Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2021 May 31;118(23):e2023617118. doi: 10.1073/pnas.2023617118

A CO2 greenhouse efficiently warmed the early Earth and decreased seawater 18O/16O before the onset of plate tectonics

Daniel Herwartz a,1, Andreas Pack b, Thorsten J Nagel c
PMCID: PMC8201855  PMID: 34074764

Significance

Due to the lower luminosity of the young Sun, climate modelers struggle to explain why the climate on early Earth was not freezing cold. This “faint young Sun paradox” is in conflict with apparently hot Archean ocean temperatures (∼70 °C) that can be estimated from the 18O/16O stable isotope ratio of chemical sediments. We show that the later temperatures had been overestimated because the 18O/16O of seawater also changed over time due to intense carbonatization and silicification of the oceanic crust, which consumes heavy 18O. Because these processes require high fluxes of CO2, greenhouse warming by a CO2-rich atmosphere appears most feasible to explain all observations.

Keywords: triple oxygen isotopes, Archean, climate, faint young Sun, plate tectonics

Abstract

The low 18O/16O stable isotope ratios (δ18O) of ancient chemical sediments imply ∼70 °C Archean oceans if the oxygen isotopic composition of seawater (sw) was similar to modern values. Models suggesting lower δ18Osw of Archean seawater due to intense continental weathering and/or low degrees of hydrothermal alteration are inconsistent with the triple oxygen isotope composition (Δ’17O) of Precambrian cherts. We show that high CO2 sequestration fluxes into the oceanic crust, associated with extensive silicification, lowered the δ18Osw of seawater on the early Earth without affecting the Δ’17O. Hence, the controversial long-term trend of increasing δ18O in chemical sediments over Earth’s history partly reflects increasing δ18Osw due to decreasing atmospheric pCO2. We suggest that δ18Osw increased from about −5‰ at 3.2 Ga to a new steady-state value close to −2‰ at 2.6 Ga, coinciding with a profound drop in pCO2 that has been suggested for this time interval. Using the moderately low δ18Osw values, a warm but not hot climate can be inferred from the δ18O of the most pristine chemical sediments. Our results are most consistent with a model in which the “faint young Sun” was efficiently counterbalanced by a high-pCO2 greenhouse atmosphere before 3 Ga.


The amount of carbon that degassed from a solidifying magma ocean on the infant Earth 4.5 Ga ago was probably similar to the amount of CO2 that is now present in the atmosphere of Venus (pCO2 ∼90 bar) (1). Subsequently, dynamic carbon cycling between the early Earth’s atmosphere (atmospheric reservoir [RA]), the ocean (RO), and the oceanic crust (ROC) reservoirs stabilized a primordial Earth–atmospheric pCO2 to about 1.5 bar (2). Decreasing pCO2 over the Earth’s history reflects the net transfer of large masses of carbon out of the atmosphere-ocean-oceanic crust (RA+O+OC) system into the mantle and the emerging continental crust reservoir (RCC), the latter providing a long-term sink for carbon in the form of inorganic carbon (carbonate rocks) and organic material (organic matter–rich shales, coal, oil, and gas) (1, 2).

Early carbon-cycle models predicted high Archean pCO2 to account for the faint young Sun paradox (2, 3), whereas direct pCO2 estimates from the end of the Archean now imply much lower atmospheric pCO2 [∼0.01 to 0.1 bar (4)]. However, a compilation of evidence from the sedimentary record implies much higher pCO2 between 3.2 to 3.0 Ga compared to the 2.9 to 2.7 Ga interval (5). These authors state that even several bars pCO2 are feasible between 3.2 to 3.0 Ga, which is not in conflict with much lower Neoarchean pCO2 estimates (Fig. 1) or paleo-atmospheric pressure estimates at 2.7 Ga [<2 bar (6) and <0.5 bar (7)]. A fundamental drop in atmospheric CO2 mixing ratio is also reflected in the observation that >3 Ga, Archean mafic crust (greenstones) is commonly characterized by very intense carbonatization and silicification that is unparalleled in their modern analogs (812). Such observations provide evidence for deep-time paleo-pCO2 fluctuations with a drastic pCO2 drop starting around 3 Ga ago (5, 13).

Fig. 1.

Fig. 1.

