S., Papineau, D., Grenne, T., Slack, J.F., Rittner, M., Pirajno, F., O’Neil, J. and Little, C.T.S. (2017) Evidence for early life in Earth’s oldest hydrothermal vent precipitates. Nature, 543, 6064, doi:10.1038/nature21377.CrossRefGoogle ScholarPubMed
Domagal-Goldman, S.D., Kasting, J.F., Johnston, D.T. and Farquhar, J. (2008) Organic haze, glaciations and multiple sulfur isotopes in the Mid-Archean Era. Earth and Planetary Science Letters, 269, 2940, doi:10.1016/j.epsl.2008.01.040.CrossRefGoogle Scholar
Dreher, C.L., Schad, M., Robbins, L.J., Konhauser, K.O., Kappler, A. and Joshi, P. (2021) Microbial processes during deposition and diagenesis of Banded Iron Formations. PalZ, 95, 593610, doi:10.1007/s12542-021-00598-z.CrossRefGoogle ScholarPubMed
Druschel, G.K., Emerson, D., Sutka, R., Suchecki, P. and Luther, G.W. (2008) Low-oxygen and chemical kinetic constraints on the geochemical niche of neutrophilic iron(II) oxidizing microorganisms. Geochimica et Cosmochimica Acta, 72, 33583370, doi:10.1016/j.gca.2008.04.035.CrossRefGoogle Scholar
Ducluzeau, A.-L., Schoepp-Cothenet, B., Van Lis, R., Baymann, F., Russell, M.J. and Nitschke, W. (2014) The evolution of respiratory O2 /NO reductases: an out-of-the-phylogenetic-box perspective. Journal of The Royal Society Interface, 11, 20140196, doi:10.1098/rsif.2014.0196.CrossRefGoogle ScholarPubMed
Duncan, M.S. and Dasgupta, R. (2017) Rise of Earth’s atmospheric oxygen controlled by efficient subduction of organic carbon. Nature Geoscience, 10, 387392, doi:10.1038/ngeo2939.CrossRefGoogle Scholar
Eigenbrode, J.L. and Freeman, K.H. (2006) Late Archean rise of aerobic microbial ecosystems. Proceedings of the National Academy of Sciences, 103, 1575915764, doi:10.1073/pnas.0607540103.CrossRefGoogle ScholarPubMed
Eigenbrode, J.L., Freeman, K.H. and Summons, R.E. (2008) Methylhopane biomarker hydrocarbons in Hamersley Province sediments provide evidence for Neoarchean aerobiosis. Earth and Planetary Science Letters, 273, 323331, doi:10.1016/j.epsl.2008.06.037.CrossRefGoogle Scholar
Elling, F.J. et al. (2025) A novel quinone biosynthetic pathway illuminates the evolution of aerobic metabolism. Proceedings of the National Academy of Sciences, 122, e2421994122, doi:10.1073/pnas.2421994122.CrossRefGoogle ScholarPubMed
Ernst, R. and Bleeker, W. (2010) Large igneous provinces (LIPs), giant dyke swarms, and mantle plumes: significance for breakup events within Canada and adjacent regions from 2.5 Ga to the Present. Canadian Journal of Earth Sciences, 47, 695739, doi:10.1139/E10-025.CrossRefGoogle Scholar
Ershov, B.G. (2021) Radiation-induced oxidation of iron in the ocean of the early Earth. Precambrian Research, 364, 106360, doi:10.1016/j.precamres.2021.106360.CrossRefGoogle Scholar
Evans, G.N., Coogan, L.A., Kaçar, B. and Seyfried, W.E. (2023) Molybdenum in basalt-hosted seafloor hydrothermal systems: Experimental, theoretical, and field sampling approaches. Geochimica et Cosmochimica Acta, 353, 2844, doi:10.1016/j.gca.2023.05.018.CrossRefGoogle Scholar
Falkowski, P.G. (1997) Evolution of the nitrogen cycle and its influence on the biological sequestration of CO2 in the ocean. Nature, 387, 272275, doi:10.1038/387272a0.CrossRefGoogle Scholar
Falkowski, P.G., Fenchel, T. and Delong, E.F. (2008) The microbial engines that drive Earth’s biogeochemical cycles. Science, 320, 10341039, doi:10.1126/science.1153213.CrossRefGoogle ScholarPubMed
Farquhar, J. (2000) Atmospheric influence of Earth’s earliest sulfur cycle. Science, 289, 756758, doi:10.1126/science.289.5480.756.CrossRefGoogle ScholarPubMed
Farquhar, J., Johnston, D.T. and Wing, B.A. (2007) Implications of conservation of mass effects on mass-dependent isotope fractionations: Influence of network structure on sulfur isotope phase space of dissimilatory sulfate reduction. Geochimica et Cosmochimica Acta, 71, 58625875, doi:10.1016/j.gca.2007.08.028.CrossRefGoogle Scholar
Farquhar, J., Savarino, J., Airieau, S. and Thiemens, M.H. (2001) Observation of wavelength-sensitive mass-independent sulfur isotope effects during SO2 photolysis: Implications for the early atmosphere. Journal of Geophysical Research, 106, 3282932839.10.1029/2000JE001437CrossRefGoogle Scholar
Field, C.B., Behrenfeld, M.J., Randerson, J.T. and Falkowski, P. (1998) primary production of the biosphere: integrating terrestrial and oceanic components. Science, 281, 237240, doi:10.1126/science.281.5374.237.CrossRefGoogle ScholarPubMed
Fischer, R.A., Nakajima, Y., Campbell, A.J., Frost, D.J., Harries, D., Langenhorst, F., Miyajima, N., Pollok, K. and Rubie, D.C. (2015) High pressure metal–silicate partitioning of Ni, Co, V, Cr, Si, and O. Geochimica et Cosmochimica Acta, 167, 177194, doi:10.1016/j.gca.2015.06.026.CrossRefGoogle Scholar
Fischer-Gödde, M., Elfers, B.-M., Münker, C., Szilas, K., Maier, W.D., Messling, N., Morishita, T., Van Kranendonk, M. and Smithies, H. (2020) Ruthenium isotope vestige of Earth’s pre-late-veneer mantle preserved in Archaean rocks. Nature, 579, 240244, doi:10.1038/s41586-020-2069-3.CrossRefGoogle ScholarPubMed
Fischer-Gödde, M. and Kleine, T. (2017) Ruthenium isotopic evidence for an inner Solar System origin of the late veneer. Nature, 541, 525527, doi:10.1038/nature21045.CrossRefGoogle ScholarPubMed
Flannery, D.T., Allwood, A.C., Summons, R.E., Williford, K.H., Abbey, W., Matys, E.D. and Ferralis, N. (2018) Spatially-resolved isotopic study of carbon trapped in ∼3.43 Ga Strelley Pool Formation stromatolites. Geochimica et Cosmochimica Acta, 223, 2135, doi:10.1016/j.gca.2017.11.028.CrossRefGoogle Scholar
Foley, S.F. (2011) A reappraisal of redox melting in the Earth’s mantle as a function of tectonic setting and time. Journal of Petrology, 52, 13631391, doi:10.1093/petrology/egq061.CrossRefGoogle Scholar
Foley, B.J. and Rizo, H. (2017) Long-term preservation of early formed mantle heterogeneity by mobile lid convection: Importance of grainsize evolution. Earth and Planetary Science Letters, 475, 94105, doi:10.1016/j.epsl.2017.07.031.CrossRefGoogle Scholar
Fournier, G.P., Moore, K.R., Rangel, L.T., Payette, J.G., Momper, L. and Bosak, T. (2021) The Archean origin of oxygenic photosynthesis and extant cyanobacterial lineages. Proceedings of the Royal Society B: Biological Sciences, 288, 20210675, doi:10.1098/rspb.2021.0675.CrossRefGoogle ScholarPubMed
François, L.M. (1986) Extensive deposition of banded iron formations was possible without photosynthesis. Nature, 320, 352354, doi:10.1038/320352a0.CrossRefGoogle Scholar
Frei, R., Gaucher, C., Poulton, S.W. and Canfield, D.E. (2009) Fluctuations in Precambrian atmospheric oxygenation recorded by chromium isotopes. Nature, 461, 250253, doi:10.1038/nature08266.CrossRefGoogle ScholarPubMed
French, K.L. et al. (2015) Reappraisal of hydrocarbon biomarkers in Archean rocks. Proceedings of the National Academy of Sciences, 112, 59155920, doi:10.1073/pnas.1419563112.CrossRefGoogle ScholarPubMed
Frost, D.J., Liebske, C., Langenhorst, F., McCammon, C.A., Trønnes, R.G. and Rubie, D.C. (2004) Experimental evidence for the existence of iron-rich metal in the Earth’s lower mantle. Nature, 428, 409412, doi:10.1038/nature02413.CrossRefGoogle ScholarPubMed
Frost, D.J., Mann, U., Asahara, Y. and Rubie, D.C. (2008) The redox state of the mantle during and just after core formation. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 366, 43154337, doi:10.1098/rsta.2008.0147.CrossRefGoogle ScholarPubMed
Frost, D.J. and McCammon, C.A. (2008) The redox state of Earth’s mantle. Annual Review of Earth and Planetary Sciences, 36, 389420, doi:10.1146/annureearth.36.031207.124322.CrossRefGoogle Scholar
Füssel, J., Lam, P., Lavik, G., Jensen, M.M., Holtappels, M., Günter, M. and Kuypers, M.M.M. (2012) Nitrite oxidation in the Namibian oxygen minimum zone. The ISME Journal, 6, 12001209, doi:10.1038/ismej.2011.178.CrossRefGoogle ScholarPubMed
Gaillard, F., Bouhifd, M.A., Füri, E., Malavergne, V., Marrocchi, Y., Noack, L., Ortenzi, G., Roskosz, M. and Vulpius, S. (2021) The diverse planetary ingassing/outgassing paths produced over billions of years of magmatic activity. Space Science Reviews, 217, 22, doi:10.1007/s11214-021-00802-1.CrossRefGoogle Scholar
Gaillard, F., Scaillet, B. and Arndt, N.T. (2011) Atmospheric oxygenation caused by a change in volcanic degassing pressure. Nature, 478, 229232, doi:10.1038/nature10460.CrossRefGoogle ScholarPubMed
Gaillard, F., Scaillet, B., Pichavant, M. and Iacono-Marziano, G. (2015) The redox geodynamics linking basalts and their mantle sources through space and time. Chemical Geology, 418, 217233, doi:10.1016/j.chemgeo.2015.07.030.CrossRefGoogle Scholar
Garcia-Pichel, F., Lombard, J., Soule, T., Dunaj, S., Wu, S.H. and Wojciechowski, M.F. (2019) Timing the evolutionary advent of cyanobacteria and the later Great Oxidation Event using gene phylogenies of a sunscreen. mBio, 10, e00561–19, doi:10.1128/mBio.00561-19.CrossRefGoogle ScholarPubMed
Garvin, J., Buick, R., Anbar, A.D., Arnold, G.L. and Kaufman, A.J. (2009) Isotopic evidence for an aerobic nitrogen cycle in the Latest Archean. Science, 323, 10451048, doi:10.1126/science.1165675.CrossRefGoogle ScholarPubMed
Gessmann, C.K., Rubie, D.C. and McCammon, C.A. (1999) Oxygen fugacity dependence of Ni, Co, Mn, Cr, V, and Si partitioning between liquid metal and magnesiowüstite at 9–18 GPa and 2200°C. Geochimica et Cosmochimica Acta, 63, 18531863, doi:10.1016/S0016-7037(99)00059-9.CrossRefGoogle Scholar
Girard, J., Amulele, G., Farla, R., Mohiuddin, A. and Karato, S. (2016) Shear deformation of bridgmanite and magnesiowüstite aggregates at lower mantle conditions. Science, 351, 144147, doi:10.1126/science.aad3113.CrossRefGoogle ScholarPubMed
Girguis, P.R., Lee, R.W., Desaulniers, N., Childress, J.J., Pospesel, M., Felbeck, H. and Zal, F. (2000) Fate of nitrate acquired by the tubeworm Riftia pachyptila. Applied and Environmental Microbiology, 66, 27832790, doi:10.1128/AEM.66.7.2783-2790.2000.CrossRefGoogle ScholarPubMed
Godfrey, L.V. and Falkowski, P.G. (2009) The cycling and redox state of nitrogen in the Archaean ocean. Nature Geoscience, 2, 725729, doi:10.1038/ngeo633.CrossRefGoogle Scholar
Goldblatt, C., Lenton, T.M. and Watson, A.J. (2006) Bistability of atmospheric oxygen and the Great Oxidation. Nature, 443, 683686, doi:10.1038/nature05169.CrossRefGoogle ScholarPubMed
Grandstaff, D. (1980) Origin of uraniferous conglomerates at Elliot Lake, Canada and Witwatersrand, South Africa: Implications for oxygen in the Precambrian atmosphere. Precambrian Research, 13, 126, doi:10.1016/0301-9268(80)90056-X.CrossRefGoogle Scholar
Grettenberger, C., Gold, D.A. and Brown, C.T. (2025) Distribution of early-branching Cyanobacteria and the potential habitats that gave rise to the earliest oxygenic phototrophs. mSphere, 10, e0101324, doi:10.1128/msphere.01013-24.CrossRefGoogle Scholar
Grettenberger, C.L. and Sumner, D.Y. (2024) Physiology, not nutrient availability, may have limited primary productivity after the emergence of oxygenic photosynthesis. Geobiology, 22, e12622, doi:10.1111/gbi.12622.CrossRefGoogle Scholar
Grewal, D.S., Dasgupta, R., Sun, C., Tsuno, K. and Costin, G. (2019) Delivery of carbon, nitrogen, and sulfur to the silicate Earth by a giant impact. Science Advances, 5, eaau3669, doi:10.1126/sciadaau3669.CrossRefGoogle Scholar
Gribaldo, S., Talla, E. and Brochier-Armanet, C. (2009) Evolution of the haem copper oxidases superfamily: a rooting tale. Trends in Biochemical Sciences, 34, 375381, doi:10.1016/j.tibs.2009.04.002.CrossRefGoogle ScholarPubMed
Gu, T., Li, M., McCammon, C. and Lee, K.K.M. (2016) Redox-induced lower mantle density contrast and effect on mantle structure and primitive oxygen. Nature Geoscience, 9, 723727, doi:10.1038/ngeo2772.CrossRefGoogle Scholar
