Skip to main content
Nature Communications logoLink to Nature Communications
. 2025 Feb 22;16:1873. doi: 10.1038/s41467-025-57091-3

Anomalous δ15N values in the Neoarchean associated with an abundant supply of hydrothermal ammonium

Ashley N Martin 1,2,, Eva E Stüeken 3, Michelle M Gehringer 4, Monika Markowska 2,5, Hubert Vonhof 5, Stefan Weyer 1, Axel Hofmann 6
PMCID: PMC11845595  PMID: 39984454

Abstract

Unusually high δ15N values in the Neoarchean sedimentary record in the time period from 2.8 to 2.6 Ga, termed the Nitrogen Isotope Event (NIE), might be explained by aerobic N cycling prior to the Great Oxidation Event (GOE). Here we report strongly positive δ15N values up to +42.5 ‰ in ~2.75 – 2.73 Ga shallow-marine carbonates from Zimbabwe. As the corresponding deeper-marine shales exhibit negative δ15N values that are explained by partial biological uptake from a large ammonium reservoir, we interpret our data to have resulted from hydrothermal upwelling of 15N-rich ammonium into shallow, partially oxic waters, consistent with uranium isotope variations. This work shows that anomalous N isotope signatures at the onset of the NIE temporally correlate with extensive volcanic and hydrothermal activity both locally and globally, which may have stimulated primary production and spurred biological innovation in the lead-up to the GOE.

Subject terms: Marine chemistry, Element cycles


Unusually elevated nitrogen isotope values in 2.75-2.73 Ga carbonates from Zimbabwe, linked to the Nitrogen Isotope Event, suggest that hydrothermal ammonium fertilised shallow marine life in some areas in the buildup to the Great Oxidation Event.

Introduction

The remarkable variation of nitrogen (N) isotope values (δ15N = [(15N/14Nsample)/(15N/14Nair) − 1] × 1000) in Neoarchean sedimentary rocks, from −11‰1 up to 50‰2,3, hints at fundamental shifts in global marine N cycling prior to the initial rise of atmospheric oxygen during the great oxidation event (GOE). The time interval that features this particular cluster of strongly positive N isotope values around 2.8 to 2.6 billion years ago (Ga) has recently been termed the nitrogen isotope event (NIE) and suggested to relate to the advent of aerobic ammonium oxidation globally4. Previously, such occurrences have been explained in terms of either partial nitrification coupled to denitrification in a marine oxygen oasis3 or NH3 volatilisation under high-pH conditions in a lacustrine setting2, whereby both scenarios invoke the loss of isotopically light N to the atmosphere. However, the operation of a limited aerobic N cycle during the Neoarchean is at odds with much of sedimentary δ15N record with an average value closer to 0‰, suggesting that the Archaean N cycle was dominated by biological N fixation utilising the molybdenum-iron nitrogenase co-factor from 3.2 Ga onwards5 with limited evidence for aerobic N cycling and the presence of nitrate until the GOE6,7. A largely anoxic Archaean world is consistent with evidence from geochemical redox proxies that indicate only transiently elevated oxygen levels in shallow marine environments811. The NIE model is also difficult to reconcile with highly negative δ15N values in deep-water shales from the ca. 2.75 Ga Manjeri Formation, Zimbabwe craton, which are explained in terms of partial N assimilation into biomass from a deep, ammonium-rich reservoir1. This explanation requires an isotopically heavy N sink that has thus far remained elusive, rendering the isotopic mass balance and our understanding of the early N cycle prior to the GOE incomplete.

The Manjeri Fm (Ngezi Group, Belingwe greenstone belt) is a relatively thin (ca. 100 m) sedimentary succession that was deposited rapidly during a marine transgression of the Zimbabwe proto-craton due to crustal extension and subsidence, which was immediately followed by extensive submarine mafic and ultramafic volcanism associated with the overlying Reliance and Zeederbergs formations at 2.746 ± 0.004 Ga1214 (Fig. 1). This period correlates with globally enhanced rates of mafic-ultramafic volcanic activity at ~2.75 Ga14,15 associated with convective mantle overturning16,17. The sedimentary Cheshire Formation was deposited following this volcanic episode, which is temporally constrained by the eruption of the Ngezi volcanics at 2.75 Ga and the crystallisation of a dolerite intrusion at ~2.71 Ga14. Although the Manjeri and Cheshire Fm carbonates exhibit positive Eu anomalies (Eu/Eu* = (Eu/0.67Sm + 0.33Tb); defined as Eu/Eu* > 1) up to 6.1, this is regarded as a primary feature of the local ambient seawater rather than reflecting close proximity to a hydrothermal vent18. Positive Eu anomalies are a common feature of detrital-free Archaean chemical sedimentary rocks and can be attributed to a Eu excess under reducing conditions in Neoarchean seawater due to increased hydrothermal activity globally19. Moreover, the Manjeri and Cheshire Fm carbonates exhibit 87Sr/86Sr ratios as low as 0.7015518, which is similar to that of the depleted mantle value at the time of deposition and indicates that carbonate sedimentation around the Zimbabwe proto-craton occurred co-eval to abundant submarine greenstone volcanism and prior to the stabilisation of most Archaean cratons18.

Fig. 1. Geological context of the Manjeri and Cheshire formations.

Fig. 1

a Regional setting of the Zimbabwe craton. b Simplified geological map showing the Archaean granitoid (yellow shaded areas) and greenstone terrain (green shaded areas) of the Zimbabwe craton. c Generalised stratigraphy of the Bulawayan Supergroup of the Belingwe greenstone belt. Figure modified from ref. 61.

Here, we investigate shallow marine limestones from the ~2.75 Ga Huntsman Quarry (Bulawayo greenstone belt), which is stratigraphically correlated with the Manjeri Fm18, and younger shallow marine limestones from the ~2.73 Ga Cheshire Fm. We focus on well-preserved carbonates with limited metamorphic alteration (lower greenschist facies metamorphism), which were screened according to their previously reported18 stable carbon (C) and oxygen (O) isotope compositions (δ13C and δ18O), and radiogenic strontium isotope (87Sr/86Sr) ratios (Table 1). We combine these existing data with measurements of bulk N and organic C isotopes (δ15Nbulk and δ13Corg), and also analyse uranium (U) isotopes to assess the redox conditions of the depositional environment.

Table 1.