Constraints on the RA+O+OC size over time with implications for pCO2. A illustrates the size of the RA+O+OC carbon reservoir and its distribution between the oceanic crust (ROC in gray), ocean water (RO in dark blue), and the atmosphere (RA in light blue). Transition of this carbon to the long-term RCC and the mantle (green) decreases the size of the RA+O+OC reservoir, which was still large at 3.2 Ga (see SI Appendix) and minimal at the onset of the global glaciations at 2.4 Ga. B shows two recent pH curves over Earth’s history to illustrate how the pH-dependent distribution of carbon between RA and RO may translate into pCO2 at a given time. (C) The panel summarizes pCO2 estimates [adopted from Catling and Zahnle (4)] and proposed pCO2 evolution curves illustrated as dashed lines: a (2), b (18), and c (proposed here). Qualitative evidence to construct curve c comes from rare evidence for glaciers (44) and lower degrees of carbonatization of oceanic crust at 3.5 Ga compared to 3.2 Ga (13), suggesting a transient interval of somewhat low pCO2 in the Paleoarchean. Low pCO2 is indicated prior to the onset of cold climates later in the rock record. The black bar at 4.5 Ga is derived from carbon-flux arguments and the primordial carbon reservoir (1, 2). The black bar at 3.2 Ga applies the same carbon-flux arguments to translate the high carbonate content observed in the oceanic crust in Pilbara into tentative pCO2 estimates (SI Appendix), which are most consistent with curve a (2).

Today, carbonatization, which is the formation of carbonates during alteration of the oceanic crust, mainly occurs in relatively cool, off-axis hydrothermal systems over the first 20 Mya after crust formation at midocean ridges (1416). Elevated degrees of carbonatization in oceanic crust from the Cretaceous and Jurassic are assigned to higher dissolved inorganic carbon at the time (14). Hence, higher pCO2 (i.e., a larger RA+O+OC) directly translates into higher degrees of carbonatization. While carbonate is mainly observed as vein fillings in the upper 300 m of oceanic crust today (14, 15), it extensively replaces glass and igneous minerals in Archean greenstones down to depths of 2 km below the ancient seafloor (810, 17). The amount of CO2 fixed in 3.2-Ga-old oceanic crust from Pilbara, Australia is estimated at 1.2 × 107 mol ⋅ m−2 (±10%) (17)—a remarkable figure that is about two orders of magnitude more compared to today (SI Appendix). Much lower degrees of carbonatization in oceanic crust are already observed at 2.6 Ga, suggesting a drastic Mesoarchean drop in pCO2 (13) (Fig. 1A).

The qualitative evidence from the sediment record (5) and from the degree of carbonatization and silicification of oceanic crust (13) has not been included in proposed pCO2 curves (4, 18) because the quantitative conversions into absolute pCO2 require some assumptions (SI Appendix). Nevertheless, pCO2 during the early Archean may have been as high as initially predicted by Kasting (2), followed by a pronounced Mesoarchean drop (13) to levels consistent with available paleo-pCO2 estimates toward the Neoarchean (4) (stippled line “c” in Fig. 1C). Further decreasing pCO2 toward modern levels partly reflects increasing ocean pH (Fig. 1) rather than a shrinking RA+O+OC (SI Appendix). Hence, only small effects of carbonatization on the δ18Osw value are expected for post-Archean seawater (16). Here, we focus on the very high carbonatization (810, 13) and silicification (11, 12) fluxes before the Mesoarchean pCO2 drop, and we model the respective effects on ancient δ18Osw.

The Effect of an Increased CO2 Flux on the δ18Osw of Seawater

The δ18Osw of seawater is controlled by the fluxes Fi of oxygen between rocks and the hydrosphere (10). Low-temperature (low-T) weathering products such as clays have ∼20 to 25‰ higher δ18O values compared to pristine silicate crust (∼5.5 to 12‰). Hence, low-T weathering on the continents (expressed as flux Fcw) and on the seafloor (Fsfw) lowers the δ18Osw value of seawater by extracting heavy 18O from the hydrosphere. Alteration of oceanic crust at hydrothermal temperatures (∼250 to 350 °C) produces rocks with ∼1 to 2‰ lower δ18O (∼4.5‰) than the pristine oceanic crust basalt (∼5.8‰), thereby adding 18O to the oceans (Fsp). Classic mass-balance models also account for minor effects from water recycling through the mantle (Fr) and the influence of continental growth (Fcg) (19, 20).

In order to explain the enigmatic shift in δ18O of ∼15‰ of chemical sediments through time (2128), extreme and difficult-to-reconcile changes of oxygen fluxes Fi are required (20, 29). As shown in Fig. 2, even a moderate shift of ocean water δ18Osw by −5‰ would necessitate very high continental weathering rates (Fcw >10 times present) or very low hydrothermal alteration fluxes (Fsp <0.15 times present). Besides the substantial implications for Precambrian paleoenvironments (21, 30), such flux manipulations would induce much higher ∆’17O (SI Appendix, Definitions) in Precambrian cherts than documented (Fig. 2). This is presently acknowledged as a strong argument against low δ18Osw as an explanation for the low δ18O of Precambrian chemical sediments (2427, 31). Most of the ∆’17Ochert data do not fall on the black equilibrium curve in Fig. 2B (2427), suggesting that these samples are affected by postdepositional alteration processes. Alteration lowers both δ18O and ∆’17O, hence most triple oxygen isotope data seem to suggest postdepositional alteration as the most viable hypothesis for the long-term shift in δ18O (2427). Due to sample-size requirements, the least altered cherts measured for ∆’17Ochert comprise somewhat lower δ18O compared to the most pristine in situ δ18O analyses (27, 31). Nevertheless, triple oxygen isotope compositions of well-preserved cherts from the Barberton Greenstone Belt, South Africa fall close to the equilibrium line consistent with hot Archean seawater (31). By combining data from the two techniques, Lowe et al. (31) conclude that “…Archean surface temperatures were well above those of the present day, perhaps as high as 66 to 76 °C.”