Guimond, C.M., Noack, L., Ortenzi, G. and Sohl, F. (2021) Low volcanic outgassing rates for a stagnant lid Archean earth with graphite-saturated magmas. Physics of the Earth and Planetary Interiors, 320, 106788, doi:10.1016/j.pepi.2021.106788.CrossRefGoogle Scholar
Gumsley, A.P., Chamberlain, K.R., Bleeker, W., Söderlund, U., de Kock, M.O., Larsson, E.R. and Bekker, A. (2017) Timing and tempo of the Great Oxidation Event. Proceedings of the National Academy of Sciences, 114, 18111816, doi:10.1073/pnas.1608824114.CrossRefGoogle ScholarPubMed
Guo, Q. et al. (2009) Reconstructing Earth’s surface oxidation across the Archean-Proterozoic transition. Geology, 37, 399402, doi:10.1130/G25423A.1.CrossRefGoogle Scholar
Gutzmer, J. and Beukes, N.J. (1996) Mineral paragenesis of the Kalahari manganese field, South Africa. Ore Geology Reviews, 11, 405428.10.1016/S0169-1368(96)00011-XCrossRefGoogle Scholar
Guy, B.M., Ono, S., Gutzmer, J., Kaufman, A.J., Lin, Y., Fogel, M.L. and Beukes, N.J. (2012) A multiple sulfur and organic carbon isotope record from non-conglomeratic sedimentary rocks of the Mesoarchean Witwatersrand Supergroup, South Africa. Precambrian Research, 216–219, 208231, doi:10.1016/j.precamres.2012.06.018.CrossRefGoogle Scholar
Haggerty, S.E. (1978) The redox state of planetary basalts. Geophysical Research Letters, 5, 443446, doi:10.1029/GL005i006p00443.CrossRefGoogle Scholar
Halevy, I. (2013) Production, preservation, and biological processing of mass-independent sulfur isotope fractionation in the Archean surface environment. Proceedings of the National Academy of Sciences, 110, 1764417649, doi:10.1073/pnas.1213148110.CrossRefGoogle ScholarPubMed
Hansel, C.M. (2017) Manganese in marine microbiology. Pp. 3783 in Advances in Microbial Physiology , Vol. 70 (Poole, R. K., editor), Elsevier, doi:10.1016/bs.ampbs.2017.01.005.Google Scholar
Hao, J., Knoll, A.H., Huang, F., Hazen, R.M. and Daniel, I. (2020) Cycling phosphorus on the Archean Earth: Part I. Continental weathering and riverine transport of phosphorus. Geochimica et Cosmochimica Acta, 273, 7084, doi:10.1016/j.gca.2020.01.027.CrossRefGoogle Scholar
Hao, J., Liu, W., Goff, J.L., Steadman, J.A., Large, R.R., Falkowski, P.G. and Yee, N. (2022) Anoxic photochemical weathering of pyrite on Archean continents. Science Advances, 8, eabn2226, doi:10.1126/sciadabn2226.CrossRefGoogle ScholarPubMed
Haqq-Misra, J., Kasting, J.F. and Lee, S. (2011) Availability of O2 and H2O2 on Pre-Photosynthetic Earth. Astrobiology, 11, 293302, doi:10.1089/ast.2010.0572.CrossRefGoogle ScholarPubMed
Harrison, T.M., Schmitt, A.K., McCulloch, M.T. and Lovera, O.M. (2008) Early (≥4.5 Ga) formation of terrestrial crust: Lu–Hf, δ18O, and Ti thermometry results for Hadean zircons. Earth and Planetary Science Letters, 268, 476486, doi:10.1016/j.epsl.2008.02.011.CrossRefGoogle Scholar
Hayes, J.M. (2001) Fractionation of the isotopes of carbon and hydrogen in biosynthetic processes. Reviews in Mineralogy and Geochemistry 43, 225277.10.2138/gsrmg.43.1.225CrossRefGoogle Scholar
Hayes, J.M. (1983) Geochemical evidence bearing on the origin of aerobiosis, a speculative interpretation. Pp. 93134 in: The Earth’s Earliest Biosphere: Its Origin and Evolution (Schopf, J.W., J.W. editor), Princeton University Press, Princeton.Google Scholar
Hayes, J.M. (1994) Global methanotrophy at the Archean-Proterozoic transition. Pp. 220236 in: Early Life on Earth, Nobel Sympsosium (Bengtson, S., editor). Columbia University Press, New York.Google Scholar
Hayes, J.M. and Waldbauer, J.R. (2006) The carbon cycle and associated redox processes through time. Philosophical Transactions of the Royal Society B: Biological Sciences, 361, 931950, doi:10.1098/rstb.2006.1840.CrossRefGoogle ScholarPubMed
Hazen, R.M., Papineau, D., Bleeker, W., Downs, R.T., Ferry, J.M., McCoy, T.J., Sverjensky, D.A. and Yang, H. (2008) Mineral evolution. American Mineralogist, 93, 16931720, doi:10.2138/am.2008.2955.CrossRefGoogle Scholar
He, H. et al. (2023) A mineral-based origin of Earth’s initial hydrogen peroxide and molecular oxygen. Proceedings of the National Academy of Sciences, 120, e2221984120, doi:10.1073/pnas.2221984120.CrossRefGoogle ScholarPubMed
He, H., Wu, X., Xian, H., Zhu, J., Yang, Y., Lv, Y., Li, Y. and Konhauser, K.O. (2021) An abiotic source of Archean hydrogen peroxide and oxygen that pre-dates oxygenic photosynthesis. Nature Communications, 12, 6611, doi:10.1038/s41467-021-26916-2.CrossRefGoogle ScholarPubMed
Head, I.M. (2016) Geomicrobiology of fossil fuels. Pp. 565622 in: Ehrlich’s Geomicrobiology (Ehrlich, H.L., Newman, D.K. and Kappler, A., editors). CRC Press, Boca Raton, pp. 635.Google Scholar
Heard, A.W., Aarons, S.M., Hofmann, A., He, X., Ireland, T., Bekker, A., Qin, L. and Dauphas, N. (2021) Anoxic continental surface weathering recorded by the 2.95 Ga Denny Dalton Paleosol (Pongola Supergroup, South Africa). Geochimica et Cosmochimica Acta, 295, 123, doi:10.1016/j.gca.2020.12.005.CrossRefGoogle Scholar
Heinrichs, T.K. and Reimer, T. (1977) A sedimentary barite deposit from the Archean Fig Tree Group of the Barberton Mountain Land (South Africa). Economic Geology, 72, 14261441.10.2113/gsecongeo.72.8.1426CrossRefGoogle Scholar
Hentschel, U. and Felbeck, H. (1993) Nitrate respiration in the hydrothermal vent tubeworm Riftia pachyptila. Nature, 366, 338340.10.1038/366338a0CrossRefGoogle Scholar
Herd, C.D.K., Borg, L.E., Jones, J.H. and Papike, J.J. (2002) Oxygen fugacity and geochemical variations in the Martian basalts: implications for Martian basalt petrogenesis and the oxidation state of the upper mantle of Mars. Geochimica et Cosmochimica Acta, 66, 20252036, doi:10.1016/S0016-7037(02)00828-1.CrossRefGoogle Scholar
Herd, C.D., Papike, J.J. and Brearley, A.J. (2001) Oxygen fugacity of Martian basalts from electron microprobe oxygen and TEM-EELS analyses of Fe-Ti oxides. American Mineralogist, 86, 10151024.10.2138/am-2001-8-908CrossRefGoogle Scholar
Herrmann, A.J., Sorwat, J., Byrne, J.M., Frankenberg-Dinkel, N. and Gehringer, M.M. (2021) Diurnal Fe(II)/Fe(III) cycling and enhanced O2 production in a simulated Archean marine oxygen oasis. Nature Communications, 12, 2069, doi:10.1038/s41467-021-22258-1.CrossRefGoogle Scholar
Herwartz, D., Pack, A., Krylov, D., Xiao, Y., Muehlenbachs, K., Sengupta, S., and Di Rocco, T. (2015) Revealing the climate of snowball Earth from Δ17O systematics of hydrothermal rocks. Proceedings of the National Academy of Sciences, 112, 53375341.10.1073/pnas.1422887112CrossRefGoogle ScholarPubMed
Herwartz, D., Pack, A. and Nagel, T.J. (2021) A CO2 greenhouse efficiently warmed the early Earth and decreased seawater 18O/16O before the onset of plate tectonics. Proceedings of the National Academy of Sciences, 118, e2023617118, doi:10.1073/pnas.2023617118.CrossRefGoogle Scholar
Herwartz, D. and Viehmann, S. (2024) Does water transparency control the banding in shallow water iron formations? South African Journal of Geology, 127, 379390, doi:10.25131/sajg.127.0014.CrossRefGoogle Scholar
Hinrichs, K.-U. (2002) Microbial fixation of methane carbon at 2.7 Ga: Was an anaerobic mechanism possible? Geochemistry, Geophysics, Geosystems, 3, 110, doi:10.1029/2001GC000286.CrossRefGoogle Scholar
Hoffmann, J.E. (2017) Oxygenation by a changing crust. Nature Geoscience, 10, 713714, doi:10.1038/ngeo3038.CrossRefGoogle Scholar
Hofmann, A., Bekker, A., Rouxel, O., Rumble, D. and Master, S. (2009) Multiple sulphur and iron isotope composition of detrital pyrite in Archaean sedimentary rocks: A new tool for provenance analysis. Earth and Planetary Science Letters, 286, 436445, doi:10.1016/j.epsl.2009.07.008.CrossRefGoogle Scholar
Hofmann, H.J., Grey, K., Hickman, A.H. and Thorpe, R.I. (1999) Origin of 3.45 Ga coniform stromatolites in Warrawoona Group, Western Australia. Geological Society of America Bulletin, 111, 12561262.10.1130/0016-7606(1999)111<1256:OOGCSI>2.3.CO;22.3.CO;2>CrossRefGoogle Scholar
Holm, G. (1989) The 13C/12C ratios of siderite and organic matter of a modern metalliferous hydrothermal sediment and their implications for banded iron formations. Chemical Geology, 77, 4145.10.1016/0009-2541(89)90013-2CrossRefGoogle Scholar
Holland, H.D. (2006) The oxygenation of the atmosphere and oceans. Philosophical Transactions of the Royal Society B: Biological Sciences, 361, 903915, doi:10.1098/rstb.2006.1838.CrossRefGoogle ScholarPubMed
Holland, H.D. (2002) Volcanic gases, black smokers, and the great oxidation event. Geochimica et Cosmochimica Acta, 66, 38113826, doi:10.1016/S0016-7037(02)00950-X.CrossRefGoogle Scholar
Holloway, J.R. and Blank, J.G. (1994) Application of experimental results to C-O-H species in natural melts. Pp. 187230 in: Volatiles in Magmas (Carroll, M.R. and Holloway, J.R., editors). De Gruyter, doi:10.1515/9781501509674-012.CrossRefGoogle Scholar
Holloway, J.R., Pan, V. and Gudmundsson, G. (1992) High-pressure fluid-absent melting experiments in the presence of graphite: oxygen fugacity, ferric/ferrous ratio and dissolved CO2. European Journal of Mineralogy, 4, 105114.10.1127/ejm/4/1/0105CrossRefGoogle Scholar
Homann, M. (2019) Earliest life on Earth: Evidence from the Barberton Greenstone Belt, South Africa. Earth-Science Reviews, 196, 102888, doi:10.1016/j.earscire2019.102888.CrossRefGoogle Scholar
Homann, M., Heubeck, C., Airo, A. and Tice, M.M. (2015) Morphological adaptations of 3.22 Ga-old tufted microbial mats to Archean coastal habitats (Moodies Group, Barberton Greenstone Belt, South Africa). Precambrian Research, 266, 4764, doi:10.1016/j.precamres.2015.04.018.CrossRefGoogle Scholar
Hopkins, M., Harrison, T.M. and Manning, C.E. (2008) Low heat flow inferred from >4 Gyr zircons suggests Hadean plate boundary interactions. Nature, 456, 493496, doi:10.1038/nature07465.CrossRefGoogle ScholarPubMed
Horan, M.F., Carlson, R.W., Walker, R.J., Jackson, M., Garçon, M. and Norman, M. (2018) Tracking Hadean processes in modern basalts with 142-Neodymium. Earth and Planetary Science Letters, 484, 184191, doi:10.1016/j.epsl.2017.12.017.CrossRefGoogle Scholar
Hu, G., Rumble, D., and Wang, P. (2003) An ultraviolet laser microprobe for the in situ analysis of multisulfur isotopes and its use in measuring Archean sulfur isotope mass-independent anomalies. Geochimica et Cosmochimica Acta, 67, 31013118, doi:10.1016/S0016-7037(02)00929-8.CrossRefGoogle Scholar
Huston, D.L. and Logan, G.A. (2004) Barite, BIFs and bugs: evidence for the evolution of the Earth’s early hydrosphere. Earth and Planetary Science Letters, 220, 4155, doi:10.1016/S0012-821X(04)00034-2.CrossRefGoogle Scholar
Ingalls, M., Grotzinger, J.P., Presen, T., Rasmussen, B. and Fischer, W.W. (2022) Carbonate-associated phosphate (CAP) indicates elevated phosphate availability in Neoarchean shallow marine environments. Geophysical Research Letters, 49, e2022GL098100, doi:10.1029/2022GL098100.CrossRefGoogle Scholar
Isson, T. and Rauzi, S. (2024) Oxygen isotope ensemble reveals Earth’s seawater, temperature, and carbon cycle history. Science, 383, 666670, doi:10.1126/science.adg1366.CrossRefGoogle ScholarPubMed
Jabłońska, J. and Tawfik, D.S. (2022) Innovation and tinkering in the evolution of oxidases. Protein Science, 31, e4310, doi:10.1002/pro.4310.CrossRefGoogle ScholarPubMed
Jabłońska, J. and Tawfik, D.S. (2021) The evolution of oxygen-utilizing enzymes suggests early biosphere oxygenation. Nature Ecology & Evolution, 5, 442448, doi:10.1038/s41559-020-01386-9.CrossRefGoogle ScholarPubMed
Jamieson, J.W., Wing, B.A., Hannington, M.D. and Farquhar, J. (2006) Evaluating isotopic equilibrium among sulfide mineral pairs in Archean ore deposits: case study from the Kidd Creek VMS deposit, Ontario, Canada. Economic Geology, 101, 10551061.10.2113/gsecongeo.101.5.1055CrossRefGoogle Scholar