Summary of sample descriptions and geochemical data for the Manjeri Fm and Cheshire Fm carbonates

Sample Unit Description CaO (wt%) MgO (wt%) δ13C
(‰)
δ18O
(‰)
87Sr/86Sr Y/Ho Ce/Ce* Eu/Eu* U
(ppb)
Th
(ppb)
Th/U δ238U
(‰)
δ234U
(‰)
Z04-17A Manjeri Fm Stromatolitic limestone 54.1 0.9 0.53 −20.32 0.701613 104.5 0.62 2.96 95 6 0.1 −0.28 ± 0.02 57.5 ± 0.4
Z04-17B Stromatolitic limestone 55.5 0.1 1.96 −16.88 0.701571 156.5 0.48 6.10 53 <1 <0.1 −0.31 ± 0.07 297.3 ± 12.8
Z04-17C Stromatolitic limestone n/a n/a 0.56* −20.52* 0.701593 113.7 0.58 2.72 35 8 0.22 −0.46 ± 0.05 32.1 ± 1.6
Z04-17D Stromatolitic limestone 46.8 1.4 −0.09 −18.49 0.701877 86.5 0.66 1.84 157 18 0.11 n/a n/a
Z04-17E Stromatolitic limestone 53.0 0.8 −0.30* −15.27* 0.701679 103.3 0.63 3.00 34 <1 <0.1 −0.52 ± 0.16 158.8 ± 11.8
Z04-17F Stromatolitic limestone 54.2 0.6 −0.54 −16.02 0.703631 60.1 0.90 1.73 880 33 <0.1 n/a n/a
Z04-27 Sheared limestone 50.6 0.4 0.62 −10.43 ‰0.703398 84.1 0.66 4.37 19 <1 <0.1 −0.43 ± 0.03 8.5 ± 0.9
BD19 Cheshire Fm Limestone 28.1 2.9 −1.11* −13.73* 0.710222 24.0 0.93 1.26 218 202 0.93 −0.10 ± 0.15 241.5 ± 10.4
BD33 Limestone n/a n/a −0.61* −13.35* n/a n/a n/a n/a n/a n/a n/a −0.22 ± 0.07 200.8 ± 7.1
BD37 Limestone 38.0 2.2 −1.97 −16.26 0.712371 28.5 0.89 2.29 33 96 2.9 −0.26 ± 0.07 243.5 ± 2.7
BD43 Limestone n/a n/a −1.24 −16.25 n/a n/a n/a n/a n/a n/a n/a −0.32 ± 0.03 130.1 ± 2.0
BD46 Limestone n/a n/a −0.47* −13.89* n/a 63.1 0.75 2.42 13 16 1.2 −0.56 ± 0.16 134.0 ± 6.3
BD49 Limestone n/a n/a −1.63* −15.02* n/a 37.5 0.89 1.80 89 190 2.1 −0.41 ± 0.01 203.9 ± 4.6
BD50 Limestone 46.8 0.3 0.06 −13.85 n/a 29.4 0.97 1.61 20 198 10.1 n/a n/a
BD51 Limestone 44.5 0.7 −0.11 -15.20 0.712061 n/a n/a n/a n/a n/a n/a n/a n/a
BD52 Stromatolitic limestone n/a n/a −0.39* -13.91* n/a n/a n/a n/a n/a n/a n/a n/a n/a

Major element, strontium (Sr) isotope and rare-earth element data sourced from Hofmann et al.18. Additional stable isotope data (δ13C and δ18O) measured in this study are indicated by asterisks (*) whereby all other data are from Hofmann et al.18. Uncertainty values for δ238U and δ234U represent 2 standard error.

Results

Bulk nitrogen and organic carbon isotope data

The δ15Nbulk values of the Cheshire and Manjeri formation carbonates are strongly positive, ranging from +26.0‰ to +35.0‰ (average = +31.1 ± 2.5‰; 1σ; n = 12) and +28.1‰ to +42.5‰ (average = +37.5 ± 6.5‰; 1σ; n = 4), respectively, with δ13Corg values ranging from −39.1‰ to −29.7‰ (Fig. 2a; Table 2). We infer a predominantly organic origin for N in our samples due to the correlation between TN and TOC contents for all samples (R2 = 0.99), which is also found when only considering the Manjeri Fm (R2 = 0.99; Fig. 2b). Although a weaker relationship is found for the Cheshire Fm (R2 = 0.37), this is likely because the latter are generally more silicified than the Manjeri Fm18. Furthermore, there is an outlier (sample BD49-E-C5), which exhibits anomalously higher TOC contents (2.28%) compared to the average value for the Cheshire Fm (1.00 wt%); excluding this sample results in a stronger relationship (R2 = 0.59). The C/N ratios range from 68 to 231 mol/mol for the Cheshire Fm and 71 to 82 mol/mol for the Manjeri Fm (Fig. 2c). This range overlaps with that measured in the Serra Sul Fm, which also exhibits similarly positive N isotope values4, but it is somewhat lower than the Tumbiana Fm samples, which exhibit very elevated C/N values up to 589 mol/mol2.

Fig. 2. Nitrogen (δ15N) and organic carbon isotope (δ13Corg) data for the Manjeri Fm and Cheshire Fm carbonates.

Fig. 2

a Plot of δ15N vs δ13Corg including data from Manjeri Fm shales1 (red-filled diamonds) and Cheshire Fm shales45 (green shaded area above plot), whereby grey-filled bars plotted outside the axes show the range of different metabolisms62. b Plot of TN vs TOC contents for the decarbonated residues. c Plot of δ15N vs C/N ratios with expected range shown for regional metamorphism effects21. Error bars represent 1σ and those not shown are smaller than the marker symbol. Source data are provided as a Source Data file. Purple- and green-filled diamonds represent data from the Manjeri Fm and Cheshire Fm carbonates, respectively.

Table 2.