Fig. 2.

Fig. 2.

Plots of ∆’17Osw versus δ18O for the various models and published chert compositions. A shows how seawater composition changes when individual fluxes are manipulated. The δ18Osw of seawater decreases for substantially lower Fsp (green) or higher Fcw (gray) but comes along with a ∆’17Osw increase. Tick marks refer to multiples of the fluxes used in the base scenario (SI Appendix). Accounting for Archean carbonatization and silicification (FCO2+SiO2) also decreases the δ18Osw of seawater (brown) but leaves ∆’17Osw largely unchanged (a). Temperatures in A indicate average precipitation temperatures (Tprecip) for carbonates and silicates. B shows how measured triple oxygen isotope compositions of cherts compare to the model. The data are derived from refs. 2527 and 31. Silica precipitated in equilibrium with modern ocean water must fall on the black equilibrium curve at the respective precipitation temperature. The greenish field illustrates ∆’17O compositions expected for cherts that precipitated from a low δ18Osw due to increasing Fcw or decreasing Fsp as modeled by Sengupta and Pack (24). When accounting for FCO2+SiO2, however, the δ18Osw of seawater decreases, and ∆’17Osw remains broadly similar, consistent with Precambrian chert ∆’17O (brown field). The solid and stippled gray curves indicate precipitation from scenarios a and b, respectively. The light brown field encompasses all possible solutions. The stippled brown line at δ18O = 23‰ indicates the highest Archean δ18O analyzed in situ by microanalytical techniques (27, 31). The red arrow depicts an expected slope for alteration.

This interpretation hinges on the assumption that δ18Osw was always close to its present-day value. The classic and apparently strong argument for a constant δ18Osw of the oceans is derived from the roughly constant δ18O composition of oceanic crust profiles through time (29). It has been proposed that low δ18Osw ocean water would drive altered oceanic crust to low values (29) as observed for rocks altered by isotopically light meteoric water (3234). However, while local crust alteration with meteoric water constitutes the case of an open system, alteration of the oceanic crust by the ocean-water reservoir as a whole should be regarded as a closed system. In order to extract 18O from seawater, a complementary reservoir must incorporate 18O. Interestingly, Archean greenstones tend to be slightly enriched in δ18O (29) and thus provided a sink for 18O. A typical feature of Archean terranes >3 Ga is their high degree of silicification, which may well explain the overall elevated δ18O observed in Archean greenstones (11, 12, 29), which act as the complementary reservoir for depleted ocean water δ18O in the scenario suggested herein.

We propose that the significance of CO2 sequestration through carbonatization along with intense silicification, both associated with high Archean pCO2, have been overlooked and that they offer an explanation for the long-standing oxygen isotope conundrum of chemical sediments deep in time.

Methods and Results

To explore how variable CO2 sequestration through carbonatization and silicification affects the oxygen isotope composition of the oceans at variable temperatures, we use Muehlenbachs’ model (20), which was recently extended for δ17Osw (24) (Fig. 2A). After slight modifications (SI Appendix), the base scenario for modern-day fluxes gives a seawater steady state δ18Osw-ss = −0.53‰, intermediate between the modern ocean and the presumed value for an ice-free world (i.e., δ18Osw ∼−1‰). The respective ∆’17Osw-ss = −9 ppm (i.e., −0.009‰) matches the measured value of −5 ppm for modern seawater within uncertainty (35) (Fig. 2A).

Volcanic CO2 equilibrates with silicates in the magma chamber before it is outgassed with near-mantle–like δ18O ∼5.5‰ (SI Appendix). It is subject to various fractionation processes but is eventually immobilized as carbonate with higher δ18O. Such carbonates, if they become subducted, are mostly volatilized, and the respective CO2 is reintroduced into the atmosphere with mantle-like δ18O. Because of the practically infinite mantle oxygen reservoir, CO2 sequestration and recycling (defined here as FCO2) extracts 18O from the hydrosphere.

Heavily silicified greenstones are common in the Archean (11, 12) simply because CO2 weathering of silicates produces both carbonate (CaCO3) and silica (SiO2). For each mole of consumed CO2 and precipitated CaCO3, 1 mol SiO2 forms (CaSiO3 + CO2 → CaCO3 + SiO2, Urey reaction) so that FSiO2 equals FCO2. As for carbonates, quartz comprises high δ18O even at elevated precipitation temperatures and removes heavy 18O from the oceans (SI Appendix). Therefore, enhanced CO2-driven carbonatization and silicification of oceanic crust provides a means of lowering the δ18O value of Archean seawater (Fig. 2A). These CO2-related fluxes, FSiO2 and FCO2, are not considered in existing models (1921, 24, 29) and could dramatically differ from the present day owing to the potentially very different atmospheric composition.