Javoy, M. (1995) The integral enstatite chondrite model of the Earth. Geophysical Research Letters, 22, 22192222, doi:10.1029/95GL02015.CrossRefGoogle Scholar
Johnson, J.E., Gerpheide, A., Lamb, M.P. and Fischer, W.W. (2014) O2 constraints from Paleoproterozoic detrital pyrite and uraninite. GSA Bulletin, 126, 813830.10.1130/B30949.1CrossRefGoogle Scholar
Johnson, J.E., Webb, S.M., Thomas, K., Ono, S., Kirschvink, J.L. and Fischer, W.W. (2013) Manganese-oxidizing photosynthesis before the rise of cyanobacteria. Proceedings of the National Academy of Sciences, 110, 1123811243, doi:10.1073/pnas.1305530110.CrossRefGoogle ScholarPubMed
Johnston, D.T., Farquhar, J., Summons, R.E., Shen, Y., Kaufman, A.J., Masterson, A.L., and Canfield, D.E. (2008) Sulfur isotope biogeochemistry of the Proterozoic McArthur Basin. Geochimica et Cosmochimica Acta, 72, 42784290, doi:10.1016/j.gca.2008.06.004.CrossRefGoogle Scholar
Johnston, D.T., Poulton, S.W., Fralick, P.W., Wing, B.A., Canfield, D.E. and Farquhar, J. (2006) Evolution of the oceanic sulfur cycle at the end of the Paleoproterozoic. Geochimica et Cosmochimica Acta, 70, 57235739, doi:10.1016/j.gca.2006.08.001.CrossRefGoogle Scholar
Jones, C., Crowe, S.A., Sturm, A., Leslie, K.L., MacLean, L.C.W., Katsev, S., Henny, C., Fowle, D.A. and Canfield, D.E. (2011) Biogeochemistry of manganese in ferruginous Lake Matano, Indonesia. Biogeosciences, 8, 29772991, doi:10.5194/bg-8-2977-2011.CrossRefGoogle Scholar
Jones, C., Nomosatryo, S., Crowe, S.A., Bjerrum, C.J. and Canfield, D.E. (2015) Iron oxides, divalent cations, silica, and the early earth phosphorus crisis. Geology, 43, 135138, doi:10.1130/G36044.1.CrossRefGoogle Scholar
Jørgensen, B.B. and Gallardo, A., 1999, Thioploca sp: filamentous sulfur bacteria with nitrate vacuoles. FEMS Microbiology Ecology, 28, 301313, doi:10.1111/j.1574-6941.1999.tb00585.x.CrossRefGoogle Scholar
Kalvelage, T. et al. (2015) Aerobic microbial respiration in oceanic oxygen minimum zones. PLoS ONE, 10, e0133526, doi:10.1371/journal.pone.0133526.CrossRefGoogle ScholarPubMed
Kalvelage, T., Jensen, M.M., Contreras, S., Revsbech, N.P., Lam, P., Günter, M., LaRoche, J., Lavik, G. and Kuypers, M.M.M. (2011) Oxygen sensitivity of anammox and coupled N-cycle processes in oxygen minimum zones. PLoS ONE, 6, e29299, doi:10.1371/journal.pone.0029299.CrossRefGoogle ScholarPubMed
Kamber, B.S. and Whitehouse, M.J. (2007) Micro-scale sulphur isotope evidence for sulphur cycling in the late Archean shallow ocean. Geobiology, 5, 517.10.1111/j.1472-4669.2006.00091.xCrossRefGoogle ScholarPubMed
Kappler, A., Bryce, C., Mansor, M., Lueder, U., Byrne, J.M. and Swanner, E.D. (2021) An evolving view on biogeochemical cycling of iron. Nature Reviews Microbiology, 19, 360374, doi:10.1038/s41579-020-00502-7.CrossRefGoogle Scholar
Kappler, A. and Newman, D.K. (2004) Formation of Fe(III)-minerals by Fe(II)-oxidizing photoautotrophic bacteria. Geochimica et Cosmochimica Acta, 68, 12171226, doi:10.1016/j.gca.2003.09.006.CrossRefGoogle Scholar
Kappler, A., Pasquero, C., Konhauser, K.O. and Newman, D.K. (2005) Deposition of banded iron formations by anoxygenic phototrophic Fe(II)-oxidizing bacteria. Geology, 33, 865868, doi:10.1130/G21658.1.CrossRefGoogle Scholar
Kasting, J.F. (1993) Earth’s early atmosphere. Science, 259, 920926.10.1126/science.11536547CrossRefGoogle ScholarPubMed
Kasting, J.F. (1987) Theoretical constraints on oxygen and carbon dioxide concentrations in the Precambrian atmosphere. Precambrian Research, 34, 205229, doi:10.1016/0301-9268(87)90001-5.CrossRefGoogle ScholarPubMed
Kasting, J.F. and Canfield, D.E. (2012) The global oxygen cycle. Pp. 93104 in Fundamentals of Geobiology (Knoll, A., Canfield, D. and Konhauser, K.O., editors). Wiley-Blackwell, West-Sussex.10.1002/9781118280874.ch7CrossRefGoogle Scholar
Kasting, J.F., Eggler, D.H. and Raeburn, S.P. (1993) Mantle redox evolution and the oxidation state of the Archean atmosphere. The Journal of Geology, 101, 245257, doi:10.1086/648219.CrossRefGoogle ScholarPubMed
Kasting, J.F., Howard, M.T., Wallmann, K., Veizer, J., Shields, G. and Jaffrés, J. (2006) Paleoclimates, ocean depth, and the oxygen isotopic composition of seawater. Earth and Planetary Science Letters, 252, 8293, doi:10.1016/j.epsl.2006.09.029.CrossRefGoogle Scholar
Kasting, J.F. and Siefert, J.L. (2002) Life and the evolution of Earth’s atmosphere. Science, 296, 10661068, doi:10.1126/science.1071184.CrossRefGoogle ScholarPubMed
Kaufman, A.J., Johnston, D.T., Farquhar, J., Masterson, A.L., Lyons, T.W., Bates, S., Anbar, A.D., Arnold, G.L., Garvin, J. and Buick, R. (2007) Late Archean biospheric oxygenation and atmospheric evolution. Science, 317, 19001903, doi:10.1126/science.1138700.CrossRefGoogle ScholarPubMed
Kemp, A.I.S. and Hawkesworth, C.J. (2014) Growth and differentiation of the continental crust from isotope studies of accessory minerals. Pp. 379421 in Treatise on Geochemistry, 2nd edition. Elsevier, doi:10.1016/B978-0-08-095975-7.00312-0.CrossRefGoogle Scholar
Kendall, B. (2021) Recent advances in geochemical paleo-oxybarometers. Annual Review of Earth and Planetary Sciences, 49, annurev-earth-071520-051637, doi:10.1146/annurev-earth-071520-051637.CrossRefGoogle Scholar
Kendall, B., Creaser, R.A., Reinhard, C.T., Lyons, T.W. and Anbar, A.D. (2015) Transient episodes of mild environmental oxygenation and oxidative continental weathering during the late Archean. Science Advances, 1, e1500777, doi:10.1126/sciad1500777.CrossRefGoogle ScholarPubMed
Kharecha, P., Kasting, J. and Siefert, J. (2005) A coupled atmosphere-ecosystem model of the early Archean Earth. Geobiology, 3, 5376, doi:10.1111/j.1472-4669.2005.00049.x.CrossRefGoogle Scholar
Kim, K.M., Qin, T., Jiang, Y.-Y., Chen, L.-L., Xiong, M., Caetano-Anollés, D., Zhang, H.-Y. and Caetano-Anollés, G. (2012) Protein domain structure uncovers the origin of aerobic metabolism and the rise of planetary oxygen. Structure, 20, 6776, doi:10.1016/j.str.2011.11.003.CrossRefGoogle ScholarPubMed
Kipp, M.A. and Stüeken, E.E. (2017) Biomass recycling and Earth’s early phosphorus cycle. Science Advances, 3, eaao4795, doi:10.1126/sciadaao4795.CrossRefGoogle ScholarPubMed
Kirschvink, J.L., Gaidos, E.J., Bertani, L.E., Beukes, N.J., Gutzmer, J., Maepa, L.N. and Steinberger, R.E. (2000) Paleoproterozoic snowball Earth: Extreme climatic and geochemical global change and its biological consequences. Proceedings of the National Academy of Sciences, 97, 14001405, doi:10.1073/pnas.97.4.1400.CrossRefGoogle ScholarPubMed
Klatt, J.M., Chennu, A., Arbic, B.K., Biddanda, B.A. and Dick, G.J. (2021) Possible link between Earth’s rotation rate and oxygenation. Nature Geoscience, 14, 564570, doi:10.1038/s41561-021-00784-3.CrossRefGoogle Scholar
Klein, C. and Beukes, N.J. (1992) Proterozoic iron formations. Developments in Precambrian Geology, 10, 383418.10.1016/S0166-2635(08)70124-5CrossRefGoogle Scholar
Knittel, K. and Boetius, A. (2009) Anaerobic oxidation of methane: progress with an unknown process. Annual Review of Microbiology, 63, 311334, doi:10.1146/annuremicro.61.080706.093130.CrossRefGoogle ScholarPubMed
Knoll, A.H., Bergmann, K.D. and Strauss, J.V. (2016) Life: the first two billion years. Philosophical Transactions of the Royal Society B: Biological Sciences, 371, 20150493, doi:10.1098/rstb.2015.0493.CrossRefGoogle ScholarPubMed
Koehler, M.C., Buick, R., Kipp, M.A., Stüeken, E.E. and Zaloumis, J. (2018) Transient surface ocean oxygenation recorded in the ∼2.66-Ga Jeerinah Formation, Australia. Proceedings of the National Academy of Sciences, 115, 77117716, doi:10.1073/pnas.1720820115.CrossRefGoogle ScholarPubMed
Konhauser, K.O. et al. (2011) Aerobic bacterial pyrite oxidation and acid rock drainage during the Great Oxidation Event. Nature, 478, 369373, doi:10.1038/nature10511.CrossRefGoogle ScholarPubMed
Konhauser, K. (2007) Introduction to Geomicrobiology. Blackwell Pub, Malden, 425 pp.Google Scholar
Konhauser, K.O. et al. (2017) Iron formations: A global record of Neoarchaean to Palaeoproterozoic environmental history. Earth-Science Reviews, 172, 140177, doi:10.1016/j.earscire2017.06.012.CrossRefGoogle Scholar
Konhauser, K.O. et al. (2018) Phytoplankton contributions to the trace-element composition of Precambrian banded iron formations. GSA Bulletin, 130, 941951, doi:10.1130/B31648.1.CrossRefGoogle Scholar
Konhauser, K.O., Amskold, L., Lalonde, S.V., Posth, N.R., Kappler, A. and Anbar, A. (2007a) Decoupling photochemical Fe(II) oxidation from shallow-water BIF deposition. Earth and Planetary Science Letters, 258, 87100, doi:10.1016/j.epsl.2007.03.026.CrossRefGoogle Scholar
Konhauser, K.O., Hamade, T., Raiswell, R., Morris, R.C., Ferris, F.G., Southam, G. and Canfield, D.E. (2002) Could bacteria have formed the Precambrian banded iron formations? Geology, 30 (12), 10791082.10.1130/0091-7613(2002)030<1079:CBHFTP>2.0.CO;22.0.CO;2>CrossRefGoogle Scholar
Konhauser, K.O., Lalonde, S.V., Amskold, L. and Holland, H.D. (2007b) Was there really an Archean phosphate crisis? Science, 315, 12341234, doi:10.1126/science.1136328.CrossRefGoogle Scholar
Konhauser, K.O., Pecoits, E., Lalonde, S.V., Papineau, D., Nisbet, E.G., Barley, M.E., Arndt, N.T., Zahnle, K. and Kamber, B.S. (2009) Oceanic nickel depletion and a methanogen famine before the Great Oxidation Event. Nature, 458, 750753, doi:10.1038/nature07858.CrossRefGoogle Scholar
Kopp, R.E., Kirschvink, J.L., Hilburn, I.A. and Nash, C.Z. (2005) The Paleoproterozoic snowball Earth: A climate disaster triggered by the evolution of oxygenic photosynthesis. Proceedings of the National Academy of Sciences, 102, 1113111136, doi:10.1073/pnas.0504878102.CrossRefGoogle ScholarPubMed
Korenaga, J. (2018) Estimating the formation age distribution of continental crust by unmixing zircon ages. Earth and Planetary Science Letters, 482, 388395, doi:10.1016/j.epsl.2017.11.039.CrossRefGoogle Scholar
Kuang, J., Morra, G., Yuen, D.A., Kusky, T., Jiang, S., Yao, H. and Qi, S. (2023) Metamorphic constraints on Archean tectonics. Precambrian Research, 397, 107195, doi:10.1016/j.precamres.2023.107195.CrossRefGoogle Scholar
Kump, L.R., and Barley, M.E. (2007) Increased subaerial volcanism and the rise of atmospheric oxygen 2.5 billion years ago. Nature, 448, 10331036, doi:10.1038/nature06058.CrossRefGoogle ScholarPubMed
Kump, L.R., Kasting, J.F. and Barley, M.E. (2001) Rise of atmospheric oxygen and the upside-down Archean mantle. Geochemistry, Geophysics, Geosystems, 2, 2000GC000114.10.1029/2000GC000114CrossRefGoogle Scholar
Kurzweil, F., Wille, M., Gantert, N., Beukes, N.J. and Schoenberg, R. (2016) Manganese oxide shuttling in pre-GOE oceans – evidence from molybdenum and iron isotopes. Earth and Planetary Science Letters, 452, 6978, doi:10.1016/j.epsl.2016.07.013.CrossRefGoogle Scholar
Laakso, T.A. and Schrag, D.P. (2014) Regulation of atmospheric oxygen during the Proterozoic. Earth and Planetary Science Letters, 388, 8191, doi:10.1016/j.epsl.2013.11.049.CrossRefGoogle Scholar
Lalonde, S.V. and Konhauser, K.O. (2015) Benthic perspective on Earth’s oldest evidence for oxygenic photosynthesis. Proceedings of the National Academy of Sciences, 112, 9951000, doi:10.1073/pnas.1415718112.CrossRefGoogle ScholarPubMed
Lam, P., Lavik, G., Jensen, M.M., Van De Vossenberg, J., Schmid, M., Woebken, D., Gutiérrez, D., Amann, R., Jetten, M.S.M. and Kuypers, M.M.M. (2009) Revising the nitrogen cycle in the Peruvian oxygen minimum zone. Proceedings of the National Academy of Sciences, 106, 47524757, doi:10.1073/pnas.0812444106.CrossRefGoogle ScholarPubMed