Nitrogen and organic carbon isotope data for the Manjeri Fm and Cheshire Fm carbonates

Sample ID Unit TNdecarb
(ppm)a
±1σ
(ppm)
δ15Nbulk (‰) ±1σ
(‰)
TOCdecarb (%)a ±1σ
(%)
δ13Corg (‰) ±1σ
(‰)
C/N (mol/mol)
BD49-E-C1

Cheshire

Fm

137 10 29.5 0.5 1.29 0.12 −36.85 0.13 110
BD49-E-C2 116 16 33.1 1.3 1.25 0.08 −37.58 0.12 126
BD49-E-C3 74 3 33.2 0.6 0.99 0.03 −39.12 0.05 157
BD49-E-C4 73 10 32.1 2.4 0.84 0.04 −38.39 0.13 133
BD49-E-C5 115 8 35.0 0.5 2.28 0.21 −38.01 0.02 231
BD49-E-C6 136 12 29.6 3.8 0.89 0.05 −36.37 0.12 76
BD49-E-C7 116 12 29.9 3.5 1.08 0.06 −37.63 0.06 109
BD49-E-C8 123 2 33.4 0.6 1.06 0.12 −37.90 0.05 100
BD52-B-C1 42 2 26.0 2.0 0.24 0.01 −31.58 0.07 68
BD52-B-C2 111 10 31.8 4.3 0.86 0.05 −31.06 0.17 91
BD52-B-C3 46 5 29.2 3.2 0.32 0.00 −30.76 0.02 81
BD52-B-C4 55 n/a 30.1 n/a 0.92 0.55 −29.67 0.61 119
ZO4-17C-G

Manjeri

Fm

5826 166 38.8 <0.1 39.03 1.13 −31.87 0.17 78
ZO4-17C-E-2 5121 47 42.5 0.2 35.79 2.36 −31.14 0.02 82
ZO4-17C-F-2 2342 n/a 28.1 n/a 14.17 0.09 −32.99 0.07 71
ZO4-17C-D 5842 229 40.7 0.9 38.45 1.25 −30.50 0.03 77

a: TNdecarb and TOCdecarb values represent the contents of the decarbonated residues.

Uranium isotope compositions

The uranium isotope composition (238U/235U, expressed as δ238U, see Eq. 1) of the Cheshire Fm and Manjeri Fm carbonates range from −0.56 ± 0.16‰ to −0.10 ± 0.15‰ (2 standard error; 2 s.e.) and δ234U values (see Eq. 2) range from +8.5 ± 0.9‰ to +297.3 ± 12.8‰ (2 s.e.; Table 1). There is a weak correlation between δ238U and the 234U/238U activity ratios (expressed relative to secular equilibrium as ‘δ234U’, where δ234Usec.eq. = 0 ‰) for the Cheshire Fm (R2 = 0.48) and no correlation for the Manjeri Fm carbonates (R2 = 0.06; Fig. 3a).

Fig. 3. Uranium isotope values and Y/Ho ratios18 of the Cheshire Fm and Manjeri Fm carbonates.

Fig. 3

a Plot of δ238U vs δ234U. b Plot of δ238U vs Y/Ho ratios. c box plots of samples with low (<29) and high (>37) Y/Ho ratios where the centre line shows the median, box limits show the upper and lower quartiles, and whiskers show 1.5 times the interquartile range. Error bars in panels b and c represent 2 standard error and those not shown are smaller than the marker symbol. Source data are provided as a Source Data file. Purple- and green-filled diamonds represent data from the Manjeri Fm and Cheshire Fm carbonates, respectively.

Discussion

The δ15Nbulk values of the Manjeri and Cheshire formations are strongly elevated in comparison to the mean value of ca. 0‰ in Archaean sedimentary rocks5,7,20. Although there is a positive relationship between δ15Nbulk and C/N ratios for both the Manjeri and Cheshire formations (Fig. 2c) that suggests some N devolatilization related to regional metamorphism, this cannot explain the highly positive δ15N values. This is because the metamorphic grade of our samples is low (lower greenschist facies) and the isotopic fractionation factor (ε) determined for greenschist-facies metamorphism is relatively small (1.5 ± 1‰)21, which can only yield values up to ca. +10‰ for a starting composition of 0‰. Moreover, there is no evidence for proximal hydrothermal processes that may have resulted in large isotope fractionation factors associated with N release from minerals at ~300 °C2224. Therefore, the primary N isotope signature of the Manjeri Fm and Cheshire Fm carbonates was likely in excess of +20‰, consistent with other examples of the NIE globally24.

Our positive δ15Nbulk values may be complementary to negative δ15N values down to −11 ‰ reported from the deep-water shales in the Manjeri Fm1, which have been explained in terms of partial ammonium assimilation from a large, deep-water ammonium reservoir. Assuming the median ε value of −14‰ for this metabolism would imply that ca. 70–90% of the dissolved ammonium pool was removed via biomass assimilation25 (Fig. 4). A necessary outcome of this hypothesis proposed by Yang et al. is the generation of a residual ammonium pool in seawater with a positive N isotope composition ranging from +18‰ to +34‰. This predicted range agrees well with our carbonate δ15N values. The highest δ15N values may also be explained by partial ammonium oxidation, as recently suggested to explain the NIE4, which implies the availability of free oxygen in the water column.

Fig. 4. The modelled residual ammonium pool following assimilation into the biomass in the deep basin and upwelling onto the shallow proto-cratonic shelf (modified from Yang et al.1).

Fig. 4

Isotope fractionation factors represent experimentally derived values from Hoch et al.25 and the grey shaded area represents the likely range for 1 − f (following Yang et al.1). Box plots show the range of measured values in the various sedimentary facies where the centre line shows the median, box limits show the upper and lower quartiles, whiskers show 1.5 times the interquartile range. Source data are provided as a Source Data file.

Similar to most Archaean carbonates26, shallow marine carbonates from the Manjeri and Cheshire formations lack true negative, shale-normalised (SN) Ce anomalies (Ce/Ce*SN = Ce/(0.5La + 0.5Pr), defined as (Ce/Ce*SN < 1), but δ238U values as low as −0.56‰ (Table 1) indicate subtle redox variations during their deposition. The lowest δ238U are considered the most reliable (maximum) estimates for the seawater δ238U value at the time of deposition because the heavy U isotope, 238U, is preferentially reduced during early sedimentary diagenesis, as demonstrated in modern Shark Bay stromatolites27 and other modern shallow marine carbonates28. Therefore, only δ238U values higher than the modern seawater value (−0.4‰) may reasonably be attributed to sedimentary diagenesis. As samples exhibit 234U–238U disequilibrium with δ234U values up to ca. +300 ‰ (Table 1), this indicates some degree of U mobility within the past ca. 1.5 Ma. However, the weak correlation between δ238U and δ234U (Fig. 3a) indicates that post-depositional alteration by weathering fluids had only a limited effect on the δ238U redox proxy. In any case, this correlation further supports the interpretation that samples with the lowest δ238U likely provide the most reliable information regarding the primary sedimentary signature.