The CO2-sequestration flux FCO2 by carbonatization of oceanic crust at 3.2 Ga is estimated as 1.5 × 1014 mol ⋅ yr−1 (17), which is two orders of magnitude larger than the modern carbonatization rate of only 0.45 to 2.4 × 1012 mol ⋅ yr−1 (14, 15). This high estimate is based on the assumption that spreading rates at the time were three times as high as those at present (36, 37), which is inconsistent with suggestions that plate tectonics only started around 3 Ga (38, 39) but is consistent with proposals that the mode of plate tectonics simply evolved over time (40). The various competing tectonic models proposed for the Archean impose large uncertainties not only on newly introduced FCO2+SiO2 but for all fluxes. Respective modifications, especially for the classic Fcw and Fsp fluxes, have previously been invoked for the Archean (21, 30) to argue for very low δ18Osw, down to −13.3‰. These particular flux modifications can be ruled out, as they induce high ∆’17Osw, which is inconsistent with the observed low ∆’17Ochert (2427). By adding the CO2-sequestration flux of Shibuya et al. (17) to our base model, we can explore how it affects the triple oxygen isotopic composition of seawater. Moderate modification of all fluxes then allows identification of feasible flux combinations that most closely resemble the measured ∆’17Ochert data.

Accounting for the respective FCO2+SiO2 yields δ18Osw-ss of seawater between −7.5 and −2‰ (Fig. 2) and mainly depends on the average temperature for carbonation and silicification within oceanic crust (Tprecip). Heavy 18O is extracted more efficiently from the hydrosphere at low Tprecip. Most notably, consideration of carbonatization and silicification fluxes FCO2+SiO2 and precipitation of the silica and carbonate at 100 to 200 °C lowers the ocean’s δ18Osw-ss but does not lead to the increase in ∆’17Osw-ss as seen in (24) (Model S1 and Fig. 2A).

Accounting for CO2 sequestration and silicification provides a much better fit to observed ∆’17O of Precambrian cherts compared to previous suggestions to lower δ18Osw without FCO2+SiO2 (Fig. 2B). Simply using the modern-day fluxes of Muehlenbachs (20) in combination with the estimated CO2 flux at 3.2 Ga (17) and Tprecip of 150 °C yields δ18Osw-ss of −3‰ (scenario “a” in Fig. 2A). The δ18Osw-ss can be forced to very low values, but the ∆’17Osw increases for such end-members (Fig. 2), making respective scenarios incompatible with chert data. A range of more realistic flux combinations cluster around a resultant δ18Osw-ss of seawater of −5‰ with near zero ∆’17O (e.g., scenario “b” in Fig. 2A; see Model S1 and SI Appendix for a sensitivity analysis). Such a scenario is best compatible with triple oxygen isotope ratios of Archean cherts (2527, 41) and shales (42).

A direct seawater estimate [δ18Osw = −1.7 ± 1.1‰ (43)] at 2.4 Ga corresponds to a time when the degree of carbonatization of oceanic crust was even lower than today (13), suggesting that slightly lower δ18Osw persisted in the Paleoproterozoic even without significant CO2 sequestration and silicification. We suggest that δ18Osw further decreased to −5 ± 2‰ when pCO2 was high at 3.2 Ga. We propose a slightly higher δ18Osw of ∼−3 ± 2‰ for the Paleoarchean (at 3.5 Ga) to account for apparently lower degrees of carbonatization (13), observations of cold climate (44), and high δ18Ocarbonate and δ18Ophosphate (45) at the time (Fig. 3). Both approximations are used in combination with direct δ18Osw estimates (43, 46, 47) to construct a seawater curve over time (Fig. 3). Shifts in δ18Osw on these timescales are viable because δ18Osw can adapt to a new steady state within a few tens of millions of years (SI Appendix).

Fig. 3.

Fig. 3.

Compiled oxygen-isotope record over Earth’s history with suggested δ18Osw evolution. The black curve from Jaffres et al. (21) is constructed through the database for δ18Ocarbonate (calcite = gray points; dolomite and other = gray diamonds) and slightly extended (stippled). Chert data compiled by Knauth (28) (green circles) is extended by data from refs. 22, 2527, and 55 (green dots). Ranges for modern and ancient δ18Ophosphate (45) are indicated by brownish boxes. The proposed seawater curve is constructed from published δ18Osw (43, 46, 47) and Archean δ18Osw proposed herein (see Methods and Results). High δ18Osw estimates (56, 57) are not included (SI Appendix). The curve and its blue margin of 2‰ are not based on statistics. Arrows indicate warming (red) and cooling (blue) trends expected from the changing size of the RA+O+OC reservoir. Temperature calibrations are anchored at δ18Osw = −5‰ for cherts and calcite and at δ18Osw = −3‰ for phosphates. Apparently, too-high δ18Ocarbonate are dolomites that generally comprise ∼2‰ higher δ18O compared to calcite.