Lammer, H. et al. (2018) Origin and evolution of the atmospheres of early Venus, Earth and Mars. The Astronomy and Astrophysics Review, 26, 2, doi:10.1007/s00159-018-0108-y.CrossRefGoogle Scholar
Latysheva, N., Junker, L., Palmer, W.J., Codd, G.A. and Barker, D. (2012) The evolution of nitrogen fixation in cyanobacteria. Bioinformatics, 28, 603606, doi:10.1093/bioinformatics/bts008.CrossRefGoogle ScholarPubMed
Learman, D.R., Voelker, B.M., Vazquez-Rodriguez, A.I. and Hansel, C.M. (2011) Formation of manganese oxides by bacterially generated superoxide. Nature Geoscience, 4, 9598, doi:10.1038/ngeo1055.CrossRefGoogle Scholar
Lee, C.-T.A., Brandon, A.D. and Norman, M. (2003) Vanadium in peridotites as a proxy for paleo-fO2 during partial melting: Prospects, limitations, and implications. Geochimica et Cosmochimica Acta, 67, 30453064.10.1016/S0016-7037(03)00268-0CrossRefGoogle Scholar
Lee, C.-T.A., Yeung, L.Y., McKenzie, N.R., Yokoyama, Y., Ozaki, K. and Lenardic, A. (2016) Two-step rise of atmospheric oxygen linked to the growth of continents. Nature Geoscience, 9, 417424, doi:10.1038/ngeo2707.CrossRefGoogle Scholar
Lee, S.H., Youn, H., Kang, S.G. and Lee, H.S. (2019) Oxygen-mediated growth enhancement of an obligate anaerobic archaeon Thermococcus onnurineus NA1. Journal of Microbiology, 57, 138142, doi:10.1007/s12275-019-8592-y.CrossRefGoogle ScholarPubMed
Lepot, K. (2020) Signatures of early microbial life from the Archean (4 to 2.5 Ga) eon. Earth-Science Reviews, 209, 103296, doi:10.1016/j.earscire2020.103296.CrossRefGoogle Scholar
Lepot, K., Williford, K.H., Philippo, P., Thomazo, C., Ushikubo, T., Kitajima, K., Mostefaoui, S. and Valley, J.W. (2019) Extreme 13C-depletions and organic sulfur content argue for S-fueled anaerobic methane oxidation in 2.72 Ga old stromatolites. Geochimica et Cosmochimica Acta, 244, 522547, doi:10.1016/j.gca.2018.10.014.CrossRefGoogle Scholar
Li, Y.-L., Konhauser, K.O. and Zhai, M. (2017) The formation of magnetite in the early Archean oceans. Earth and Planetary Science Letters, 466, 103114, doi:10.1016/j.epsl.2017.03.013.CrossRefGoogle Scholar
Li, Z.-X.A. and Lee, C.-T.A. (2004) The constancy of upper mantle fO2 through time inferred from V/Sc ratios in basalts. Earth and Planetary Science Letters, 228, 483493, doi:10.1016/j.epsl.2004.10.006.Google Scholar
Little, C.T.S. et al. (2021) A late Paleoproterozoic (1.74 Ga) deep-sea, low-temperature, iron-oxidizing microbial hydrothermal vent community from Arizona, USA. Geobiology, 19, 228249, doi:10.1111/gbi.12434.CrossRefGoogle ScholarPubMed
Liu, W., Hao, J., Elzinga, E.J., Piotrowiak, P., Nanda, V., Yee, N. and Falkowski, P.G. (2020) Anoxic photogeochemical oxidation of manganese carbonate yields manganese oxide. Proceedings of the National Academy of Sciences, 117, 2269822704, doi:10.1073/pnas.2002175117.CrossRefGoogle ScholarPubMed
Loucks, R.R., Fiorentini, M.L. and Henríquez, G.J. (2020) New magmatic oxybarometer using trace elements in zircon. Journal of Petrology, 61, egaa034, doi:10.1093/petrology/egaa034.CrossRefGoogle Scholar
Lowe, D.R., Drabon, N. and Byerly, G.R. (2019) Crustal fracturing, unconformities, and barite deposition, 3.26–3.23 Ga, Barberton Greenstone Belt, South Africa. Precambrian Research, 327, 3446, doi:10.1016/j.precamres.2019.02.024.CrossRefGoogle Scholar
Luo, G., Ono, S., Beukes, N.J., Wang, D.T., Xie, S. and Summons, R.E. (2016) Rapid oxygenation of Earth’s atmosphere 2.33 billion years ago. Science Advances, 2, e1600134, doi:10.1126/sciad1600134.CrossRefGoogle ScholarPubMed
Luther, G.W., Findlay, A.J., MacDonald, D.J., Owings, S.M., Hanson, T.E., Beinart, R.A. and Girguis, P.R. (2011) Thermodynamics and kinetics of sulfide oxidation by oxygen: A look at inorganically controlled reactions and biologically mediated processes in the environment. Frontiers in Microbiology, 2, doi:10.3389/fmicb.2011.00062.CrossRefGoogle Scholar
Luther, G.W., Sundby, B., Lewis, B.L., Brendel, P.J. and Silverberg, N. (1997) Interactions of manganese with the nitrogen cycle. Alternative pathways to dinitrogen. Geochimica et Cosmochimica Acta, 61, 40434052, doi:10.1016/S0016-7037(97)00239-1.CrossRefGoogle Scholar
Lyons, T.W., Diamond, C.W., Planavsky, N.J., Reinhard, C.T. and Li, C. (2021) Oxygenation, life, and the planetary system during Earth’s middle history: An overview. Astrobiology, 21, 906923, doi:10.1089/ast.2020.2418.CrossRefGoogle ScholarPubMed
Lyons, T.W., Reinhard, C.T. and Planavsky, N.J. (2014) The rise of oxygen in Earth’s early ocean and atmosphere. Nature, 506, 307315, doi:10.1038/nature13068.CrossRefGoogle ScholarPubMed
Lyons, T.W., Tino, C.J., Fournier, G.P., Anderson, R.E., Leavitt, W.D., Konhauser, K.O. and Stüeken, E.E. (2024) Co-evolution of early Earth environments and microbial life. Nature Reviews Microbiology, doi:10.1038/s41579-024-01044-y.CrossRefGoogle Scholar
Maisch, M., Lueder, U., Laufer, K., Scholze, C., Kappler, A. and Schmidt, C. (2019) Contribution of microaerophilic iron(II)-oxidizers to iron(III) mineral formation. Environmental Science & Technology, 53, 81978204, doi:10.1021/acs.est.9b01531.CrossRefGoogle ScholarPubMed
Mänd, K., Robbins, L.J., Planavsky, N.J., Bekke, A. and Konhauser, K.O. (2021) iron formations as palaeoenvironmental archives in: Elements in Geochemical Tracers in Earth System Science. Cambridge University Press.Google Scholar
Mandernack, K.W., Mills, C.T., Johnson, C.A., Rahn, T. and Kinney, C. (2009) The δ15N and δ18O values of N2O produced during the co-oxidation of ammonia by methanotrophic bacteria. Chemical Geology, 267, 96107, doi:10.1016/j.chemgeo.2009.06.008.CrossRefGoogle Scholar
Mao, H.K. and Bell, P.M. (1977) Disproportionation equilibrium in iron-bearing systems at pressures above 100 kbar with applications to chemistry of the Earth’s mantle. Pp. 236249 in: Energetics of Geological Processes: Hans Ramberg on his 60th birthday (Saxena, S.K., Bhattacharji, S., Annersten, H. and Stephansson, O., editors). Springer, Berlin Heidelberg, doi:10.1007/978-3-642-86574-9_12.CrossRefGoogle Scholar
Mariotti, A., Germon, J.C., Hubert, P., Kaiser, P., Letolle, R., Tardieux, A. and Tardieux, P. (1981) Experimental determination of nitrogen kinetic isotope fractionation: Some principles; illustration for the denitrification and nitrification processes. Plant and Soil, 62, 413430, doi:10.1007/BF02374138.CrossRefGoogle Scholar
Martin, A.N., Stüeken, E.E., Michaud, J.A.-S., Münker, C., Weyer, S., Van Hees, E.H.P. and Gehringer, M.M. (2024) Mechanisms of nitrogen isotope fractionation at an ancient black smoker in the 2.7 Ga Abitibi greenstone belt, Canada. Geology, 52, 181186, doi:10.1130/G51689.1.CrossRefGoogle Scholar
Marty, B. (2012) The origins and concentrations of water, carbon, nitrogen and noble gases on Earth. Earth and Planetary Science Letters, 313–314, 5666, doi:10.1016/j.epsl.2011.10.040.CrossRefGoogle Scholar
McAllister, S.M., Moore, R.M., Gartman, A., Luther, G.W., Emerson, D. and Chan, C.S. (2019) The Fe(II)-oxidizing Zetaproteobacteria : historical, ecological and genomic perspectives. FEMS Microbiology Ecology, 95, doi:10.1093/femsec/fiz015.CrossRefGoogle ScholarPubMed
McCammon, C. (2005) The paradox of mantle redox. Science, 308, 807808, doi:10.1126/science.1110532.CrossRefGoogle ScholarPubMed
Meadows, S. et al. (2018) Exoplanet biosignatures: Understanding oxygen as a biosignature in the context of its environment. Astrobiology, 18, 630662, doi:10.1089/ast.2017.1727.CrossRefGoogle ScholarPubMed
Mhlanga, X.R., Tsikos, H., Lee, B., Rouxel, O.J., Boyce, A.C., Harris, C. and Lyons, T.W. (2023) The Palaeoproterozoic Hotazel BIF-Mn formation as an archive of Earth’s earliest oxygenation. Earth-Science Reviews, 240, 104389, doi:10.1016/j.earscire2023.104389.CrossRefGoogle Scholar
Mikhail, S. and Sverjensky, D.A. (2014) Nitrogen speciation in upper mantle fluids and the origin of Earth’s nitrogen-rich atmosphere. Nature Geoscience, 7, 816819, doi:10.1038/ngeo2271.CrossRefGoogle Scholar
Mills, D.B., Boyle, R.A., Daines, S.J., Sperling, E.A., Pisani, D., Donoghue, P.C.J. and Lenton, T.M. (2022) Eukaryogenesis and oxygen in Earth history. Nature Ecology & Evolution, 6, 520532, doi:10.1038/s41559-022-01733-y.CrossRefGoogle ScholarPubMed
Milucka, J., Kirf, M., Lu, L., Krupke, A., Lam, P., Littmann, S., Kuypers, M.M. and Schubert, C.J. (2015) Methane oxidation coupled to oxygenic photosynthesis in anoxic waters. The ISME Journal, 9, 19912002, doi:10.1038/ismej.2015.12.CrossRefGoogle ScholarPubMed
Mloszewska, A.M., Cole, D.B., Planavsky, N.J., Kappler, A., Whitford, D.S., Owttrim, G.W. and Konhauser, K.O. (2018) UV radiation limited the expansion of cyanobacteria in early marine photic environments. Nature Communications, 9, 3088, doi:10.1038/s41467-018-05520-x.CrossRefGoogle ScholarPubMed
Mojzsis, S.J., Arrhenius, G., McKeegan, K.D., Harrison, T.M., Nutman, A.P, and Friend, C.R.L. (1996) Evidence for life on Earth before 3,800 million years ago. Nature, 384, 5559, doi:10.1038/384055a0.CrossRefGoogle ScholarPubMed
Mojzsis, S.J., Coath, C.D., Greenwood, J.P., McKeegan, K.D. and Harrison, T.M. (2003) Mass-independent isotope effects in Archean (2.5 to 3.8 Ga) sedimentary sulfides determined by ion microprobe analysis. Geochimica et Cosmochimica Acta, 67, 16351658, doi:10.1016/S0016-7037(03)00059-0.CrossRefGoogle Scholar
Muhling, J.R. and Rasmussen, B. (2020) Widespread deposition of greenalite to form Banded Iron Formations before the Great Oxidation Event. Precambrian Research, 339, 105619, doi:10.1016/j.precamres.2020.105619.CrossRefGoogle Scholar
Mukhopadhyay, S. (2012) Early differentiation and volatile accretion recorded in deep-mantle neon and xenon. Nature, 486, 101104, doi:10.1038/nature11141.CrossRefGoogle ScholarPubMed
Mulder, A., Graaf, A.A., Robertson, L.A. and Kuenen, J.G. (1995) Anaerobic ammonium oxidation discovered in a denitrifying fluidized bed reactor. FEMS Microbiology Ecology, 16, 177184, doi:10.1111/j.1574-6941.1995.tb00281.x.CrossRefGoogle Scholar
Mundl, A., Touboul, M., Jackson, M.G., Day, J.M.D., Kurz, M.D., Lekic, V., Helz, R.T. and Walker, R.J. (2017) Tungsten-182 heterogeneity in modern ocean island basalts. Science, 356, 6669, doi:10.1126/science.aal4179.CrossRefGoogle ScholarPubMed
Mundl-Petermeier, A., Viehmann, S., Tusch, J., Bau, M., Kurzweil, F. and Münker, C. (2022) Earth’s geodynamic evolution constrained by 182W in Archean seawater. Nature Communications, 13, 2701, doi:10.1038/s41467-022-30423-3.CrossRefGoogle ScholarPubMed
Nabhan, S., Marin-Carbonne, J., Mason, P.R.D. and Heubeck, C. (2020) 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, 18, 426444, doi:10.1111/gbi.12393.CrossRefGoogle ScholarPubMed
Navarro-González, R., Molina, M.J. and Molina, L.T. (1998) Nitrogen fixation by volcanic lightning in the early Earth. Geophysical Research Letters, 25, 31233126.10.1029/98GL02423CrossRefGoogle Scholar
Nealson, K.H., Tebo, B.M. and Rosson, R.A. (1988) Occurrence and mechanisms of microbial oxidation of manganese. Pp. 279318 in: Advances in Applied Microbiology, Vol. 33, Elsevier, doi:10.1016/S0065-2164(08)70209-0.Google Scholar
Nicklas, R.W., Puchtel, I.S. and Ash, R.D. (2016) High-precision determination of the oxidation state of komatiite lavas using vanadium liquid-mineral partitioning. Chemical Geology, 433, 3645, doi:10.1016/j.chemgeo.2016.04.011.CrossRefGoogle Scholar
Nicklas, R.W., Puchtel, I.S. and Ash, R.D. (2018) Redox state of the Archean mantle: Evidence from V partitioning in 3.5–2.4 Ga komatiites. Geochimica et Cosmochimica Acta, 222, 447466, doi:10.1016/j.gca.2017.11.002.CrossRefGoogle Scholar