The reliability of the δ238U redox proxy in ancient carbonates can be further examined by considering the rare-earth element + yttrium (REY) patterns. For instance, the Y and holmium (Ho) elemental ratio, which remains constant at the chondritic Y/Ho ratio of 26 to 28 during most geological processes but is fractionated in aqueous marine environments29,30. This results in modern seawater exhibiting a superchondritic Y/Ho ratio (>28) that is considered to represent a primary seawater signal in ancient carbonates31,32. Despite Y/Ho and U isotope representing two different chemical systems, they may record signals from the same source, i.e. a primary seawater signal, as both U isotope and REY signatures are typically well preserved during carbonate diagenesis33. A two-tailed t-test reveals that carbonates from the Manjeri and Cheshire formations with lower Y/Ho (<28) exhibit significantly higher δ238U values (mean ± 1 s.d. = −0.23 ± 0.09 ‰) than samples with Y/Ho greater than 37 (p = 0.01; mean ± 1 s.d. = −0.42 ± 0.12 ‰). This is consistent with a greater influence of detrital material for carbonates with higher δ238U and lower Y/Ho. According to the δ238U values of samples with a Y/Ho greater than 37, our data indicate that the local Neoarchean seawater that covered the Zimbabwe proto-craton possibly varied from a modern-like δ238U value of −0.4‰34 to a minimum of around -0.6‰ (Fig. 3b, c). Although these δ238U values overlap with the average value for modern open seawater (ca. −0.4‰34), we do not propose that the average oxidation state of the Neoarchean Ocean was similar to present. Lower δ238U values relative to modern seawater have also been interpreted to represent the onset of mildly oxidative weathering in other Precambrian sedimentary rocks35,36, whereas in Phanerozoic sedimentary rocks, lower δ238U values are typically associated with a relative increase in the extent of seafloor anoxia and preferential reduction of 238U (refs. 37,38). We stress that the main significance of lower δ238U in Neoarchean carbonates is the implied presence of oxidised U6+ in the water column. Mildly oxidising redox conditions may be reasonable if the local levels of dissolved oxygen in oxygen oases were related to the productivity of oxygen-producing cyanobacteria39, which would be regulated by the supply of dissolved nutrients delivered from the upwelling of deep waters.

The lack of true Ce anomalies despite variations in U isotopes may be reconciled by considering the redox potentials of these elements under aqueous conditions, whereby Ce3+ is oxidised to Ce4+ by oxygen at ca. +1 V at a circumneutral pH with low Ce3+ concentrations found in seawater40, whereas the two-step oxidation of U4+ to U6+ may occur at less oxidising conditions up to +0.3 V (ref. 41), which may constrain the upper limit of the redox potential in the shallow marine environment. Importantly, ammonium oxidation can occur at a redox potential of around +0.4 V under a circumneutral pH42, which lies between the reduction potentials of Ce4+/Ce3+ and U6+/U4+. Thus, the strongly elevated δ15N values suggest that the marine redox environment was at least transiently oxidising enough for ammonium oxidation to occur. This is plausible given that both modelling43 and laboratory experiments39 with cyanobacteria suggest that oxygenic photosynthesis can locally yield dissolved oxygen concentrations up to ~10 μM under Archaean conditions. However, molecular clock estimates suggest that most modern clades of ammonium oxidising bacteria and archaea emerged after the GOE in the Paleoproterozoic44. The low δ13Corg of the shallow-marine carbonates from Cheshire and Manjeri Fm (−39.1‰ to −29.7‰; Table 2) compared to the higher δ13Corg values in the deep-marine Manjeri Fm shales1,45 (Fig. 2a) suggest that methanotrophy was restricted to shallow waters around the Zimbabwe proto-craton, possibly due to limited sulfate availability in deeper, more reducing waters. As there is a large isotope fractionation effect associated with methanotrophic ammonium oxidation to N2O, this may yield elevated δ15N values46, especially in combination with the upwelling of a 15N-rich pool of residual ammonium due to biological ammonium assimilation in the deep basin1.

Positive δ15Nbulk values coupled to low δ13Corg values may provide a signature of ammonium oxidation by methanotrophs under an abundant supply of methane in a hydrothermally influenced setting. Hydrothermal fluids are typically rich in ammonium (and methane) when circulating through sediment-covered oceanic ridges with modern hydrothermal fluids exhibiting elevated concentrations up to 16 mM4749. In the Neoarchean, some hydrothermal vent fluids also had sufficiently high ammonium concentrations to facilitate the partial N utilisation by abiotic or biotic processes, such as those in the 2.7 Ga Abitibi basin22. Ammonium may also be sourced via the remineralisation of organic matter in marine sediments in the absence of hydrothermal fluid flow as, for instance, in the modern Black Sea. However, this process is inconsistent with our N isotope data because nutrient-rich reservoirs that accumulate due to intense basin stratification imply a limited upwelling of nutrients to the surface50. In contrast, buoyant hydrothermal vent plumes, which are typically warmer and less dense than the surrounding seawater, enable limited nutrients such as Fe to reach surface waters and stimulate primary productivity in the modern oceans on timescales of ~100 yr51,52. A similar scenario could therefore be plausible for hydrothermal ammonium in the Archaean. As ammonium may have been a limiting nutrient in the Neoarchean Ocean, it was probably rapidly scavenged from the water column, preferentially removing 14N via ammonium assimilation and archiving negative N isotope signature in the deep-water shales1. The remaining pool of 15N-rich hydrothermal ammonium could then have reached the shallow marine environment via upwelling and fuelled surface biological productivity in the form of microbial communities associated with stromatolites under weakly oxidising conditions due to oxygenic photosynthesis by cyanobacteria (Fig. 5).

Fig. 5. Conceptual model of the submerged Zimbabwe proto-craton at ca.

Fig. 5

2.75 Ga to explain the coupled positive and negative nitrogen isotope values in terms of hydrothermal ammonium upwelling. Hydrothermal fluids rich in dissolved ammonium (NH4+) and other key nutrients are released in the deep basin and accumulate in the deeper waters, which are assimilated by biological organisms and produce negative nitrogen isotopes values in deep water sediments that eventually form shale rocks. The remaining dissolved NH4+ that reaches the surface due to upwelling processes is enriched in 15N and archived in shallow-water carbonates.