Discussion

Even with the moderately low δ18Osw of seawater and warm precipitation temperatures proposed here, alteration is still required to explain the triple oxygen isotopic compositions of most Precambrian cherts (Fig. 2). Bedded cherts are frequently interpreted to reflect precipitation from warm/hot bottom water brines or hydrothermal plumes (2527). Cherts of superior-type banded iron formations, however, certainly originate from primary precipitates from ambient seawater temperatures on continental shelves. Surface waters within epicontinental seaways may have been low in δ18O due to a partly meteoric origin (29), and respective ∆’17O would have been only slightly elevated. Alternatively, isotopic exchange between iron phases (with low δ18OFeOx) and silica bands (with high δ18OopalA) decreased δ18Ochert and generated ∆’17Ochert below the equilibrium curve (41). Alteration with (meteoric) water resembles a similar mixing mechanism that has been suggested to lower ∆’17Ochert (2527). The variable combinations of these processes induce the considerable scatter in the oxygen-isotopic composition of chemical sediments through time (Fig. 3). Collectively, the Archean δ18O record can be rationalized by a combination of the classic explanations: 1) a moderately low δ18Osw, 2) warm ocean temperatures, and 3) alteration.

Although even the most pristine Archean cherts are altered to some extent, their oxygen isotope composition reflects lower δ18O of Archean ocean water as a result of high degrees of carbonatization (810, 13) and silicification at >3 Ga ago (11, 12). This hypothesis needs further verification from high-precision triple oxygen isotope analyses of other types of chemical sediments. Eventually, concepts to “see through” diagenesis (48) can reveal the most viable triple oxygen isotope composition of Archean seawater. Analyzing the triple oxygen isotopic composition of apparently unaltered phosphates (45) (but see ref. 31) may even put direct constraints both on Archean seawater ∆’17Osw and precipitation temperatures. Knowledge of the exact ∆’17Osw would further restrict feasible Archean oxygen fluxes including the relative proportion of the CO2-sequestration flux. Respective constraints may also provide information on the tectonic regime (e.g., sagduction versus subduction, high versus low spreading rates, etc.) that is active at a given time (3640).

From the model presented here, lowering the seawater δ18O requires high CO2 fluxes, which are most viable for a high-pCO2 atmosphere. The exact level of pCO2 in the Archean most probably fluctuated quite a bit as it has done for the past billion years. Large fluctuations, such as the proposed Mesoarchean drop in pCO2 (5, 13), should induce a significant shift in δ18Osw. Indeed, a respective shift is broadly recorded in the oxygen isotope record (Fig. 3), implying that carbonatization and silicification processes had indeed been large enough to significantly decrease the δ18Osw.

Even with the moderately low δ18Osw proposed herein, the temperatures estimated from chemical sediments imply a very warm climate on the early Earth, which is consistent with the rare evidence for ice before 3 Ga ago (44). With respect to the vigorously debated faint young Sun paradox (3, 49, 50), this implies generally high concentrations of greenhouse gases for most of the Archean. Our study supports the original idea that CO2 and not CH4 counterbalanced the lower luminosity of the faint young Sun (2, 3), at least prior to the stark Mesoarchean drop in pCO2 (5, 13). The apparently warmer temperatures compared to the Phanerozoic leads to the question of how pCO2 can be maintained at high levels in a constantly warm climate, where it accelerates CO2 weathering, which leads to a decrease in pCO2 and keeps Earth’s surface in a temperature window that is “just right” (51).

Carbon transfer from the RA+O+OC reservoirs to the RCC and the mantle (i.e., to the long-term storage reservoirs) was probably inefficient on the early Earth. Continental crust was less abundant, and thus the continental reservoir was small. The residence times for carbon on early continents was probably shorter than today due to efficient recycling of thin continental-crust margins and acidic rain (due to high pCO2). Sagduction tectonics, proposed for the early Earth (39), is associated with high thermal gradients leading to efficient CO2 recycling and thus inefficient carbon storage in the mantle.