Nicklas, R.W., Puchtel, I.S., Ash, R.D., Piccoli, P.M., Hanski, E., Nisbet, E.G., Waterton, P., Pearson, D.G. and Anbar, A.D. (2019) Secular mantle oxidation across the Archean-Proterozoic boundary: Evidence from V partitioning in komatiites and picrites. Geochimica et Cosmochimica Acta, 250, 4975, doi:10.1016/j.gca.2019.01.037.CrossRefGoogle Scholar
Nims, C., Lafond, J., Alleon, J., Templeton, A.S. and Cosmidis, J. (2021) Organic biomorphs may be better preserved than microorganisms in early Earth sediments. Geology, 49, 629634, doi:10.1130/G48152.1.CrossRefGoogle Scholar
Ohmoto, H., Watanabe, Y., Ikemi, H., Poulson, S.R. and Taylor, B.E. (2006a) Sulphur isotope evidence for an oxic Archaean atmosphere. Nature, 442, 908911.10.1038/nature05044CrossRefGoogle Scholar
Ohmoto, H., Watanabe, Y., Yamaguchi, K.E., Naraoka, H., Haruna, M., Kakegawa, T., Hayashi, K. and Kato, Y. (2006b) Chemical and biological evolution of early Earth: Constraints from banded iron formations in: Evolution of Early Earth’s Atmosphere, Hydrosphere, and Biosphere – Constraints from Ore Deposits. Geological Society of America, doi:10.1130/2006.1198(17).Google Scholar
Olson, H.C., Drabon, N. and Johnston, D.T. (2022) Oxygen isotope insights into the Archean ocean and atmosphere. Earth and Planetary Science Letters, 591, 117603, doi:10.1016/j.epsl.2022.117603.CrossRefGoogle Scholar
St.C, O’Neill H. (1991) The origin of the moon and the early history of the earth—A chemical model. Part 2: The earth. Geochimica et Cosmochimica Acta, 55, 11591172, doi:10.1016/0016-7037(91)90169-6.Google Scholar
O’Neill, C. and Aulbach, S. (2022) Destabilization of deep oxidized mantle drove the Great Oxidation Event. Science Advances, 8, eabg1626, doi:10.1126/sciadabg1626.CrossRefGoogle ScholarPubMed
St.C, O’Neill H. and Wall, J. (1987) The olivine–orthopyroxene–spinel oxygen geobarometer, the nickel precipitation curve, and the oxygen fugacity of the Earth’s upper mantle. Journal of Petrology, 28, 11691191, doi:10.1093/petrology/28.6.1169.Google Scholar
Ono, S. (2017) Photochemistry of sulfur dioxide and the origin of mass-independent isotope fractionation in earth’s atmosphere. Annual Review of Earth and Planetary Sciences, 45, 301329, doi:10.1146/annurev-earth-060115-012324.CrossRefGoogle Scholar
Ono, S., Beukes, N.J. and Rumble, D. (2009) Origin of two distinct multiple-sulfur isotope compositions of pyrite in the 2.5Ga Klein Naute Formation, Griqualand West Basin, South Africa. Precambrian Research, 169, 4857, doi:10.1016/j.precamres.2008.10.012.CrossRefGoogle Scholar
Ono, S., Beukes, N.J., Rumble, D. and Fogel, M.L. (2006) 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. South African Journal of Geology, 109, 97108.10.2113/gssajg.109.1-2.97CrossRefGoogle Scholar
Ono, S., Eigenbrode, J.L., Pavlov, A.A., Kharecha, P., Rumble, D., Kasting, J.F. and Freeman, K.H., (2003) New insights into Archean sulfur cycle from mass-independent sulfur isotope records from the Hamersley Basin, Australia. Earth and Planetary Science Letters, 213, 1530, doi:10.1016/S0012-821X(03)00295-4.CrossRefGoogle Scholar
Ortenzi, G. et al. (2020) Mantle redox state drives outgassing chemistry and atmospheric composition of rocky planets. Scientific Reports, 10, 10907, doi:10.1038/s41598-020-67751-7.CrossRefGoogle ScholarPubMed
Ossa Ossa, F. et al. (2022) Moderate levels of oxygenation during the late stage of Earth’s Great Oxidation Event. Earth and Planetary Science Letters, 594, 117716, doi:10.1016/j.epsl.2022.117716.CrossRefGoogle Scholar
Ossa Ossa, F., Hofmann, A., Spangenberg, J.E., Poulton, S.W., Stüeken, E.E., Schoenberg, R., Eickmann, B., Wille, M., Butler, M., and Bekker, A. (2019) Limited oxygen production in the Mesoarchean ocean. Proceedings of the National Academy of Sciences, 116, 66476652, doi:10.1073/pnas.1818762116.CrossRefGoogle ScholarPubMed
Ossa Ossa, F., Hofmann, A., Vidal, O., Kramers, J.D., Belyanin, G. and Cavalazzi, B. (2016) Unusual manganese enrichment in the Mesoarchean Mozaan Group, Pongola Supergroup, South Africa. Precambrian Research, 281, 414433, doi:10.1016/j.precamres.2016.06.009.CrossRefGoogle Scholar
Ossa Ossa, F., Hofmann, A., Wille, M., Spangenberg, J.E., Bekker, A., Poulton, S.W., Eickmann, B. and Schoenberg, R. (2018) Aerobic iron and manganese cycling in a redox-stratified Mesoarchean epicontinental sea. Earth and Planetary Science Letters, 500, 2840, doi:10.1016/j.epsl.2018.07.044.CrossRefGoogle Scholar
Ostrander, C.M., Johnson, A.C. and Anbar, A.D. (2021) Earth’s First Redox Revolution. Annual Review of Earth and Planetary Sciences, 49, 337366, doi:10.1146/annurev-earth-072020-055249.CrossRefGoogle Scholar
Ostrander, C.M., Nielsen, S.G., Owens, J.D., Kendall, B., Gordon, G.W., Romaniello, S.J. and Anbar, A.D. (2019) Fully oxygenated water columns over continental shelves before the Great Oxidation Event. Nature Geoscience, 12, 186191, doi:10.1038/s41561-019-0309-7.CrossRefGoogle ScholarPubMed
Oswald, K., Milucka, J., Brand, A., Littmann, S., Wehrli, B., Kuypers, M.M.M. and Schubert, C.J. (2015) Light-dependent aerobic methane oxidation reduces methane emissions from seasonally stratified lakes. PLoS ONE, 10, e0132574, doi:10.1371/journal.pone.0132574.CrossRefGoogle ScholarPubMed
Ozaki, K., Thompson, K.J., Simister, R.L., Crowe, S.A. and Reinhard, C.T. (2019) Anoxygenic photosynthesis and the delayed oxygenation of Earth’s atmosphere. Nature Communications, 10, 3026, doi:10.1038/s41467-019-10872-z.CrossRefGoogle ScholarPubMed
Pahlevan, K., Schaefer, L. and Hirschmann, M.M. (2019) Hydrogen isotopic evidence for early oxidation of silicate Earth. Earth and Planetary Science Letters, 526, 115770, doi:10.1016/j.epsl.2019.115770.CrossRefGoogle ScholarPubMed
Papineau, D., Mojzsis, S.J., Coath, C.D., Karhu, J.A. and McKeegan, K.D. (2005) Multiple sulfur isotopes of sulfides from sediments in the aftermath of Paleoproterozoic glaciations. Geochimica et Cosmochimica Acta, 69, 50335060, doi:10.1016/j.gca.2005.07.005.CrossRefGoogle Scholar
Papineau, D., Mojzsis, S.J. and Schmitt, A.K. (2007) Multiple sulfur isotopes from Paleoproterozoic Huronian interglacial sediments and the rise of atmospheric oxygen. Earth and Planetary Science Letters, 255, 188212, doi:10.1016/j.epsl.2006.12.015.CrossRefGoogle Scholar
Papineau, D., She, Z., Dodd, M.S., Iacoviello, F., Slack, J.F., Hauri, E., Shearing, P. and Little, C.T.S. (2022) Metabolically diverse primordial microbial communities in Earth’s oldest seafloor-hydrothermal jasper: Science Advances, 8, eabm2296, doi:10.1126/sciadabm2296.CrossRefGoogle ScholarPubMed
Parsons, C., Stüeken, E.E., Rosen, C.J., Mateos, K. and Anderson, R.E. (2021) Radiation of nitrogen-metabolizing enzymes across the tree of life tracks environmental transitions in Earth history. Geobiology, 19, 1834, doi:10.1111/gbi.12419.CrossRefGoogle ScholarPubMed
Partridge, M.A., Golding, S.D., Baublys, K.A. and Young, E. (2008) Pyrite paragenesis and multiple sulfur isotope distribution in late Archean and early Paleoproterozoic Hamersley Basin sediments. Earth and Planetary Science Letters, 272, 4149, doi:10.1016/j.epsl.2008.03.051.CrossRefGoogle Scholar
Pavlov, A.A., Brown, L.L. and Kasting, J.F. (2001) UV shielding of NH3 and O2 by organic hazes in the Archean atmosphere. Journal of Geophysical Research: Planets, 106, 2326723287, doi:10.1029/2000JE001448.CrossRefGoogle Scholar
Pavlov, A.A. and Kasting, J.F. (2002) Mass-independent fractionation of sulfur isotopes in Archean sediments: Strong evidence for an anoxic Archean atmosphere. Astrobiology, 2, 2741, doi:10.1089/153110702753621321.CrossRefGoogle ScholarPubMed
Peck, W.H., Valley, J.W., Wilde, S.A. and Graham, C.M. (2001) Oxygen isotope ratios and rare earth elements in 3.3 to 4.4 Ga zircons: Ion microprobe evidence for high δ18O continental crust and oceans in the Early Archean. Geochimica et Cosmochimica Acta, 65, 42154229.10.1016/S0016-7037(01)00711-6CrossRefGoogle Scholar
Pecoits, E., Smith, M.L., Catling, D.C., Philippot, P., Kappler, A. and Konhauser, K.O. (2015) Atmospheric hydrogen peroxide and Eoarchean iron formations. Geobiology, 13, 114, doi:10.1111/gbi.12116.CrossRefGoogle ScholarPubMed
Pellerin, A., Thomazo, C., Ader, M., Marin-Carbonne, J., Alleon, J., Vennin, E. and Hofmann, A. (2023) Iron-mediated anaerobic ammonium-oxidation recorded in the early Archean ferruginous ocean. Geobiology, 21, 277289.10.1111/gbi.12540CrossRefGoogle ScholarPubMed
Pellerin, A., Thomazo, C., Ader, M., Rossignol, C., Rego, E.S., Busigny, V. and Philippot, P. (2024) Neoarchaean oxygen-based nitrogen cycle en route to the Great Oxidation Event. Nature, 633, 365370, doi:10.1038/s41586-024-07842-x.CrossRefGoogle Scholar
Philippot, P., Van Zuilen, M., Lepot, K., Thomazo, C., Farquhar, J. and Van Kranendonk, M.J., (2007) Early Archaean microorganisms preferred elemental sulfur, not sulfate. Science, 317, 15341537, doi:10.1126/science.1145861.CrossRefGoogle Scholar
Picard, A., Kappler, A., Schmid, G., Quaroni, L. and Obst, M. (2015) Experimental diagenesis of organo-mineral structures formed by microaerophilic Fe(II)-oxidizing bacteria. Nature Communications, 6, 6277, doi:10.1038/ncomms7277.CrossRefGoogle ScholarPubMed
Pirajno, F. and Yu, H.-C. (2021) Cycles of hydrothermal activity, precipitation of chemical sediments, with special reference to Algoma-type BIF. Gondwana Research, 100, 251260, doi:10.1016/j.gr.2021.02.012.CrossRefGoogle Scholar
Planavsky, N.J. et al. (2014) Evidence for oxygenic photosynthesis half a billion years before the Great Oxidation Event. Nature Geoscience, 7, 283286, doi:10.1038/ngeo2122.CrossRefGoogle Scholar
Postgate, J.R. (1998) Nitrogen fixation. Cambridge University Press, London.Google Scholar
Poulton, S.W., Bekker, A., Cumming, M., Zerkle, A.L., Canfield, D.E. and Johnston, D.T. (2021) A 200-million-year delay in permanent atmospheric oxygenation. Nature, 592, 232236, doi:10.1038/s41586-021-03393-7.CrossRefGoogle ScholarPubMed
Ramdohr, P. (1958) New observations of the ores of the Witwatersrand in South Africa and their genetic significance. Geological Society of South Africa Transactions, 61, 150.Google Scholar
Rasmussen, B. and Buick, R. (1999) Redox state of the Archean atmosphere: Evidence from detrital heavy minerals in ca. 3250–2750 Ma sandstones from the Pilbara Craton, Australia. Geology, 27, 115, doi:10.1130/0091-7613(1999)027<0115:RSOTAA>2.3.CO;2.2.3.CO;2>CrossRefGoogle Scholar
Rasmussen, B., Fletcher, I.R., Brocks, J.J. and Kilburn, M.R. (2008) Reassessing the first appearance of eukaryotes and cyanobacteria. Nature, 455, 11011104, doi:10.1038/nature07381.CrossRefGoogle ScholarPubMed
Rasmussen, B., Muhling, J.R., Suvorova, A. and Fischer, W.W. (2021) Apatite nanoparticles in 3.46–2.46 Ga iron formations: Evidence for phosphorus-rich hydrothermal plumes on early Earth. Geology, 49, 647651, doi:10.1130/G48374.1.CrossRefGoogle Scholar
Rasmussen, B., Muhling, J.R., Tosca, N.J. and Fischer, W.W. (2023) Did nutrient-rich oceans fuel Earth’s oxygenation? Geology, 51(5), 444448, doi:10.1130/G50835.1.CrossRefGoogle Scholar
Reinhard, C.T. and Planavsky, N.J. (2011) Mineralogical constraints on Precambrian pCO2. Nature, 474, E1, doi:10.1038/nature09959.CrossRefGoogle ScholarPubMed
Reinhard, C.T. and Planavsky, N.J. (2022) The History of Ocean Oxygenation. Annual Review of Marine Science, 14, 331353, doi:10.1146/annurev-marine-031721-104005.CrossRefGoogle ScholarPubMed
Reinhard, C.T., Planavsky, N.J., Gill, B.C., Ozaki, K., Robbins, L.J., Lyons, T.W., Fischer, W.W., Wang, C., Cole, D.B. and Konhauser, K.O. (2017) Evolution of the global phosphorus cycle. Nature, 541, 386389, doi:10.1038/nature20772.CrossRefGoogle ScholarPubMed
Reinhard, C.T., Raiswell, R., Scott, C., Anbar, A.D. and Lyons, T.W. (2009) A Late Archean sulfidic sea stimulated by early oxidative weathering of the continents. Science, 326, 713716, doi:10.1126/science.1176711.CrossRefGoogle ScholarPubMed
Righter, K. and Drake, M.J., (1996) Core formation in Earth’s Moon, Mars, and Vesta. Icarus, 124, 513529, doi:10.1006/icar.1996.0227.CrossRefGoogle Scholar
Righter, K., Yang, H., Costin, G. and Downs, R.T. (2008) Oxygen fugacity in the Martian mantle controlled by carbon: New constraints from the nakhlite MIL 03346. Meteoritics & Planetary Science, 43, 17091723, doi:10.1111/j.1945-5100.2008.tb00638.x.CrossRefGoogle Scholar
Ringwood, A.E. (1979) Origin of the Earth and Moon. Springer Science & Business Media, New York, 295 pp.10.1007/978-1-4612-6167-4CrossRefGoogle Scholar
Rios-Del Toro, E.E., Valenzuela, E.I., López-Lozano, N.E., Cortés-Martínez, M.G., Sánchez-Rodríguez, M.A., Calvario-Martínez, O., Sánchez-Carrillo, S. and Cervantes, F.J. (2018) Anaerobic ammonium oxidation linked to sulfate and ferric iron reduction fuels nitrogen loss in marine sediments. Biodegradation, 29, 429442, doi:10.1007/s10532-018-9839-8.CrossRefGoogle ScholarPubMed
Rizo, H., Boyet, M., Blichert-Toft, J. and Rosing, M.T. (2013) Early mantle dynamics inferred from 142Nd variations in Archean rocks from southwest Greenland. Earth and Planetary Science Letters, 377–378, 324335, doi:10.1016/j.epsl.2013.07.012.CrossRefGoogle Scholar
Rizo, H., Walker, R.J., Carlson, R.W., Horan, M.F., Mukhopadhyay, S., Manthos, V., Francis, D. and Jackson, M.G. (2016a) Preservation of Earth-forming events in the tungsten isotopic composition of modern flood basalts. Science, 352, 809812, doi:10.1126/science.aad8563.CrossRefGoogle Scholar
Rizo, H., Walker, R.J., Carlson, R.W., Touboul, M., Horan, M.F., Puchtel, I.S., Boyet, M. and Rosing, M.T. (2016b) Early Earth differentiation investigated through 142Nd, 182W, and highly siderophile element abundances in samples from Isua, Greenland. Geochimica et Cosmochimica Acta, 175, 319336, doi:10.1016/j.gca.2015.12.007.CrossRefGoogle Scholar
Robbins, L.J. et al. (2023) Manganese oxides, Earth surface oxygenation, and the rise of oxygenic photosynthesis. Earth-Science Reviews, 239, 104368, doi:10.1016/j.earscire2023.104368.CrossRefGoogle Scholar
Roger, A.J. and Susko, E. (2018) Molecular clocks provide little information to date methanogenic Archaea. Nature Ecology & Evolution, 2, 16761677, doi:10.1038/s41559-018-0687-z.CrossRefGoogle ScholarPubMed
Rollinson, H., Adetunji, J., Lenaz, D. and Szilas, K. (2017) Archaean chromitites show constant Fe3+/ΣFe in Earth’s asthenospheric mantle since 3.8 Ga. Lithos, 282–283, 316325, doi:10.1016/j.lithos.2017.03.020.CrossRefGoogle Scholar
Rubie, D.C., Frost, D.J., Mann, U., Asahara, Y., Nimmo, F., Tsuno, K., Kegler, P., Holzheid, A. and Palme, H. (2011) Heterogeneous accretion, composition and core–mantle differentiation of the Earth. Earth and Planetary Science Letters, 301, 3142, doi:10.1016/j.epsl.2010.11.030.CrossRefGoogle Scholar
Rubie, D.C., Jacobson, S.A., Morbidelli, A., O’Brien, D.P., Young, E.D., De Vries, J., Nimmo, F., Palme, H. and Frost, D.J. (2015) Accretion and differentiation of the terrestrial planets with implications for the compositions of early-formed Solar System bodies and accretion of water. Icarus, 248, 89108, doi:10.1016/j.icarus.2014.10.015.CrossRefGoogle Scholar
Ruff, S.E. et al. (2023) Hydrogen and dark oxygen drive microbial productivity in diverse groundwater ecosystems. Nature Communications, 14, 3194, doi:10.1038/s41467-023-38523-4.CrossRefGoogle ScholarPubMed
Runge, E.A., Mansor, M., Kappler, A. and Duda, J.-P. (2023) Microbial biosignatures in ancient hydrothermal sulfides. Geobiology, 21, 355377.10.1111/gbi.12539CrossRefGoogle ScholarPubMed
Rush, J.D. and Bielski, B.H.J. (1985) Pulse radiolytic studies of the reaction of perhydroxyl/superoxide O2- with iron(II)/iron(III) ions. The reactivity of HO2/O2- with ferric ions and its implication on the occurrence of the Haber–Weiss reaction. The Journal of Physical Chemistry, 89, 50625066, doi:10.1021/j100269a035.CrossRefGoogle Scholar
Rye, R. and Holland, H.D. (1998) Paleosols and the evolution of atmospheric oxygen: a critical review. American Journal of Science, 298, 621672.10.2475/ajs.298.8.621CrossRefGoogle ScholarPubMed
Sánchez-Baracaldo, P. (2015) Origin of marine planktonic cyanobacteria. Scientific Reports, 5, 17418, doi:10.1038/srep17418.CrossRefGoogle ScholarPubMed
Sánchez-Baracaldo, P., Bianchini, G., Di Cesare, A., Callieri, C. and Chrismas, N.A.M. (2019) Insights Into the Evolution of Picocyanobacteria and Phycoerythrin Genes (mpeBA and cpeBA). Frontiers in Microbiology, 10, 45, doi:10.3389/fmicb.2019.00045.CrossRefGoogle ScholarPubMed
Sánchez-Baracaldo, P., Raven, J.A., Pisani, D. and Knoll, A.H. (2017) Early photosynthetic eukaryotes inhabited low-salinity habitats. Proceedings of the National Academy of Sciences, 114, doi:10.1073/pnas.1620089114.CrossRefGoogle ScholarPubMed
Sánchez-Baracaldo, P., Ridgwell, A. and Raven, J.A. (2014) A Neoproterozoic Transition in the Marine Nitrogen Cycle. Current Biology, 24, 652657, doi:10.1016/j.cub.2014.01.041.CrossRefGoogle ScholarPubMed
Scaillet, B. and Gaillard, F. (2011) Redox state of early magmas. Nature, 480, 4849, doi:10.1038/480048a.CrossRefGoogle ScholarPubMed
Schaefer, L. and Elkins-Tanton, L.T. (2018) Magma oceans as a critical stage in the tectonic development of rocky planets. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 376, 20180109, doi:10.1098/rsta.2018.0109.CrossRefGoogle ScholarPubMed
Schidlowski, M. (2001) Carbon isotopes as biogeochemical recorders of life over 3.8 Ga of Earth history: evolution of a concept. Precambrian Research, 106, 117134, doi:10.1016/S0301-9268(00)00128-5.CrossRefGoogle Scholar
Schidlowski, M. (1981) Uraniferous constituents of the Witwatersrand conglomerates: ore-microscopic observations and implications for the Witwatersrand metallogeny.Google Scholar
Schippers, A., Neretin, L.N., Lavik, G., Leipe, T. and Pollehne, F. (2005) Manganese(II) oxidation driven by lateral oxygen intrusions in the western Black Sea. Geochimica et Cosmochimica Acta, 69, 22412252, doi:10.1016/j.gca.2004.10.016.CrossRefGoogle Scholar
Schirrmeister, B.E., Gugger, M. and Donoghue, P.C.J. (2015) Cyanobacteria and the Great Oxidation Event: evidence from genes and fossils. Palaeontology, 58, 769785, doi:10.1111/pala.12178.CrossRefGoogle ScholarPubMed
Schirrmeister, B.E., Sanchez-Baracaldo, P. and Wacey, D. (2016) Cyanobacterial evolution during the Precambrian. International Journal of Astrobiology, 15, 187204, doi:10.1017/S1473550415000579.CrossRefGoogle Scholar
Schönbächler, M., Carlson, R.W., Horan, M.F., Mock, T.D. and Hauri, E.H. (2010) Heterogeneous accretion and the moderately volatile element budget of Earth. Science, 328, 884887, doi:10.1126/science.1186239.CrossRefGoogle ScholarPubMed
Schopf, J.W. (1993) Microfossils of the Early Archean Apex Chert: New evidence of the antiquity of life. Nature, 260, 640646.Google ScholarPubMed
Schröder, S., Bedorf, D., Beukes, N.J. and Gutzmer, J. (2011) From BIF to red beds: Sedimentology and sequence stratigraphy of the Paleoproterozoic Koegas Subgroup (South Africa). Sedimentary Geology, 236, 2544, doi:10.1016/j.sedgeo.2010.11.007.CrossRefGoogle Scholar
Schwieterman, E.W. et al. (2018) Exoplanet biosignatures: a review of remotely detectable signs of Life. Astrobiology, 18, 663708, doi:10.1089/ast.2017.1729.CrossRefGoogle ScholarPubMed
Scott, C.T., Bekker, A., Reinhard, C.T., Schnetger, B., Krapež, B., Rumble, D. and Lyons, T.W. (2011) Late Archean euxinic conditions before the rise of atmospheric oxygen. Geology, 39, 119122, doi:10.1130/G31571.1.CrossRefGoogle Scholar
Scott, C., Lyons, T.W., Bekker, A., Shen, Y., Poulton, S.W., Chu, X. and Anbar, A.D. (2008) Tracing the stepwise oxygenation of the Proterozoic ocean. Nature, 452, 456459, doi:10.1038/nature06811.CrossRefGoogle ScholarPubMed
Segerer, A., Neuner, A., Kristjansson, J.K. and Stetter, K.O. (1986) Acidianus infernus gen. nov., sp. nov., and Acidianus brierleyi Comb. nov.: Facultatively aerobic, extremely acidophilic thermophilic sulfur-metabolizing Archaebacteria. International Journal of Systematic Bacteriology, 36, 559564, doi:10.1099/00207713-36-4-559.CrossRefGoogle Scholar
Sekine, Y. et al. (2011) Manganese enrichment in the Gowganda Formation of the Huronian Supergroup: A highly oxidizing shallow-marine environment after the last Huronian glaciation: Earth and Planetary Science Letters, 307, 201210, doi:10.1016/j.epsl.2011.05.001.CrossRefGoogle Scholar
Sengupta, S. and Pack, A. (2018) Triple oxygen isotope mass balance for the Earth’s oceans with application to Archean cherts. Chemical Geology, 495, 1826, doi:10.1016/j.chemgeo.2018.07.012.CrossRefGoogle Scholar
Sharp, Z.D., McCubbin, F.M. and Shearer, C.K. (2013) A hydrogen-based oxidation mechanism relevant to planetary formation. Earth and Planetary Science Letters, 380, 8897, doi:10.1016/j.epsl.2013.08.015.CrossRefGoogle Scholar
Shen, Y., Farquhar, J., Masterson, A., Kaufman, A.J. and Buick, R. (2009) Evaluating the role of microbial sulfate reduction in the early Archean using quadruple isotope systematics. Earth and Planetary Science Letters, 279, 383391, doi:10.1016/j.epsl.2009.01.018.CrossRefGoogle Scholar
Sherwood Lollar, B. and McCollom, T.M. (2006) Biosignatures and abiotic constraints on early life. Nature, 444, E18, doi:10.1038/nature05499.CrossRefGoogle Scholar
Shi, T. and Falkowski, P.G. (2008) Genome evolution in cyanobacteria: The stable core and the variable shell. Proceedings of the National Academy of Sciences, 105, 25102515, doi:10.1073/pnas.0711165105.CrossRefGoogle ScholarPubMed
Shibuya, T., Komiya, T., Nakamura, K., Takai, K. and Maruyama, S. (2010) Highly alkaline, high-temperature hydrothermal fluids in the early Archean ocean. Precambrian Research, 182, 230238, doi:10.1016/j.precamres.2010.08.011.CrossRefGoogle Scholar
Siebert, J., Badro, J., Antonangeli, D. and Ryerson, F.J. (2013) Terrestrial accretion under oxidizing conditions. Science, 339, 11941197, doi:10.1126/science.1227923.CrossRefGoogle ScholarPubMed
Smart, K.A., Tappe, S., Stern, R.A., Webb, S.J. and Ashwal, L.D. (2016) Early Archaean tectonics and mantle redox recorded in Witwatersrand diamonds. Nature Geoscience, 9, 255259, doi:10.1038/ngeo2628.CrossRefGoogle Scholar
Smit, M.A. and Mezger, K. (2017) Earth’s early O2 cycle suppressed by primitive continents. Nature Geoscience, 10, 788792, doi:10.1038/ngeo3030.CrossRefGoogle Scholar
Smith, A.J.B. (2018) The iron formations of Southern Africa. Pp. 469491 in: Geology of Southwest Gondwana (Siegesmund, S., Basei, M.A.S., Oyhantçabal, P., and Oriolo, S., editors), Springer International Publishing, Cham, doi:10.1007/978-3-319-68920-3_17.CrossRefGoogle Scholar
Smith, A.J.B. and Beukes, N.J. (2023) The paleoenvironmental implications of pre-Great Oxidation Event manganese deposition in the Mesoarchean Ijzermijn Iron Formation Bed, Mozaan Group, Pongola Supergroup, South Africa. Precambrian Research, 384, 106922, doi:10.1016/j.precamres.2022.106922.CrossRefGoogle Scholar
Smithies, R.H., Champion, D.C., Van Kranendonk, M.J., Howard, H.M. and Hickman, A.H. (2005) Modern-style subduction processes in the Mesoarchaean: Geochemical evidence from the 3.12 Ga Whundo intra-oceanic arc. Earth and Planetary Science Letters, 231, 221237, doi:10.1016/j.epsl.2004.12.026.CrossRefGoogle Scholar
Søgaard, E.G., Medenwaldt, R. and Abraham-Peskir, J.V. (2000) Conditions and rates of biotic and abiotic iron precipitation in selected Danish freshwater plants and microscopic analysis of precipitate morphology. Water Research, 34, 26752682, doi:10.1016/S0043-1354(00)00002-6.CrossRefGoogle Scholar
Sommer, T. et al., (2017) Bacteria-induced mixing in natural waters. Geophysical Research Letters, 44, 94249432.10.1002/2017GL074868CrossRefGoogle Scholar