Overall, the occurrence of strongly positive δ15N in the ca. 2.75 Manjeri Fm and 2.73 Ga Cheshire Fm in the Zimbabwe craton further supports the global nature of the NIE, which includes the 2.68 Ga Serra Sul Fm (Amazonian craton)4 and 2.72 Ga Tumbiana Fm (Pilbara craton)2,3. The link between elevated δ15N values in shallow water carbonates and unusually negative δ15N values in deep-water shales via ammonium assimilation may also explain both the extremely high values and greater variability of δ15N values during the NIE. Furthermore, we show that the NIE began at least some 30 Ma earlier at around 2.75 Ga (Fig. 6), which temporally correlates with globally enhanced volcanism caused by increased magmatic production associated with mantle overturning14,16,17. Although global conditions during the Neoarchean were primed to supply enhanced fluxes of ammonium from hydrothermal vents in marine environments, not all places would have received enough ammonium to produce those extremely positive δ15N values. However, distal hydrothermal fluxes of recycled ammonium were likely supplemented by key nutrients like methane and dissolved P (ref. 53) that could have triggered biological productivity during this time. In addition, enhanced fluxes of biologically useful transition metals such as copper, molybdenum and zinc22,54 may have simultaneously catalysed novel biological diversification, thereby triggering the necessary conditions for the onset of the NIE and the expansion of life in the buildup to the GOE.

Fig. 6.

Fig. 6

Nitrogen isotope data (δ15N) for well-preserved (sub-greenschist to lower-greenschist facies) Neoarchean sedimentary rocks deposited between 2.80 and 2.45 Ga. Data were compiled by Stüeken et al.63 and supplemented with additional data from the Manjeri Fm and Cheshire Fm carbonates (this study), Manjeri Fm shales1 and the Serra Sul Fm4. Source data are provided as a Source Data file. Filled diamonds and circles represent δ15Nbulk and δ15Nker data from various localities deposited during the Nitrogen Isotope Event, respectively, where purple: Manjeri Fm carbonates, green: Cheshire Fm carbonates, red: Manjeri Fm shales, pink: Serra Sul Fm, and blue: Tumbiana Fm and Kylena Fm.

Methods

Stable isotope analyses

Nitrogen and carbon isotope values (δ15Nbulk and δ13Corg) and their abundances of carbonate powders were analysed at the University of St Andrews (as previously described in ref. 2). Samples were first decarbonated by heating with 2 M HCl (reagent grade) at 70 °C overnight and centrifuged to remove the acid. Before drying in a closed oven, samples were washed three times with 18.2 MΩ/cm DI-H2O to remove acidic residues. An appropriate amount of dry sample residue was then weighed into tin capsules and analysed with an elemental analyser for flash combustion (EA-IsoLink) coupled to a continuous-flow isotope-ratio mass spectrometer (MAT253 CF-IRMS) via a Conflo IV (all Thermo Fisher)55. Isotopic values were calibrated with the international reference materials USGS-40 and USGS-41 with USGS-62 as a secondary standard, which yielded average δ15Nbulk and δ13Corg values of +20.26 ± 0.18‰ (n = 17; 1σ) and δ13Corg = −14.77 ± 0.07‰ (n = 6; 1σ), respectively, which are consistent with previously reported values of +20.17‰ and −14.79‰, respectively. Devonian shale SDo-1 was also processed through the entire procedure and yielded average δ15Nbulk and δ13Corg values of −0.44‰ ± 0.24‰ (n = 4; 1σ) and −30.16 ± 0.29‰ (n = 4; 1σ), respectively. Fourteen carbonate samples were analysed in duplicate, yielding average reproducibilies of ±1.70‰ and ±0.09‰ for δ15Nbulk and δ13Corg (1σ), respectively. The TN and TOC contents of the decarbonated residues (TNdecarb and TOCdecarb) were determined from peak areas of the IRMS analysis and calibrated with a series of USGS-41 measurements. Carbonate contents of each sample were estimated by weighing an aliquot of powder before and after treatment with 2 M HCl. Where we report total concentrations, decarbonated denotes the siliciclastic fraction remaining after treatment with HCl, as reported by Thomazo et al.3. Isotopic ratios are reported relative to atmospheric air for δ15Nbulk and VPDB for δ13Corg.

Stable carbon and oxygen isotope values (δ13C and δ18O) of carbonate powders were analysed following methods previously described in Pallacks et al.56 on a Thermo Delta V mass spectrometer equipped with a GASBENCH-II preparation device at the Max Planck Institute for Chemistry. Approximately ~20 to 50 μg of CaCO3 sample was placed in a He-filled 12 ml exetainer vial and digested in water-free H3PO4 at a temperature of 70 °C. Subsequently, the CO2–He gas mixture is transported to the GASBENCH in Helium carrier gas. In the GASBENCH, water vapour and various gaseous compounds are separated from the He-CO2 mixture prior to sending it to the mass spectrometer. Isotope values are reported as δ13C and δ18O values relative to Vienna Pee Dee Belemnite (VPDB). A total of 20 replicates of two in-house CaCO3 standards are analysed in each run of 55 samples. CaCO3 standard weights are chosen so that they span the entire range of sample weights of the samples. After correction of isotope effects related to sample size, the reproducibility of these standards typically is better than 0.1‰ (1σ) for δ18O and for δ13C.

Uranium isotope measurements

Uranium isotope measurements were conducted following methods previously described in Martin et al.27 and are briefly given here. Depending on sample availability and previously measured U concentrations, approximately ~300–1000 mg of stromatolite powder was leached with 20 mL 2 M HCl at room temperature for 24 h. The samples were centrifuged and the solutions were retained for analyses. Prior to column chromatography, the samples were evaporated at 80 °C to incipient dryness and a U double spike (IRMM-3636a)57 was added to the samples, targeting a 236U/235U of ~3 and a molar U sample-spike ratio of ~20–25. To separate U from the carbonate matrix, column chromatography was conducted according to Weyer et al.58 using the Eichrom UTEVA resin and 150–300 ng U was typically loaded. Following column chromatography, 0.1 mL HNO3 (65%) and 0.1 mL H2O2 (30%) were added and evaporated at 80 °C to incipient dryness. The residue was then redissolved in 3% (v/v) HNO3 to yield final solutions with U concentrations ranging from 50 to 100 ppb.

Isotopic measurements were conducted using a Thermo Scientific™ Neptune Plus™ in low-resolution mode with a Cetac Aridus 2 sample introduction system (dry plasma conditions) at LUH following Noordman et al. 38. A standard Ni H sampler cone and X skimmer cone setup typically achieved >1 V/ppb sensitivity for 238U. The 233U, 235U and 236U isotopes were measured using Faraday detectors with 1011 Ω resistors and 238U was measured with a 1010 Ω resistor whereby 234U isotope was measured with a 1013 Ω resistor. The abundance sensitivity of 238U on 236U was monitored to ensure it was <1 ppm. Instrumental mass bias was corrected using the 233U/236U ratio according to the exponential law. Measurement sequences were performed using a standard-sample-bracketing method relative to a CRM-112A standard solution to calculate δ238U (Eq. 1) and δ234U are given according to Eq. 2 relative to the secular equilibrium (SE) of 234U/238U = 54.891 ± 0.094 × 106 (2σ)59. Uranium isotope ratios are reported according to convention using delta notation (in ‰), given as:

δ238U=[(U238/U235)sample/(U238/U235)CRM112A1]*1000 1
δU234=[(U234/U238)sample/(U234/U238)s.e.1]*1000 2

All δ-values of samples represent triplicate measurements where uncertainty values represent 2σ standard error (2 s.e.) for both δ238U and δ234U. Reference materials were measured throughout the measurement sequence to monitor the instrument performance and a limestone (JLs; Geological Survey of Japan) was also processed with each batch of samples for column chromatography. The average δ238U values of IRMM-184, Reimep-18a and JLs were −1.17 ± 0.04‰ (2σ, n = 9), −0.25 ± 0.07‰ (2σ, n = 9), and −0.36 ± 0.08‰ (2σ, n = 3), respectively, and their average δ234U values were −28.0 ± 1.0‰, 34.4 ± 2.3‰, and 34.3 ± 0.5‰, which are all consistent with reported values60. Total procedure blanks from leaching and column chromatography were <4 ng and no blank corrections were applied to the data.

Supplementary information

Peer Review File (600.5KB, pdf)

Source data

Source Data (24.9KB, xlsx)

Acknowledgements

Funding for A.N.M. and S.W. (WE 2850/17-1), and M.M.G. (GE 2558/4-1) was provided by the German Research Foundation (DFG) priority programme SPP-1833 Building a Habitable Earth. A.N.M. was supported by additional funding from the DFG priority programme SPP-2238 Dynamics of Ore Metals Enrichment (MA 9571-3-1). E.E.S. acknowledges support from a UK Natural Environment Research Council (NERC) Frontiers grant (NE/V010824/1) and a Leverhulme Trust grant (RPG-2022-313). M.M. acknowledges funding from a Royal Society Award (URF\R1\231546). A.H. acknowledges logistical support from the Department of Geology, University of Zimbabwe.

Author contributions

A.N.M. and E.E.S. performed the nitrogen isotope measurements. A.N.M. conducted the uranium isotope analyses. M.M. and H.V. obtained inorganic carbon and oxygen stable isotope data. A.N.M., M.M.G., E.E.S., S.W. & A.H. developed the concept and designed the experimental approach. A.N.M. wrote the initial draft. All authors contributed to reviewing and editing the paper at all stages.

Peer review

Peer review information

Nature Communications thanks Genming Luo and Magali Ader for their contribution to the peer review of this work. A peer review file is available.

Data availability

The stable isotope data generated in this study for nitrogen, organic carbon and uranium are provided in the Source Data file and uploaded in a Figshare repository (10.6084/m9.figshare.27632010). Source data are provided with this paper.

Competing interests

The authors declare no competing interests.

Ethics

We affirm that all geological materials were collected in a responsible manner and in accordance with relevant permits and local laws. A local researcher (Axel Hofmann) collected the samples and is a co-author of this paper. Local and regional research relevant to this study has been cited where appropriate.

Footnotes

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

The online version contains supplementary material available at 10.1038/s41467-025-57091-3.