Modern-style plate tectonics provides a mechanism to accrete land masses and to thicken continental crust (52), which generates topography and subaerial land surfaces. The two negative feedbacks for pCO2 are the following: 1) the establishment of stable, long-term storage reservoirs for carbon and 2) an increased supply of weatherable rocks, enhancing continental CO2-weathering fluxes. Subaerial land mass is assumed to be small before the Neoarchean, with the large-scale emergence of continents above sea level proposed to occur at 2.5 Ga (53). The RA+O+OC carbon reservoir drastically decreased in size in the preceding several hundred million years, suggesting that carbon was already relocated to the long-term storage reservoir on evolving continental platforms (38, 40, 52) and that carbon transfer to the mantle became more efficient within proto-Phanerozoic–type subduction zones. Hence, sharply decreasing pCO2 could be a consequence of a rapid or gradual switch from vertical sagduction–type tectonics to a horizontal subduction regime (39), generating thick, silica-rich continental crust from about 3 Ga onwards (52).

Conclusions

In this contribution, we propose that the intense carbonatization and silica precipitation due to high CO2 fluxes in the Archean led to lower δ18Osw values of seawater without affecting the Δ’17O. The shift in δ18Osw is mirrored in the lower δ18O of early Archean chemical sediments. The decreasing intensity of carbonatization and silica precipitation around ∼3 Ga shifted the isotope composition of the ocean to similar values as seen in the oceans today. On a smaller scale, however, variable CO2 fluxes in the Proterozoic and Phanerozoic also induce variations in FCO2+SiO2 and thus in seawater δ18Osw. Therefore, long-term fluctuations in the δ18O of chemical sediments do not only reflect temperature but also changing δ18Osw. The apparent temperature change estimated for long-term climate fluctuations (e.g., ref. 54) are thus systematically overestimated.

Supplementary Material

Supplementary File
pnas.2023617118.sapp.pdf (438.5KB, pdf)
Supplementary File
pnas.2023617118.sd01.xlsx (21.5KB, xlsx)

Footnotes

The authors declare no competing interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.2023617118/-/DCSupplemental.

Data Availability

All data are included in the article and/or supporting information.

Change History

June 3, 2021: The supporting information has been updated.