Soo, R.M., Hemp, J., Parks, D.H., Fischer, W.W. and Hugenholtz, P. (2017) On the origins of oxygenic photosynthesis and aerobic respiration in Cyanobacteria. Science, 355, 14361440, doi:10.1126/science.aal3794.CrossRefGoogle ScholarPubMed
Stagno, V., and Aulbach, S. (2021) Redox processes before, during, and after Earth’s accretion affecting the deep carbon cycle in: Magma Redox Geochemistry (Moretti, R. and Neuville, D.R., editors), Geophysical Monograph Series Vol. 266, AGU & Wiley, Hoboken, USA.Google Scholar
Stagno, V. and Fei, Y. (2020) The redox boundaries of Earth’s interior. Elements, 16, 167172, doi:10.2138/gselements.16.3.167.CrossRefGoogle Scholar
Stockey, R.G., Cole, D.B., Farrell, U.C., Agić, H., Boag, T.H., Brocks, J.J., Canfield, D.E., Cheng, M., Crockford, P.W., Cui, H., Dahl, T.W., Del Mouro, L., Dewing, K., Dornbos, S.Q., Emmings, J.F., Gaines, R.R., Gibson, T.M., Gill, B.C., Gilleaudeau, G.J.Sperling, E.A. (2024) Sustained increases in atmospheric oxygen and marine productivity in the Neoproterozoic and Palaeozoic eras. Nature Geoscience, 17(7), 667674. https://doi.org/10.1038/s41561-024-01479-1.CrossRefGoogle Scholar
Stolper, D.A., Revsbech, N.P. and Canfield, D.E. (2010) Aerobic growth at nanomolar oxygen concentrations. Proceedings of the National Academy of Sciences, 107, 1875518760, doi:10.1073/pnas.1013435107.CrossRefGoogle ScholarPubMed
Stone, J., Edgar, J.O., Gould, J.A. and Telling, J. (2022) Tectonically-driven oxidant production in the hot biosphere. Nature Communications, 13, 4529, doi:10.1038/s41467-022-32129-y.CrossRefGoogle ScholarPubMed
Stüeken, E.E., Boocock, T.J., Robinson, A., Mikhail, S. and Johnson, B.W. (2021) Hydrothermal recycling of sedimentary ammonium into oceanic crust and the Archean ocean at 3.24 Ga. Geology, 49, 822826, doi:10.1130/G48844.1.CrossRefGoogle Scholar
Stüeken, E.E. and Buick, R. (2018) Environmental control on microbial diversification and methane production in the Mesoarchean. Precambrian Research, 304, 6472, doi:10.1016/j.precamres.2017.11.003.CrossRefGoogle Scholar
Stüeken, E.E., Buick, R., Anderson, R.E., Baross, J.A., Planavsky, N.J. and Lyons, T.W. (2017) Environmental niches and metabolic diversity in Neoarchean lakes. Geobiology, 15, 767783, doi:10.1111/gbi.12251.CrossRefGoogle ScholarPubMed
Stüeken, E.E., Buick, R., Guy, B.M. and Koehler, M.C. (2015a) Isotopic evidence for biological nitrogen fixation by molybdenum-nitrogenase from 3.2 Gyr. Nature, 520, 666669, doi:10.1038/nature14180.CrossRefGoogle Scholar
Stüeken, E.E., Buick, R. and Schauer, A.J. (2015b) Nitrogen isotope evidence for alkaline lakes on late Archean continents. Earth and Planetary Science Letters, 411, 110, doi:10.1016/j.epsl.2014.11.037.CrossRefGoogle Scholar
Stüeken, E.E., Kipp, M.A., Koehler, M.C. and Buick, R. (2016) The evolution of Earth’s biogeochemical nitrogen cycle. Earth-Science Reviews, 160, 220239, doi:10.1016/j.earscire2016.07.007.CrossRefGoogle Scholar
Stüeken, E.E., Pellerin, A., Thomazo, C., Johnson, B.W., Duncanson, S. and Schoepfer, S.D. (2024) Marine biogeochemical nitrogen cycling through Earth’s history. Nature Reviews Earth & Environment, doi:10.1038/s43017-024-00591-5.CrossRefGoogle Scholar
Sun, L., Lechte, M., Shi, X., Zhou, X., Zhou, L., Feng, H., Xie, B., Wu, M. and Tang, D. (2022) Hexagonal magnetite in Algoma-type banded iron formations of the ca. 2.52 Ga Baizhiyan Formation, North China: Evidence for a green rust precursor? American Mineralogist, 107, 970984.10.2138/am-2022-7981CrossRefGoogle Scholar
Swanner, E.D., Mloszewska, A.M., Cirpka, O.A., Schoenberg, R., Konhauser, K.O. and Kappler, A. (2015a) Modulation of oxygen production in Archaean oceans by episodes of Fe(II) toxicity. Nature Geoscience, 8, 126130, doi:10.1038/ngeo2327.CrossRefGoogle Scholar
Swanner, E.D., Wu, W., Hao, L., Wüstner, M.L., Obst, M., Moran, D.M., McIlvin, M.R., Saito, M.A. and Kappler, A. (2015b) Physiology, Fe(II) oxidation, and Fe mineral formation by a marine planktonic cyanobacterium grown under ferruginous conditions. Frontiers in Earth Science, 3, doi:10.3389/feart.2015.00060.CrossRefGoogle Scholar
Tamaura, Y., Yoshida, T. and Katsura, T. (1984) The synthesis of green rust II (FeIII1-FeII2) and its spontaneous transformation into Fe3O4. Bulletin of the Chemical Society of Japan, 57, 24112416.10.1246/bcsj.57.2411CrossRefGoogle Scholar
Tatzel, M., Frings, P.J., Oelze, M., Herwartz, D., Lünsdorf, N.K. and Wiedenbeck, M. (2022) Chert oxygen isotope ratios are driven by Earth’s thermal evolution. Proceedings of the National Academy of Sciences, 119, e2213076119, doi:10.1073/pnas.2213076119.CrossRefGoogle ScholarPubMed
Tebo, B.M. (1991) Manganese(II) oxidation in the suboxic zone of the Black Sea. Deep Sea Research Part A. Oceanographic Research Papers, 38, S883S905, doi:10.1016/S0198-0149(10)80015-9.CrossRefGoogle Scholar
Tebo, B.M., Bargar, J.R., Clement, B.G., Dick, G.J., Murray, K.J., Parker, D., Verity, R. and Webb, S.M. (2004) Biogenic manganese oxides: properties and mechanisms of formation. Annual Review of Earth and Planetary Sciences, 32, 287328, doi:10.1146/annureearth.32.101802.120213.CrossRefGoogle Scholar
Tebo, B.M., Johnson, H.A., McCarthy, J.K. and Templeton, A.S. (2005) Geomicrobiology of manganese(II) oxidation. Trends in Microbiology, 13, 421428, doi:10.1016/j.tim.2005.07.009.CrossRefGoogle ScholarPubMed
Thamdrup, B., Dalsgaard, T. and Revsbech, N.P. (2012) Widespread functional anoxia in the oxygen minimum zone of the Eastern South Pacific. Deep Sea Research Part I: Oceanographic Research Papers, 65, 3645, doi:10.1016/j.dsr.2012.03.001.CrossRefGoogle Scholar
Thomazo, C., Ader, M., Farquhar, J. and Philippot, P. (2009a) Methanotrophs regulated atmospheric sulfur isotope anomalies during the Mesoarchean (Tumbiana Formation, Western Australia). Earth and Planetary Science Letters, 279, 6575, doi:10.1016/j.epsl.2008.12.036.CrossRefGoogle Scholar
Thomazo, C., Ader, M. and Philippot, P. (2011) Extreme 15N-enrichments in 2.72-Gyr-old sediments: evidence for a turning point in the nitrogen cycle. Geobiology, 9, 107120.10.1111/j.1472-4669.2011.00271.xCrossRefGoogle ScholarPubMed
Thomazo, C., Pinti, D.L., Busigny, V., Ader, M., Hashizume, K. and Philippot, P. (2009b) Biological activity and the Earth’s surface evolution: Insights from carbon, sulfur, nitrogen and iron stable isotopes in the rock record. Comptes Rendus Palevol, 8, 665678, doi:10.1016/j.crp2009.02.003.CrossRefGoogle Scholar
Thompson, K.J. et al. (2019) Photoferrotrophy, deposition of banded iron formations, and methane production in Archean oceans. Science Advances, 5, eaav2869, doi:10.1126/sciadaav2869.CrossRefGoogle ScholarPubMed
Thorpe, R.I. (1979) A sedimentary barite deposit from the Archean Fig Tree Group of the Barberton Mountain Land (South Africa); discussion. Economic Geology, 74, 700702.10.2113/gsecongeo.74.3.700CrossRefGoogle Scholar
Towe, K.M. (1990) Aerobic respiration in the Archaean? Nature, 348, 5456, doi:10.1038/348054a0.CrossRefGoogle ScholarPubMed
Trail, D., Watson, E.B. and Tailby, N.D. (2011) The oxidation state of Hadean magmas and implications for early Earth’s atmosphere. Nature, 480, 7982, doi:10.1038/nature10655.CrossRefGoogle ScholarPubMed
Trendall, A.F. (2002) The significance of iron-formation in the Precambrian stratigraphic record. Pp. 3366 in: Precambrian Sedimentary Environments (Altermann, W. and Corcoran, P.L., editors), Wiley, doi:10.1002/9781444304312.ch3.CrossRefGoogle Scholar
Tsikos, H., Beukes, N.J., Moore, J.M. and Harris, C. (2003) Deposition, diagenesis, and secondary enrichment of metals in the Paleoproterozoic Hotazel Iron Formation, Kalahari manganese field, South Africa. Economic Geology, 98, 14491462.Google Scholar
Tusch, J., Hoffmann, J.E., Hasenstab, E., Fischer-Gödde, M., Marien, C.S., Wilson, A.H. and Münker, C. (2022) Long-term preservation of Hadean protocrust in Earth’s mantle. Proceedings of the National Academy of Sciences, 119, e2120241119, doi:10.1073/pnas.2120241119.CrossRefGoogle ScholarPubMed
Tusch, J., Münker, C., Hasenstab, E., Jansen, M., Marien, C.S., Kurzweil, F., Van Kranendonk, M.J., Smithies, H., Maier, W. and Garbe-Schönberg, D. (2021) Convective isolation of Hadean mantle reservoirs through Archean time. Proceedings of the National Academy of Sciences, 118, e2012626118, doi:10.1073/pnas.2012626118.CrossRefGoogle ScholarPubMed
Ueno, Y., Ono, S., Rumble, D. and Maruyama, S. (2008) Quadruple sulfur isotope analysis of ca. 3.5Ga Dresser Formation: New evidence for microbial sulfate reduction in the early Archean. Geochimica et Cosmochimica Acta, 72, 56755691, doi:10.1016/j.gca.2008.08.026.CrossRefGoogle Scholar
Ueno, Y., Yamada, K., Yoshida, N., Maruyama, S. and Isozaki, Y., 2006, Evidence from fluid inclusions for microbial methanogenesis in the early Archaean era: Nature, 440, 516519, doi:10.1038/nature04584.CrossRefGoogle ScholarPubMed
Valley, J.W., Peck, W.H., King, E.M. and Wilde, S.A. (2002) A cool early Earth. Geology, 30, 351354.10.1130/0091-7613(2002)030<0351:ACEE>2.0.CO;22.0.CO;2>CrossRefGoogle Scholar
Van De Graaf, A.A., Mulder, A., De Bruijn, P., Jetten, M.S., Robertson, L.A. and Kuenen, J.G. (1995) Anaerobic oxidation of ammonium is a biologically mediated process. Applied and Environmental Microbiology, 61, 12461251, doi:10.1128/aem.61.4.1246-1251.1995.CrossRefGoogle ScholarPubMed
Van Kranendonk, M.J. (2011) Onset of Plate Tectonics. Science, 333, 413414, doi:10.1126/science.1208766.CrossRefGoogle ScholarPubMed
Van Kranendonk, M.J., Smithies, H.R., Hickman, A.H. and Champion, D.C. (2007) Review: secular tectonic evolution of Archean continental crust: interplay between horizontal and vertical processes in the formation of the Pilbara Craton, Australia. Terra Nova, 19, 138, doi:10.1111/j.1365-3121.2006.00723.x.CrossRefGoogle Scholar
Wacey, D., Saunders, M., Brasier, M.D. and Kilburn, M.R. (2011) Earliest microbially mediated pyrite oxidation in ~3.4 billion-year-old sediments. Earth and Planetary Science Letters, 301, 393402, doi:10.1016/j.epsl.2010.11.025.CrossRefGoogle Scholar
Wade, J. and Wood, B.J. (2005) Core formation and the oxidation state of the Earth. Earth and Planetary Science Letters, 236, 7895, doi:10.1016/j.epsl.2005.05.017.CrossRefGoogle Scholar
Wade, J. and Wood, B.J. (2001) The Earth’s ‘missing’ niobium may be in the core. Nature, 409, 7578, doi:10.1038/35051064.CrossRefGoogle ScholarPubMed
Wadhwa, M. (2008) Redox conditions on small bodies, the Moon and Mars. Reviews in Mineralogy and Geochemistry, 68, 493510.10.2138/rmg.2008.68.17CrossRefGoogle Scholar
Wadhwa, M. (2001) Redox state of Mars’ upper mantle and crust from Eu anomalies in shergottite pyroxenes. Science, 291, 15271530, doi:10.1126/science.1057594.CrossRefGoogle ScholarPubMed
Waldbauer, J.R., Sherman, L.S., Sumner, D.Y. and Summons, R.E. (2009) Late Archean molecular fossils from the Transvaal Supergroup record the antiquity of microbial diversity and aerobiosis. Precambrian Research, 169, 2847, doi:10.1016/j.precamres.2008.10.011.CrossRefGoogle Scholar
Wallmann, K. (2001) The geological water cycle and the evolution of marine δ18O values. Geochimica et Cosmochimica Acta, 65, 24692485.10.1016/S0016-7037(01)00603-2CrossRefGoogle Scholar
Walter, M.R., Buick, R., and Dunlop, J.S.R. (1980) Stromatolites 3,400–3,500 Myr old from the North Pole area, Western Australia. Nature, 284, 443445, doi:10.1038/284443a0.CrossRefGoogle Scholar
Walton, C.R., Ewens, S., Coates, J.D., Blake, R.E., Planavsky, N.J., Reinhard, C., Ju, P., Hao, J. and Pasek, M.A. (2023) Phosphorus availability on the early Earth and the impacts of life. Nature Geoscience, 16, 399409, doi:10.1038/s41561-023-01167-6.CrossRefGoogle Scholar