References

  • 1.Yang, J. et al. Ammonium availability in the Late Archaean nitrogen cycle. Nat. Geosci.12, 553–557 (2019). [Google Scholar]
  • 2.Stüeken, E. E., Buick, R. & Schauer, A. J. Nitrogen isotope evidence for alkaline lakes on late Archean continents. Earth Planet. Sci. Lett.411, 1–10 (2015). [Google Scholar]
  • 3.Thomazo, C., Ader, M. & Philippot, P. Extreme 15N-enrichments in 2.72-Gyr-old sediments: evidence for a turning point in the nitrogen cycle. Geobiology9, 107–120 (2011). [DOI] [PubMed] [Google Scholar]
  • 4.Pellerin, A. et al. Neoarchaean oxygen-based nitrogen cycle en route to the Great Oxidation Event. Nature10.1038/s41586-024-07842-x (2024). [DOI] [PubMed]
  • 5.Stüeken, E. E., Buick, R., Guy, B. M. & Koehler, M. C. Isotopic evidence for biological nitrogen fixation by molybdenum-nitrogenase from 3.2 Gyr. Nature520, 666–669 (2015). [DOI] [PubMed] [Google Scholar]
  • 6.Zerkle, A. L. et al. Onset of the aerobic nitrogen cycle during the Great Oxidation Event. Nature542, 465–467 (2017). [DOI] [PubMed] [Google Scholar]
  • 7.Mettam, C. et al. Anaerobic nitrogen cycling on a Neoarchaean ocean margin. Earth Planet. Sci. Lett.527, 115800 (2019). [Google Scholar]
  • 8.Eickmann, B. et al. Isotopic evidence for oxygenated Mesoarchaean shallow oceans. Nat. Geosci.11, 133–138 (2018). [Google Scholar]
  • 9.Ossa Ossa, F. et al. Aerobic iron and manganese cycling in a redox-stratified Mesoarchean epicontinental sea. Earth Planet. Sci. Lett.500, 28–40 (2018). [Google Scholar]
  • 10.Planavsky, N. J. et al. Evidence for oxygenic photosynthesis half a billion years before the Great Oxidation Event. Nat. Geosci.7, 283–286 (2014). [Google Scholar]
  • 11.Anbar, A. D. et al. A whiff of oxygen before the Great Oxidation Event? Science317, 1903–1906 (2007). [DOI] [PubMed] [Google Scholar]
  • 12.Prendergast, M. D. & Wingate, M. T. D. Zircon geochronology of late Archean komatiitic sills and their felsic country rocks, south-central Zimbabwe: a revised age for the Reliance komatiitic event and its implications. Precambrian Res.229, 105–124 (2013). [Google Scholar]
  • 13.Prendergast, M. D. The Bulawayan Supergroup: a late Archaean passive margin-related large igneous province in the Zimbabwe craton. J. Geol. Soc.161, 431–445 (2004). [Google Scholar]
  • 14.Hofmann, A., Li, X., Bradley, D. C. & Humbert, F. The Belingwe greenstone belt, Zimbabwe—some new age constraints and implications for greenstone belt evolution. Int. Geol. Rev. 10.1080/00206814.2024.2309477 (2024).
  • 15.Kump, L. R. & Barley, M. E. Increased subaerial volcanism and the rise of atmospheric oxygen 2.5 billion years ago. Nature448, 1033–1036 (2007). [DOI] [PubMed] [Google Scholar]
  • 16.Griffin, W. L. et al. The world turns over: Hadean–Archean crust–mantle evolution. Lithos. 189, 2–15 (2014). [Google Scholar]
  • 17.Campbell, I. H. & Hill, R. I. A two-stage model for the formation of the granite-greenstone terrains of the Kalgoorlie-Norseman area, Western Australia. Earth Planet. Sci. Lett.90, 11–25 (1988). [Google Scholar]
  • 18.Hofmann, A. et al. A craton-wide geochemical study of Neoarchaean carbonate rocks of Zimbabwe. Chem. Geol.634, 121587 (2023). [Google Scholar]
  • 19.Viehmann, S., Bau, M., Hoffmann, J. E. & Münker, C. Geochemistry of the Krivoy Rog Banded Iron Formation, Ukraine, and the impact of peak episodes of increased global magmatic activity on the trace element composition of Precambrian seawater. Precambrian Res.270, 165–180 (2015). [Google Scholar]
  • 20.Ossa, F. O. et al. Limited oxygen production in the Mesoarchean ocean. Proc. Natl Acad. Sci. USA116, 6647–6652 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Bebout, G. E. & Fogel, M. L. Nitrogen-isotope compositions of metasedimentary rocks in the Catalina Schist, California: Implications for metamorphic devolatilization history. Geochim. Cosmochim. Acta56, 2839–2849 (1992). [Google Scholar]
  • 22.Martin, A. N. et al. Mechanisms of nitrogen isotope fractionation at an ancient black smoker in the 2.7 Ga Abitibi greenstone belt, Canada. Geology52, 181–186 (2024). [Google Scholar]
  • 23.Scalan, R. S. The Isotopic Composition, Concentration, and Chemical State of the Nitrogen in Igneous Rocks. (University of Arkansas, Fayetteville, 1958).
  • 24.Hanschmann, G. Berechnung von Isotopieeffekten auf Quantenchemischer Grundlage am Beispiel stickstoffhaltiger Molecule. Zfl-Mitteilungen41, 19–39 (1981). [Google Scholar]
  • 25.Hoch, P. & Kirchman, D. L. Isotope fractionation associated with ammonium uptake by a marine bacterium. Limnol. Oceanogr.37, 1447–1459 (1992). [Google Scholar]
  • 26.Zhang, K. & Shields, G. A. Sedimentary Ce anomalies: secular change and implications for paleoenvironmental evolution. Earth-Sci. Rev. 229, 104015 (2022). [Google Scholar]
  • 27.Martin, A. N., Markowska, M., Chivas, A. R. & Weyer, S. Assessing the reliability of modern marine stromatolites as archives for the uranium isotope paleoredox proxy. Geochim. Cosmochim. Acta345, 75–89 (2023). [Google Scholar]
  • 28.Romaniello, S. J., Herrmann, A. D. & Anbar, A. D. Uranium concentrations and 238U/235U isotope ratios in modern carbonates from the Bahamas: assessing a novel paleoredox proxy. Chem. Geol.362, 305–316 (2013). [Google Scholar]
  • 29.Nozaki, Y., Zhang, J. & Amakawa, H. The fractionation between Y and Ho in the marine environment. Earth Planet. Sci. Lett.148, 329–340 (1997). [Google Scholar]
  • 30.Bau, M. & Dulski, P. Comparing yttrium and rare earths in hydrothermal fluids from the Mid-Atlantic Ridge: implications for Y and REE behaviour during near-vent mixing and for the Y/Ho ratio of Proterozoic seawater. Chem. Geol.155, 77–90 (1999). [Google Scholar]
  • 31.Kamber, B. S., Webb, G. E. & Gallagher, M. The rare earth element signal in Archaean microbial carbonate: information on ocean redox and biogenicity. J. Geol. Soc.171, 745–763 (2014). [Google Scholar]
  • 32.Kamber, B. S. & Webb, G. E. The geochemistry of late Archaean microbial carbonate: implications for ocean chemistry and continental erosion history. Geochim. Cosmochim. Acta65, 2509–2525 (2001). [Google Scholar]
  • 33.Hood, Av. S., Planavsky, N. J., Wallace, M. W. & Wang, X. The effects of diagenesis on geochemical paleoredox proxies in sedimentary carbonates. Geochim. Cosmochim. Acta232, 265–287 (2018). [Google Scholar]
  • 34.Andersen, M. B., Stirling, C. H. & Weyer, S. Uranium isotope fractionation. Rev. Mineral. Geochem.82, 799–850 (2017). [Google Scholar]
  • 35.Brüske, A. et al. The onset of oxidative weathering traced by uranium isotopes. Precambrian Res.338, 1–16 (2020). [Google Scholar]
  • 36.Wang, X. et al. A Mesoarchean shift in uranium isotope systematics. Geochim. Cosmochim. Acta238, 438–452 (2018). [Google Scholar]
  • 37.Brennecka, G. A., Herrmann, A. D., Algeo, T. J. & Anbar, A. D. Rapid expansion of oceanic anoxia immediately before the end-Permian mass extinction. Proc. Natl. Acad. Sci. USA108, 17631–17634 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Noordmann, J. et al. Uranium and molybdenum isotope systematics in modern euxinic basins: case studies from the central Baltic Sea and the Kyllaren fjord (Norway). Chem. Geol.396, 182–195 (2015). [Google Scholar]
  • 39.Herrmann, A. J., Sorwat, J., Byrne, J. M., Frankenberg-Dinkel, N. & Gehringer, M. M. Diurnal Fe(II)/Fe(III) cycling and enhanced O2 production in a simulated Archean marine oxygen oasis. Nat. Commun.12, 2069 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Pourbaix, M. Atlas of Electrochemical Equilibria in Aqueous Solutions. (National Association of Corrosion Engineers, Houston, Texas, U.S., 1974).
  • 41.Langmuir, D. Uranium solution-mineral equilibria at low temperatures with applications to sedimentary ore deposits. Geochim. Cosmochim. Acta42, 547–569 (1978). [Google Scholar]
  • 42.Langmuir, D. Aqueous Environmental Geochemistry. (Prentice Hall, Upper Saddle River, New Jersey, U.S., 1997).
  • 43.Olson, S. L., Kump, L. R. & Kasting, J. F. Quantifying the areal extent and dissolved oxygen concentrations of Archean oxygen oases. Chem. Geol.362, 35–43 (2013). [Google Scholar]
  • 44.Liao, T., Wang, S., Hao, Z., Stüeken, E. E. & Luo, H. Dating ammonia-oxidizing bacteria with abundant eukaryotic fossils. Mol. Biol. Evol. 41, msae096 (2024). [DOI] [PMC free article] [PubMed]
  • 45.Thomazo, C., Nisbet, E. G., Grassineau, N. V., Peters, M. & Strauss, H. Multiple sulfur and carbon isotope composition of sediments from the Belingwe Greenstone Belt (Zimbabwe): a biogenic methane regulation on mass independent fractionation of sulfur during the Neoarchean? Geochim. Cosmochim. Acta121, 120–138 (2013). [Google Scholar]
  • 46.Mandernack, K. W., Mills, C. T., Johnson, C. A., Rahn, T. & Kinney, C. The δ15N and δ18O values of N2O produced during the co-oxidation of ammonia by methanotrophic bacteria. Chem. Geol.267, 96–107 (2009). [Google Scholar]
  • 47.Von Damm, K. L., Edmond, J. M., Measures, C. I. & Grant, B. Chemistry of submarine hydrothermal solutions at Guaymas Basin, Gulf of California. Geochim. Cosmochim. Acta49, 2221–2237 (1985). [Google Scholar]
  • 48.Lilley, M. D. et al. Anomalous CH4 and NH4+ concentrations at an unsedimented mid-ocean-ridge hydrothermal system. Nature364, 45–47 (1993). [Google Scholar]
  • 49.Bourbonnais, A., Lehmann, M. F., Butterfield, D. A. & Juniper, S. K. Subseafloor nitrogen transformations in diffuse hydrothermal vent fluids of the Juan de Fuca Ridge evidenced by the isotopic composition of nitrate and ammonium. Geochem. Geophys. Geosystems13, Q02T01 (2012).
  • 50.Grégoire, M. & Lacroix, G. Exchange processes and nitrogen cycling on the shelf and continental slope of the Black Sea basin. Glob. Biogeochem. Cycles17, 1073 (2003).
  • 51.Jenkins, W. J. Using excess 3He to estimate Southern Ocean upwelling time scales. Geophys. Res. Lett.47, e2020GL087266 (2020). [Google Scholar]
  • 52.Ardyna, M. et al. Hydrothermal vents trigger massive phytoplankton blooms in the Southern Ocean. Nat. Commun.10, 2451 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Rasmussen, B., Muhling, J. R., Suvorova, A. & Fischer, W. W. Apatite nanoparticles in 3.46–2.46 Ga iron formations: evidence for phosphorus-rich hydrothermal plumes on early Earth. Geology49, 647–651 (2021). [Google Scholar]
  • 54.Stüeken, E. E. Hydrothermal vents and organic ligands sustained the Precambrian copper budget. Geochem. Perspect. Lett. 10.7185/geochemlet.2037 (2020).
  • 55.Stüeken, E. E. et al. Optimized switch-over between CHNS abundance and CNS isotope ratio analyses by elemental analyzer-isotope ratio mass spectrometry: application to six geological reference materials. Rapid Commun. Mass Spectrom.34, 1–11 (2020). [DOI] [PubMed] [Google Scholar]
  • 56.Pallacks, S. et al. Anthropogenic acidification of surface waters drives decreased biogenic calcification in the Mediterranean Sea. Commun. Earth Environ.4, 1–10 (2023).37325084 [Google Scholar]
  • 57.Richter, S. et al. The isotopic composition of natural uranium samples—measurements using the new n(233U)/n(236U) double spike IRMM-3636. Int. J. Mass Spectrom.269, 145–148 (2008). [Google Scholar]
  • 58.Weyer, S. et al. Natural fractionation of 238U/235U. Geochim. Cosmochim. Acta72, 345–359 (2008). [Google Scholar]
  • 59.Cheng, H. et al. The half-lives of uranium-234 and thorium-230. Chem. Geol.169, 17–33 (2000). [Google Scholar]
  • 60.Richter, S. et al. New average values for the n(238U)/n(235U) isotope ratios of natural uranium standards. Int. J. Mass Spectrom.295, 94–97 (2010). [Google Scholar]
  • 61.Hofmann, A. & Kusky, T. The Belingwe greenstone belt: ensialic or oceanic? in Developments in Precambrian Geology vol. 13 487–538 (Elsevier, 2004).
  • 62.Planavsky, N., Partin, C. & Bekker, A. Carbon Isotopes as a Geochemical Tracer. in Encyclopedia of Astrobiology 366–371 (Springer, Berlin Heidelberg, Germany, 2015). 10.1007/978-3-662-44185-5_228.
  • 63.Stüeken, E. E., Kipp, M. A., Koehler, M. C. & Buick, R. The evolution of Earth’s biogeochemical nitrogen cycle. Earth-Sci. Rev.160, 220–239 (2016). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Peer Review File (600.5KB, pdf)
Source Data (24.9KB, xlsx)

Data Availability Statement

The stable isotope data generated in this study for nitrogen, organic carbon and uranium are provided in the Source Data file and uploaded in a Figshare repository (10.6084/m9.figshare.27632010). Source data are provided with this paper.


Articles from Nature Communications are provided here courtesy of Nature Publishing Group

RESOURCES