References

  • 1.Walker J. C. G., Carbon dioxide on the early earth. Orig. Life Evol. Biosph. 16, 117–127 (1985). [DOI] [PubMed] [Google Scholar]
  • 2.Kasting J. F., Theoretical constraints on oxygen and carbon dioxide concentrations in the Precambrian atmosphere. Precambrian Res. 34, 205–229 (1987). [DOI] [PubMed] [Google Scholar]
  • 3.Feulner G., The faint young Sun problem. Rev. Geophys. 50, RG2006 (2012). [Google Scholar]
  • 4.Catling D. C., Zahnle K. J., The Archean atmosphere. Sci. Adv. 6, eaax1420 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Lowe D. R., Tice M. M., Geologic evidence for Archean atmospheric and climatic evolution: Fluctuating levels of CO2, CH4, and O2 with an overriding tectonic control. Geology 32, 493 (2004). [Google Scholar]
  • 6.Som S. M., Catling D. C., Harnmeijer J. P., Polivka P. M., Buick R., Air density 2.7 billion years ago limited to less than twice modern levels by fossil raindrop imprints. Nature 484, 359–362 (2012). [DOI] [PubMed] [Google Scholar]
  • 7.Som S. M., et al., Earth’s air pressure 2.7 billion years ago constrained to less than half of modern levels. Nat. Geosci. 9, 448–451 (2016). [Google Scholar]
  • 8.Kitajima K., Maruyama S., Utsunomiya S., Liou J. G., Seafloor hydrothermal alteration at an Archaean mid-ocean ridge. J. Metamorph. Geol. 19, 583–599 (2001). [Google Scholar]
  • 9.Nakamura K., Kato Y., Carbonatization of oceanic crust by the seafloor hydrothermal activity and its significance as a CO2 sink in the Early Archean. Geochim. Cosmochim. Acta 68, 4595–4618 (2004). [Google Scholar]
  • 10.Shibuya T., et al., Middle Archean ocean ridge hydrothermal metamorphism and alteration recorded in the Cleaverville area, Pilbara Craton, Western Australia. J. Metamorph. Geol. 25, 751–767 (2007). [Google Scholar]
  • 11.Abraham K., et al., Coupled silicon–oxygen isotope fractionation traces Archaean silicification. Earth Planet. Sci. Lett. 301, 222–230 (2011). [Google Scholar]
  • 12.Hofmann A., Harris C., Silica alteration zones in the Barberton greenstone belt: A window into subseafloor processes 3.5–3.3 Ga ago. Chem. Geol. 257, 221–239 (2008). [Google Scholar]
  • 13.Shibuya T., et al., Weak hydrothermal carbonation of the Ongeluk volcanics: Evidence for low CO2 concentrations in seawater and atmosphere during the Paleoproterozoic global glaciation. Prog. Earth Planet. Sci. 4, 31 (2017). [Google Scholar]
  • 14.Gillis K. M., Coogan L. A., Secular variation in carbon uptake into the ocean crust. Earth Planet. Sci. Lett. 302, 385–392 (2011). [Google Scholar]
  • 15.Alt J. C., Teagle D. A. H., The uptake of carbon during alteration of ocean crust. Geochim. Cosmochim. Acta 63, 1527–1535 (1999). [Google Scholar]
  • 16.Coogan L. A., Daëron M., Gillis K. M., Seafloor weathering and the oxygen isotope ratio in seawater: Insight from whole-rock δ18O and carbonate δ18O and Δ47 from the Troodos ophiolite. Earth Planet. Sci. Lett. 508, 41–50 (2019). [Google Scholar]
  • 17.Shibuya T., et al., Depth variation of carbon and oxygen isotopes of calcites in Archean altered upperoceanic crust: Implications for the CO2 flux from ocean to oceanic crust in the Archean. Earth Planet. Sci. Lett. 321–322, 64–73 (2012). [Google Scholar]
  • 18.Krissansen-Totton J., Arney G. N., Catling D. C., Constraining the climate and ocean pH of the early Earth with a geological carbon cycle model. Proc. Natl. Acad. Sci. U.S.A. 115, 4105–4110 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Muehlenbachs K., Clayton R. N., Oxygen isotope composition of the oceanic crust and its bearing on seawater. J. Geophys. Res. 81, 4365–4369 (1976). [Google Scholar]
  • 20.Muehlenbachs K., The oxygen isotopic composition of the oceans, sediments and the seafloor. Chem. Geol. 145, 263–273 (1998). [Google Scholar]
  • 21.Jaffrés J. B. D., Shields G. A., Wallmann K., The oxygen isotope evolution of seawater: A critical review of a long-standing controversy and an improved geological water cycle model for the past 3.4 billion years. Earth Sci. Rev. 83, 83–122 (2007). [Google Scholar]
  • 22.Hren M. T., Tice M. M., Chamberlain C. P., Oxygen and hydrogen isotope evidence for a temperate climate 3.42 billion years ago. Nature 462, 205–208 (2009). [DOI] [PubMed] [Google Scholar]
  • 23.Galili N., et al., The geologic history of seawater oxygen isotopes from marine iron oxides. Science 365, 469–473 (2019). [DOI] [PubMed] [Google Scholar]
  • 24.Sengupta S., Pack A., Triple oxygen isotope mass balance for the Earth’s oceans with application to Archean cherts. Chem. Geol. 495, 18–26 (2018). [Google Scholar]
  • 25.Sengupta S., et al., Triple oxygen isotopes of cherts through time. Chem. Geol. 554, 119789 (2020). [Google Scholar]
  • 26.Liljestrand F. L., et al., The triple oxygen isotope composition of Precambrian chert. Earth Planet. Sci. Lett. 537, 116167 (2020). [Google Scholar]
  • 27.Zakharov D.O., Marin-Carbonne J., Alleon J., Bindeman I.N., Triple oxygen isotope trend recorded by Precambrian cherts: A perspective from combined bulk and in situ secondary ion probe measurements. Rev. Mineral. Geochem. 86, 323–365 (2021). [Google Scholar]
  • 28.Knauth L. P., “Temperature and salinity history of the Precambrian ocean: Implications for the course of microbial evolution” in Geobiology: Objectives, Concepts, Perspectives, Noffke N., Ed. (Elsevier, 2005), pp. 53–69. [Google Scholar]
  • 29.Gregory R. T., Ophiolites and global geochemical cycles: Implications for the isotopic evolution of seawater. Geol. Soc. Lond. Spec. Publ. 218, 353–368 (2003). [Google Scholar]