Wang, X. et al. (2018) A Mesoarchean shift in uranium isotope systematics. Geochimica et Cosmochimica Acta, 238, 438452, doi:10.1016/j.gca.2018.07.024.CrossRefGoogle Scholar
Wang, M., Jiang, Y.-Y., Kim, K.M., Qu, G., Ji, H.-F., Mittenthal, J.E., Zhang, H.-Y. and Caetano-Anollés, G. (2011) A universal molecular clock of protein folds and its power in tracing the early history of aerobic metabolism and planet oxygenation. Molecular Biology and Evolution, 28, 567582, doi:10.1093/molbev/msq232.CrossRefGoogle ScholarPubMed
Wang, X., Ossa Ossa, F., Hofmann, A., Agangi, A., Paprika, D. and Planavsky, N.J. (2020) Uranium isotope evidence for Mesoarchean biological oxygen production in shallow marine and continental settings. Earth and Planetary Science Letters, 551, 116583, doi:10.1016/j.epsl.2020.116583.CrossRefGoogle Scholar
Wänke, H., Dreibus, G. and Jagoutz, E. (1984) Mantle chemistry and accretion history of the Earth. Pp. 124 in: Archaean Geochemistry (Kröner, A., Hanson, G.N., and Goodwin, A.M., editors). Springer, Berlin, Heidelberg, doi:10.1007/978-3-642-70001-9_1.Google Scholar
Warke, M.R., Di Rocco, T., Zerkle, A.L., Lepland, A., Prave, A.R., Martin, A.P., Ueno, Y., Condon, D.J. and Claire, M.W. (2020) The Great Oxidation Event preceded a Paleoproterozoic ‘snowball Earth’. Proceedings of the National Academy of Sciences, 117, 1331413320, doi:10.1073/pnas.2003090117.CrossRefGoogle ScholarPubMed
Watanabe, Y. and Tajika, E. (2021) Atmospheric oxygenation of the early earth and earth-like planets driven by competition between land and seafloor weathering. Earth, Planets and Space, 73, 188, doi:10.1186/s40623-021-01527-9.CrossRefGoogle Scholar
Whitehill, A.R. and Ono, S. (2012) Excitation band dependence of sulfur isotope mass-independent fractionation during photochemistry of sulfur dioxide using broadband light sources. Geochimica et Cosmochimica Acta, 94, 238253, doi:10.1016/j.gca.2012.06.014.CrossRefGoogle Scholar
Whitehill, A.R., Xie, C., Hu, X., Xie, D., Guo, H. and Ono, S. (2013) Vibronic origin of sulfur mass-independent isotope effect in photoexcitation of SO2 and the implications to the early earth’s atmosphere. Proceedings of the National Academy of Sciences, 110, 1769717702, doi:10.1073/pnas.1306979110.CrossRefGoogle Scholar
Whitehouse, M.J., Kamber, B.S., Fedo, C.M. and Lepland, A. (2005) Integrated Pb- and S-isotope investigation of sulphide minerals from the early Archaean of southwest Greenland. Chemical Geology, 222, 112131, doi:10.1016/j.chemgeo.2005.06.004.CrossRefGoogle Scholar
Williams, R.J.P. and Fraústo Da Silva, J.J.R. (2003) Evolution was chemically constrained. Journal of Theoretical Biology, 220, 323343, doi:10.1006/jtbi.2003.3152.CrossRefGoogle ScholarPubMed
Wolfe, J.M. and Fournier, G.P. (2018) Horizontal gene transfer constrains the timing of methanogen evolution. Nature Ecology & Evolution, 2, 897903, doi:10.1038/s41559-018-0513-7.CrossRefGoogle ScholarPubMed
Wong, M.L., Charnay, B.D., Gao, P., Yung, Y.L. and Russell, M.J. (2017) Nitrogen oxides in early Earth’s atmosphere as electron acceptors for life’s emergence. Astrobiology, 17, 975983, doi:10.1089/ast.2016.1473.CrossRefGoogle ScholarPubMed
Wood, B.J., Bryndzia, L.T. and Johnson, K.E. (1990) Mantle oxidation state and its relationship to tectonic environment and fluid speciation. Science, 248, 337345, doi:10.1126/science.248.4953.337.CrossRefGoogle ScholarPubMed
Wood, B.J. and Virgo, D. (1989) Upper mantle oxidation state: Ferric iron contents of lherzolite spinels by 57Fe Mössbauer spectroscopy and resultant oxygen fugacities. Geochimica et Cosmochimica Acta, 53, 12771291.10.1016/0016-7037(89)90062-8CrossRefGoogle Scholar
Xu, D., Qin, Z., Wang, X., Li, J., Shi, X., Tang, D. and Liu, J. (2023) Extensive sea-floor oxygenation during the early Mesoproterozoic. Geochimica et Cosmochimica Acta, 354, 186196, doi:10.1016/j.gca.2023.06.007.CrossRefGoogle Scholar
Yang, X., Liu, H. and Zhang, K. (2022) Redox geodynamics in Earth’s interior. Science China Earth Sciences, 65, 624640, doi:10.1007/s11430-021-9864-8.CrossRefGoogle Scholar
Yang, W.H., Weber, K.A. and Silver, W.L. (2012) Nitrogen loss from soil through anaerobic ammonium oxidation coupled to iron reduction. Nature Geoscience, 5, 538541, doi:10.1038/ngeo1530.CrossRefGoogle Scholar
Yu, H. and Leadbetter, J.R., (2020) Bacterial chemolithoautotrophy via manganese oxidation. Nature, 583, 453458, doi:10.1038/s41586-020-2468-5.CrossRefGoogle ScholarPubMed
Zahnle, K., Claire, M. and Catling, D. (2006) The loss of mass-independent fractionation in sulfur due to a Palaeoproterozoic collapse of atmospheric methane. Geobiology, 4, 271283.10.1111/j.1472-4669.2006.00085.xCrossRefGoogle Scholar
Zakharov, D.O., Bindeman, I.N., Tanaka, R., Friðleifsson, G.Ó., Reed, M.H. and Hampton, R.L., (2019) Triple oxygen isotope systematics as a tracer of fluids in the crust: A study from modern geothermal systems of Iceland. Chemical Geology, 530, 119312, doi:10.1016/j.chemgeo.2019.119312.CrossRefGoogle Scholar
Zehr, J.P. and Kudela, R.M. (2011) Nitrogen cycle of the open ocean: from genes to ecosystems. Annual Review of Marine Science, 3, 197225, doi:10.1146/annurev-marine-120709-142819.CrossRefGoogle ScholarPubMed
Zerkle, A.L., House, C.H., Cox, R.P. and Canfield, D.E. (2006) Metal limitation of cyanobacterial N 2 fixation and implications for the Precambrian nitrogen cycle. Geobiology, 4, 285297, doi:10.1111/j.1472-4669.2006.00082.x.CrossRefGoogle Scholar
Zhang, X., Shu, D., Han, J., Zhang, Z., Liu, J. and Fu, D. (2014) Triggers for the Cambrian explosion: Hypotheses and problems. Gondwana Research, 25, 896909, doi:10.1016/j.gr.2013.06.001.CrossRefGoogle Scholar
Zhao, M., Mills, B.J.W., Homoky, W.B. and Peacock, C.L. (2023) Oxygenation of the Earth aided by mineral–organic carbon preservation. Nature Geoscience, 16, 262267, doi:10.1038/s41561-023-01133-2.CrossRefGoogle Scholar
Figure 0

Table 1. Net reactions of microbial metabolism discussed in this paper (Konhauser, 2007).

Figure 1

Figure 1. Commonly used mineral redox buffers and their relationship to ƒO2 plotted over temperature. Quartz-fayalite-magnetite QFM) and nickel-nickel-oxide (NiNiO) depict oxidised conditions, while iron-wustite (IW) and quartz-iron-fayalite (QIF) represent reduced conditions.

Figure 2

Figure 2. Evolution of the Earth´s redox state for different formation stages. The changing ƒO2 is indicated by the deviation in log units from the quartz-fayalite-magnetite (QFM) buffer and is explained in the text. The colours range from blue (reduced) to red (oxidised). The Earth is assumed to become more oxidised with time, with the most reduced values during the accretion period before core formation. It is thought that during the magma ocean period, ƒO2 evolved towards more oxidised values. The modern Earth is comparatively oxidised, with a decreasing redox state with depth (after McCammon, 2005). See the text for references on the redox state of the early Earth.

Figure 3

Figure 3. Literature estimates regarding the ƒO2 of the upper mantle on early Earth (after Aulbach and Stagno, 2016; Schaefer and Elkins-Tanton, 2018; Stagno and Aulbach, 2021). The shaded rectangles represent the results of the ƒO2 estimated from individual studies (normalised to the QFM buffer) over the respective ages of the samples examined. The references for the horizontal rectangles are Canil (1997), Delano (2001), Li and Lee (2004), Nicklas et al. (2018, 2019), Aulbach and Stagno (2016) (AS), Aulbach et al. (2017). The crosses are data points from Nicklas et al. (2018) and Nicklas et al. (2019) that represent an estimated ƒO2 based on the redox-dependent partitioning of vanadium between liquidus olivine and melt. The squares are orogenic eclogites, the circles are mantle eclogites and the diamond is a mid-ocean ridge ophiolite from Aulbach and Stagno (2016). The symbols display the ƒO2 (corrected to 1 GPa) calculated from V/Sc ratios. The vertical error bars are predicted 1σ errors of the V/Sc ratios (representing 1σ of the mean per sample suite) and the horizontal error bars show age ranges or 1σ errors for isochron ages from the literature. The red star shows the calculated ƒO2 of the modern MORB and the arrow points toward the estimated ƒO2 of the uppermost mantle according to Trail et al. (2011) of samples from 4.4 Ga. QFM, quartz-fayalite-magnetite.

Figure 4

Figure 4. Oxidation and oxygenation of the upper mantle and the surface environment over time. (a) Calculated ƒO2 of samples derived from the upper mantle after Aulbach and Stagno (2016) and Stagno and Aulbach (2021) (see Fig. 3 for the legend). The rectangles at the top of the figure display some important geodynamic events: extensive formation of continental crust ca. 3.5–2.4 Ga (Collerson and Kamber, 1999; Huston and Logan, 2004), onset of modern style plate tectonics (PT) ca. 3.2–3.0 Ga (Smithies et al., 2005; Van Kranendonk et al., 2007; Van Kranendonk, 2011; Duncan and Dasgupta, 2017; Kuang et al., 2023), major crustal growth ca. 2.7 Ga (Gaillard et al., 2011), intrusion of large igneous provinces (LIPs) 2.5–2.4 Ga (Ernst and Bleeker, 2010; Gumsley et al., 2017), first glaciations (Ice) ca. 2.4 Ga (Kirschvink et al., 2000; Gumsley et al., 2017). (b) Partial pressure of specific gas species over time after Catling and Zahnle (2020). The two grey dashed lines indicate the partial pressure of CO2. The upper line (K) is after Kasting (1987) and Herwartz et al. (2021), and the lower line (CZ) is after Catling and Zahnle (2020). (c) Geological evidence for oxygenation of the surface environment is explained in detail in the text. The red star indicates early whiffs of oxygen at 3.0 Ga. WO, more abundant later whiffs of oxygen; BS, black shales; Red, red beds (see the main text for references). Mass-independent sulfur isotope fractionation (MIF-S) marks the positive and negative excursions of Δ33S in ‰ (after Ono, 2017; see also the main text). The data for banded iron formations (BIFs) (S, = superior-type; A, algoma-type) and sulfate deposits (barite) is from Huston and Logan (2004). Reddish colours display evidence of oxygenation, while blueish colours indicate reduced conditions. The boxes without colours are deposits discussed in the literature as possible hints for redox conditions, even though the general opinion is that they cannot be used as redox proxies. (d) Timetable for the emergence of the microbial metabolic processes discussed in this review. Solid lines represent well-established timeframes. Dashed lines represent tentative timeframes. Question marks signify highly uncertain periods. Adapted from Lepot (2020) and modified based on references in the text. The red bar across the whole figure indicates the timing of the GOE (2.5–2.3 Ga; see text for references). GOE, Great Oxygenation Event; QFM, quartz-fayalite-magnetite

Figure 5

Figure 5. Schematic of microbial O2 sources and sinks on the late Archean Earth. Oxygenic photosynthesis is the major O2 source. Biomass from primary productivity, anaerobic respiration (i.e. dissimilatory reduction of NO3-, Mn(IV), Fe(III), SO42-), methanogenesis and abiotic sources (not indicated) yield diverse O2 sinks (i.e. Corg, NH4+, Mn(II), Fe(II), H2S, CH4). Aerobic microorganisms couple the oxidation of these sinks to the reduction of O2, forming the microbial O2 sink. The geochemical zonation on the left was redrawn from Canfield and Thamdrup (2009). Note that this zonation reflects the decreasing energy yield of the corresponding respiration process and may strongly overlap in natural environments, therefore it does not necessarily match the depth profile of the indicated chemical species.