  • 30.Kasting J. F., et al., Paleoclimates, ocean depth, and the oxygen isotopic composition of seawater. Earth Planet. Sci. Lett. 252, 82–93 (2006). [Google Scholar]
  • 31.Lowe D. R., Ibarra D. E., Drabon N., Chamberlain C. P., Constraints on surface temperature 3.4 billion years ago based on triple oxygen isotopes of cherts from the Barberton Greenstone Belt, South Africa, and the problem of sample selection. Am. J. Sci. 320, 790–814 (2020). [Google Scholar]
  • 32.Wenner D. B., Taylor H. P., D/H and O18/O16 studies of serpentinization of ultramaflc rocks. Geochim. Cosmochim. Acta 38, 1255–1286 (1974). [Google Scholar]
  • 33.Criss R. E., Taylor H. P. Jr, An 18O/16O and D/H study of Tertiary hydrothermal systems in the southern half of the Idaho batholith. Bull. Geol. Soc. Am. 94, 640–663 (1983). [Google Scholar]
  • 34.Chamberlain C. P., et al., Triple oxygen isotopes of meteoric hydrothermal systems – implications for palaeoaltimetry. Geochem. Perspect. Lett. 15, 6–9 (2020). [Google Scholar]
  • 35.Luz B., Barkan E., Variations of 17O/16O and 18O/16O in meteoric waters. Geochim. Cosmochim. Acta 74, 6276–6286 (2010). [Google Scholar]
  • 36.Komiya T., Material circulation model including chemical differentiation within the mantle and secular variation of temperature and composition of the mantle. Phys. Earth Planet. Inter. 146, 333–367 (2004). [Google Scholar]
  • 37.Sleep N. H., Windley B. F., Archean plate tectonics: Constraints and inferences. J. Geol. 90, 363–379 (1982). [Google Scholar]
  • 38.Cawood P. A., et al., Geological archive of the onset of plate tectonics. Philos. Trans.- Royal Soc., Math. Phys. Eng. Sci. 376, 20170405 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Van Kranendonk M. J., Hugh Smithies R., Hickman A. H., Champion D. C., 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, 1–38 (2007). [Google Scholar]
  • 40.Windley B. F., Kusky T., Polat A., Onset of plate tectonics by the Eoarchean. Precambrian Res. 352, 105980 (2021). [Google Scholar]
  • 41.Levin N. E., Raub T. D., Dauphas N., Eiler J. M., Triple oxygen isotope variations in sedimentary rocks. Geochim. Cosmochim. Acta 139, 173–189 (2014). [Google Scholar]
  • 42.Bindeman I. N., Triple oxygen isotopes in evolving continental crust, granites, and clastic sediments. Rev. Mineral. Geochem. 86, 241–290 (2021). [Google Scholar]
  • 43.Zakharov D. O., Bindeman I. N., Triple oxygen and hydrogen isotopic study of hydrothermally altered rocks from the 2.43–2.41 Ga Vetreny belt, Russia: An insight into the early Paleoproterozoic seawater. Geochim. Cosmochim. Acta 248, 185–209 (2019). [Google Scholar]
  • 44.de Wit M. J., Furnes H., 3.5-Ga hydrothermal fields and diamictites in the Barberton Greenstone Belt-Paleoarchean crust in cold environments. Sci. Adv. 2, e1500368 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Blake R. E., Chang S. J., Lepland A., Phosphate oxygen isotopic evidence for a temperate and biologically active Archaean ocean. Nature 464, 1029–1032 (2010). [DOI] [PubMed] [Google Scholar]
  • 46.Hodel F., et al., Fossil black smoker yields oxygen isotopic composition of Neoproterozoic seawater. Nat. Commun. 9, 1453 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Peters S. T. M., et al., >2.7 Ga metamorphic peridotites from southeast Greenland record the oxygen isotope composition of Archean seawater. Earth Planet. Sci. Lett. 544, 116331 (2020). [Google Scholar]
  • 48.Wostbrock J. A. G., et al., Calibration of carbonate-water triple oxygen isotope fractionation: Seeing through diagenesis in ancient carbonates. Geochim. Cosmochim. Acta 288, 369–388 (2020). [Google Scholar]
  • 49.Rosing M. T., Bird D. K., Sleep N. H., Bjerrum C. J., No climate paradox under the faint early Sun. Nature 464, 744–747 (2010). [DOI] [PubMed] [Google Scholar]
  • 50.Goldblatt C., Zahnle K. J., Faint young Sun paradox remains. Nature 474, E1 (2011). [DOI] [PubMed] [Google Scholar]
  • 51.Kasting J. F., The Goldilocks planet? How silicate weathering maintains Earth “just right”. Elements 15, 235–240 (2019). [Google Scholar]
  • 52.Dhuime B., Wuestefeld A., Hawkesworth C. J., Emergence of modern continental crust about 3 billion years ago. Nat. Geosci. 8, 552–555 (2015). [Google Scholar]
  • 53.Bindeman I. N., et al., Rapid emergence of subaerial landmasses and onset of a modern hydrologic cycle 2.5 billion years ago. Nature 557, 545–548 (2018). [DOI] [PubMed] [Google Scholar]
  • 54.Zachos J., Pagani M., Sloan L., Thomas E., Billups K., Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292, 686–693 (2001). [DOI] [PubMed] [Google Scholar]
  • 55.Sharma S. D., Patil D. J., Srinivasan R., Gopalan K., Very high 18o enrichment in Archean cherts from South India: Implications for Archean ocean temperature. Terra Nova 6, 385–390 (1994). [Google Scholar]
  • 56.Pope E. C., Bird D. K., Rosing M. T., Isotope composition and volume of Earth’s early oceans. Proc. Natl. Acad. Sci. U.S.A. 109, 4371–4376 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Johnson B. W., Wing B. A., Limited Archaean continental emergence reflected in an early Archaean 18O-enriched ocean. Nat. Geosci. 13, 243–248 (2020). [Google Scholar]

Associated Data

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

Supplementary Materials

Supplementary File
pnas.2023617118.sapp.pdf (438.5KB, pdf)
Supplementary File
pnas.2023617118.sd01.xlsx (21.5KB, xlsx)

Data Availability Statement

All data are included in the article and/or supporting information.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES