Isotope Composition and Volume of Earth's Early Oceans: A, B, C, 1 A, C A, B, C

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Isotope composition and volume of Earths early oceans

Emily C. Popea,b,c,1, Dennis K. Birda,c, and Minik T. Rosinga,b,c


a Department of Geological and Environmental Sciences, Stanford University, Stanford, California 94305; bNatural History Museum of Denmark, University of Copenhagen, ster Voldgade 5-7, 1350 Kbenhavn K, Denmark; and cNordic Center for Earth Evolution, ster Voldgade 5-7, 1350 Kbenhavn K, Denmark

Edited by Robert N. Clayton, University of Chicago, Chicago, IL, and approved January 19, 2012 (received for review September 23, 2011)

Oxygen and hydrogen isotope compositions of Earths seawater are controlled by volatile fluxes among mantle, lithospheric (oceanic and continental crust), and atmospheric reservoirs. Throughout geologic time the oxygen mass budget was likely conserved within these Earth system reservoirs, but hydrogens was not, as it can escape to space. Isotopic properties of serpentine from the approximately 3.8 Ga Isua Supracrustal Belt in West Greenland are used to characterize hydrogen and oxygen isotope compositions of ancient seawater. Archaean oceans were depleted in deuterium [expressed as D relative to Vienna standard mean ocean water (VSMOW)] by at most 25 5, but oxygen isotope ratios were comparable to modern oceans. Mass balance of the global hydrogen budget constrains the contribution of continental growth and planetary hydrogen loss to the secular evolution of hydrogen isotope ratios in Earths oceans. Our calculations predict that the oceans of early Earth were up to 26% more voluminous, and atmospheric CH4 and CO2 concentrations determined from limits on hydrogen escape to space are consistent with clement conditions on Archaean Earth.
Archaean faint early sun hydrogen escape hydrosphere serpentine

he geologic record provides scant definitive evidence of early Archaean seawater chemistry. Oxygen and hydrogen isotope compositions of oceanic sedimentary rocks provide a commonly used proxy record, but interpretations remain inconclusive (14). Average 18 O values of Archaean marine sediments (cherts and carbonates) are tens of per mil lower than present-day values, and systematically increase over geologic time (2, 3). This trend has been attributed either to lower 18 OSEA WATER in the Archaean (1, 3), or to ocean temperatures up to approximately 70 C (58). Both interpretations have been questioned based on the susceptibility of chemical sediments to diagenetic alteration over geologic time (2, 4), and for the latter, a lack of geologic evidence for the extreme greenhouse gas concentrations required to sustain such high temperatures under a less luminous young Sun (9). Minerals formed by metasomatic reactions between seawater and oceanic lithosphere provide an alternative proxy for early ocean chemistry (1014). Here we report hydrogen and oxygen isotope compositions of 114 silicate mineral separates from the ca. 3.8 Ga Isua Supracrustal Belt (ISB) of West Greenland, including metamorphosed basalts, gabbros, and ultramafic rocks that represent fragments of Eoarchaean oceanic crust (see Fig. S1, Table S2). Minerals formed during heterogeneous late-stage CO2 - or meteoric-dominated metamorphic overprinting are distinguished by characteristic trends in their coupled D and 18 O values (Fig. S2). In contrast, serpentine minerals formed by reaction of seawater with ultramafic rocks (peridotites) preserved in a low-strain enclave of the ISB (15) define a different isotopic trend. Here we use these serpentines to constrain DH and 18 O16 O ratios of the Eoarchaean ocean. Serpentine Isotope Geochemistry Serpentine-group minerals are hydrous magnesian silicates commonly formed by hydrothermal infiltration of seawater into ultramafic oceanic crust, hydration of the mantle wedge by slabwww.pnas.org/cgi/doi/10.1073/pnas.1115705109

derived fluids, or fluid-rock interaction in continentally emplaced ultramafic lithologies (1618). Serpentinized oceanic peridotites are dominantly composed of low-temperature (<300 C) polymorphs lizardite and chrysotile, and to a lesser extent antigorite, which forms at temperatures >250 C (17). In modern oceanic serpentinites, antigorites have D between 46 and 30 (Figs. 1 and 2A, filled squares), values that indicate equilibrium with modern-day seawater at temperatures between 150 and 300 C (19). In contrast, chrysotiles and lizardites have D predominantly <50 (Figs. 1 and 2A, open squares). They likely formed by recrystallization in the presence of low-D fluids originating from dehydration reactions in hydrothermally altered oceanic crust (20, 21). Due to its more refractory nature (12, 22, 23), antigorite retains its primary D value, even when lizardite and chrysotile reequilibrate (21) (Fig. 1). Oxygen isotopes in oceanic serpentines range from 0.8 to 12.4, with the majority between 3 and 7 (Fig. 2B). Isotopic variation is due to the temperature dependence of serpentine-water 18 O fractionation (19) and the exchange of oxygen isotopes between the fluids and basaltic ocean crust (24). This can be seen in the 18 O of fluids venting at mid-ocean ridge hydrothermal systems, which are typically enriched in 18 O relative to seawater by 2 to 3 (2426). Serpentinites in ophiolites range in age from Tertiary to the Eoarchaean, and show a wider range in D than their presentday oceanic counterparts (Fig. 3), often due to postemplacement isotope exchange with crustal fluids, including meteoric waters. The highest D values, particularly those of antigorite, are commonly considered to have preserved their original isotopic compositions established during serpentinization by seawater in oceanic environments (10, 12, 27), and have been used to make inferences about the extent to which DSEA WATER changes over geologic time (11). Serpentines in the Isua Supracrustal Belt The Isua Supracrustal Belt represents the best preserved oceanic crust of Eoarchaean age (15, 28, 29). It has undergone regional greenschist- to amphibolite-facies metamorphism (SI Text), but both geochemical (15) and structural preservation (14) of protolith material occurs within localized low-strain exposures. For example, original bedding is visible in low-13 C carbon-rich marine sediments (30), and pillow lavas and sheeted dike structures associated with metagabbroic and serpentinized ultramafic rocks indicate that the ISB contains preserved fragments of oceanic crust (14, 29).
Author contributions: E.C.P., D.K.B., and M.T.R. designed research; E.C.P. performed research; E.C.P., D.K.B., and M.T.R. analyzed data; E.C.P., D.K.B., and M.T.R. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. Freely available online through the PNAS open access option.
1

To whom correspondence should be addressed. E-mail: emily@snm.ku.dk.

This article contains supporting information online at www.pnas.org/lookup/suppl/ doi:10.1073/pnas.1115705109/-/DCSupplemental.

PNAS Early Edition

1 of 6

EARTH, ATMOSPHERIC, AND PLANETARY SCIENCES

Fig. 1. Hydrogen and oxygen isotope values of Isua serpentines. ISB serpentines purple circles (Table 1). Shown relative to: modern ocean serpentinites [gray region; filled squaresantigorite, open squareslizardite and chrysotile, hatched squaresundifferentiated serpentine; from Hess Deep [East Pacific Ridge] (52), the Marianas Trench (23, 53), the Puerto Rican Trench, the Mid-Atlantic Ridge and the Blanco Fracture Zone (16)], the meteoric water line (MWL) and modern seawater composition (VSMOW), and the isotope composition of serpentine that would be in equilibrium with (1) modern-day seawater (black line), (2) Archaean seawater from this study (D 25 5; 18 O 2.3; purple region), and (3) Archaean seawater as proposed by Hren et al. (3) (D 60; 18 O 10; gray line). Based on the temperature-dependent fractionation of Saccocia et al. (19). Errors for Isua serpentines are 1 (SD) or analytical error, whichever is larger (see Table 1, Table S1 and Analytical Methods; N 24 for D, N 12 for 18 O).

Hydrogen and oxygen isotopes of the analyzed minerals distinguish two metasomatic events that have selectively altered ISB lithologies (Fig. S2). First, CO2 -rich fluids likely associated with ca. 3.81 3.74 Ga granitoid intrusions (14, 15, 28) intensely metasomatized ultramafic lithologies and adjacent amphibolites and felsic schists along the western and southern portions of the ISB. A second, later-stage meteoric-overprinting in the ISB is restricted to highly deformed regions and along contacts with late-stage intrusions (14). Isotopically, the first metasomatic event is distinguished by a large increase in 18 O (up to 18) with an insignificant change in D, and the second resulted in a progressive decrease in D (as low as 129) with very little change in 18 O. The isotopic trends associated with both events apparently originate at a common protolith isotopic composition (Fig. S2). Serpentinites are preserved in regions of the ISB not affected by the CO2 -metasomatic event (Fig. S1) (28). Antigorite, lizardite, and chrysotile occur with varying modes, and have a cumulative range in D of 99 to 53 and 18 O of 0.1 to 5.6 (Table 1; Fig. 1, purple circles). Serpentines with the highest D (samples from low-strain outcrops) exhibit little variation in 18 O, but samples with D 80 (highly deformed samples or those composed predominantly of chrysotile/lizardite) have 18 O that decrease with decreasing D. This hockey-stick trend in values of D and 18 O of hydrothermal minerals is characteristic of progressive meteoric-hydrothermal alteration (31), where decreasing D and 18 O values of crustal rocks reflects increasing water-rock (WR) ratios. Hydrogen isotopes in the rock are more significantly affected by secondary water-rock reaction than oxygen isotopes because of the greater oxygen content in rock-forming minerals. Thus at low WR mass ratios, DROCK will reflect isotope exchange with meteoric fluids but 18 OROCK will not, whereas at elevated WR ratios, both the 18 OROCK and DROCK will change.
2 of 6 www.pnas.org/cgi/doi/10.1073/pnas.1115705109

Fig. 2. Isotopic features of modern ocean serpentine. A. Hydrogen isotope composition of serpentine from various discrete locations (16, 21, 23, 52, 53). B. Histogram of 18 O values of modern ocean serpentinites (blue) (16, 21, 23, 5256). Also shown is the distribution of 18 O of ISB serpentinites (purple).

Hypothetical water-rock reaction paths between the most deuterium-rich serpentine (D 53, 18 O 4.8 to 5.6) and meteoric water (D 90, 18 O 12.5) are shown in Fig. 4. DMETEORIC-WATER was estimated based on approximately 40 fractionation (an approximate DMINERAL-H2 O for Fe-Mg minerals at 100300 C; SI Text) between meteoric water and ISB mineral samples from regions associated with late-stage intrusions or mylonitization. The lowest D value of all minerals analyzed in the ISB is 129 (Table S2). 18 OMETEORICWATER was calculated using the equation for the meteoric water line (32). With the exception of two outliers (16-05, 07-07, Table 1), the hydrogen and oxygen isotope properties of ISB serpentines are well approximated by a WR reaction path where temperatures are approximately 100200 C (Fig. 4). For samples with D > 85, WR ratios were likely 0.1 by weight, regardless of the temperature of isotope exchange. In view of these observations, we interpret serpentines with D > 85 as preserving oxygen isotope values close to their original composition.
Pope et al.

Fig. 3. D of serpentine in modern ocean crust and in ophiolites as a function of time. Antigorite solid squares, lizardite and chrysotile open squares, whole rock data crossed squares. Oceanic serpentines are from the same references as Fig. 1. Older serpentines are from ophiolite or greenstone belt sequences (1012, 16, 27, 5759, this study). Serpentine-forming fluids (D 25 5) at 3.8 Ga are the minimum estimate of Archaean seawater, based on equilibrium fractionation of unaltered ISB antigorite serpentine. Blue line represents the approximate increase in seawater D with time (calculated from average of values in Table 2) resulting from gradual, steady-state continental growth, and hydrogen escape before the rise of oxygen. The different slopes before and after approximately 2.4 Ga represent the elimination of appreciable hydrogen escape after the rise of oxygen (Fig. S7). We note that seawater evolution as represented by this line is approximately parallel to maximum D of serpentine over geologic time (purple line). Phan Phanerozoic.

The highest DSERPENTINE value recorded in the ISB provides a minimum estimate of their primary hydrogen isotope composition. However, it is possible that the serpentines with D 53 do not represent the initial rock composition as we model in Fig. 4, but have also been altered by meteoric water interaction. To test this, reaction paths between the same meteoric fluid and a hypothetical starting rock with D 30 (uppermost D of modern ocean antigorites) were also calculated. These paths are shown in Fig. S4A, and are less consistent with the Isua serpentine data than those in Fig. 4, but cannot be unequivocally ruled out.
Table 1. Isotope analyses of Isua serpentinites
Sample 03-10B 03-10C 15-05 03-10A 01-10 28-05 940094 07-08A 940095 09-08 20-07A 16-05 08-08 810000 07-07 06-07AA Mineral* A A+L A L L U A+L U U A+L L A+L L+C A U A+C D 53 53 56 56 61 69 69 73 74 78 84 85 91 95 98 99 1 0.5 0.7 1.9 1.3 1.8 2.2 2.2 1.1 0.2 1.3 4.0 2.1 1.5 1.9 1.2 0.2 18 O 4.6 4.1 4.9 5.3 4.1 5.1 3.5 4.8 4.5 5.2 4.1 5.6 2.9 4.4 0.1 2.7 1 0.28 - 0.07 1.13 0.42 0.35 0.35 0.42 0.78 0.42 0.42

We conclude that samples with the highest D (56 to 53) and 18 O between 4.1 and 4.9 (Fig. 4), have undergone minimal late-stage metasomatism (WR 0.01), and retain an isotope signature of seawater serpentinization of Eoarchaean oceanic crust. Critical to our interpretation are the paragenetic, petrofabric, and geochemical features of the most deuterium-rich ISB serpentine samples. These samples exhibit primary pseudomorph textures after peridotitic olivine (Fig. S5), and contain modally abundant antigorite that has a tendency to seal pathways for secondary fluids and inhibit hydrogen isotope exchange (SI Text) (12, 22, 23). We note that the four samples with D 56 are all from a region in the ISB that has been structurally and geochemically characterized as a well-preserved fragment of oceanic crust (Figs. S1, S3) (14, 29). Pons et al. (33) compared the source region for these serpentinized ultramafics to serpentinite mud volcanoes of the Mariana forearc on the basis of zinc isotopes. The presence of magnetite in the ISB serpentinites also indicates formation by low-temperature (50300 C) serpentinization on the ocean floor, as opposed to formation during high-grade serpentinization of the mantle wedge during crustal accretion (18, 34) (SI Text). Over the range of temperatures typical for antigorite formation in oceanic settings (250300 C, SI Text), samples 15-05, 03-10A, 03-10B, and 03-10C (Table 1) would form in equilibrium with fluids having a D between 28 and 21 (25 5, incorporating analytical error) and 18 O between 0.8 and 3.8 (Fig. 1, blue cross) (19). We consider these values to represent a minimum (the most negative possible) estimate for D of Eoarchaean seawater. Additionally, we note our modeled oxygen isotope ratios for early Archaean oceans (Fig. 1) are not significantly different from those of modern-day seawater (18 OVSMOW 0). As in modern hydrothermal alteration of ocean crust, ISB serpentine-forming fluids may be up to approximately 2.8 higher than starting seawater 18 O, because of isotopic exchange with oceanic crust (25). From Fig. 2B, it is clear that the 18 O distribution in ISB serpentinites (purple values) is consistent with that of modern ocean serpentinites (blue values), particularly for samples where D > 85. Models for a Global Hydrogen Budget Geologic processes that control the secular increase in DSEA WATER from 25 5 in the Eoarchaean to 0 in modern oceans can be evaluated through mass balance constraints on
PNAS Early Edition 3 of 6

Relative to VSMOW. *Primary phase: A Antigorite, L Lizardite, C Chrysotile, U all serpentine polymorphs. 1 1SD for N 24 samples. Does not include analytical error. 1 not applicable; N 1. Further details of analyses in Table S1.

Pope et al.

EARTH, ATMOSPHERIC, AND PLANETARY SCIENCES

Fig. 4. ISB serpentine data shown relative to hypothetical reaction paths between serpentines and meteoric water. Starting composition of rock is D 53 and 18 O 4.1 to 5.3 (highest D and 18 O values of ISB serpentines) and meteoric water is D 90 and 18 O 12.5. Reaction paths shown for varying temperature conditions, with water-rock ratios (WR) denoted. Errors for Isua serpentines are 1 (SD) or analytical error, whichever is larger (see Table 1, Table S1 and Analytical Methods; N 24 for D, N 12 for 18 O).

fluxes between global reservoirs and hydrogen loss to space (35). Global reservoirs include water contained in the modern ocean, trapped in hydrous minerals of the continental crust and sediment, in surficial and glacial waters, in the biosphere, in the hydrated ocean crust and in the mantle (Table S3). Relative to the other hydrogen reservoirs, the volume of water trapped in the biosphere is negligibly small. We assume that the extent of hydration of ocean crust by reaction with seawater has likely remained approximately constant over geologic time. In addition, we tentatively consider a net zero flux of water into and out of the mantle, as modern ingassing fluxes in subduction zones and outgassing fluxes at mid-ocean ridges are comparable (see Table S4, SI Text). Three model scenarios that consider fluxes among the remaining hydrogen reservoirs are presented in Fig. 5 as a function of total ocean mass and DSEA WATER , and are compared with values inferred from ISB serpentinites. In scenarios A and B, Earth behaves as a closed system for hydrogen. In A, all the hydrogen currently sequestered in the cryosphere, biosphere, surface- and ground-waters, sediments, and continental crust is allocated to the ocean reservoir, characterizing seawater before continental growth sequestered deuterium-poor fluids from the oceans into hydrous minerals (SI Text). In scenario B, 100% of modern continental crust is present, but it assumes nonglacial conditions. In these two scenarios, the ocean would have been 4 to 20% larger than modern oceans, and lower in deuterium by 918. Model scenario C is an estimate of the effect of hydrogen escape to space via methane photolysis in the preoxygenic Archaean atmosphere (36), given DCH4 -H2 O 150 to 300 (37), and diffusionlimited hydrogen escape, where no DH fractionation occurs (36, 38) (SI Text). The amount of hydrogen escape required to oxidize Earths crust indicates that preescape oceans would have been approximately 6% larger, and have a D 10 to 20 lower than Vienna standard mean ocean water (VSMOW). Here we consider the effects of both hydrogen sequestration in terrestrial reservoirs and hydrogen escape to space to account

Fig. 5. Hydrogen isotope composition and mass of ocean in the early Archaean. The terrestrial hydrogen budget [A, B; e.g., Lcuyer et al. (35)] and Catling et al. (36) data (C) represents the required D and mass of the ocean to satisfy the parameters of their models. The terrestrial hydrogen budget model represents a primordial ocean in which either (A) continental crust is not yet present (D 18), or (B) where 100% of modern continental mass exists (D 9). Model (C) estimates that oxidization of all modern continents, oceans and the atmosphere due to H-escape caused 6% mass loss from the ocean and enriched the ocean in deuterium by approximately 10 to 20 (36). Point (D) shows the mass of the ocean required to satisfy the Archaean DSEA WATER values predicted by chert analyses from Hren et al. (3), given hydrogen loss via escape to space from photolysis of methane (range in ocean mass results from uncertainty in methane-water fractionation). These data are shown relative to the minimum D values of the Archaean ocean determined from serpentine analyses (this study, Table 2), and the ocean mass as a function of the volume of continents present at 3.8 Ga (0 to 70%).

for the secular increase of DSEA WATER from 25 5 at ca. 3.8 Ga (see SI Text for model parameters, calculations, and uncertainties). The calculated amount of hydrogen that must be lost via escape to accommodate the estimated maximum and minimum DEOARCHAEAN OCEAN values based on ISB serpentinites (30 and 20) are compiled in Table 2, assuming different amounts of continental crust at approximately 3.8 Ga. If significant continental mass existed at approximately 3.8 Ga, the isotopic shift of seawater since that time would be dominantly controlled by hydrogen escape. If no continental crust existed at approximately 3.8 Ga, less hydrogen would have to escape to space to result in the observed change in D, as some low-D fluids would be incorporated into gradually growing continental crust. For our modeled range of DSEA WATER , ocean volumes would have been 9 to 26% greater than present-day, requiring a mass loss of H to space of between approximately 2 1021 mol and 18 1021 mol (Table 2; Fig. 5). Such a difference in mass may have a significant impact on current hypotheses of early Earth processes, including shallow mid-ocean ridge-crests and associated models for 18 O buffering of the Archaean ocean (1), continental freeboard in early Earth, and rates of heat and volatile transport via ocean spreading ridges (39). We consider the values determined from our hydrogen mass balance model to be reasonable first order approximations based upon average masses and isotope compositions of major hydrogen reservoirs on Earth (SI Text), which should be supported by additional evidence from the geologic record. Namely, changes in the isotopic composition of seawater with time will be apparent in other well-preserved rock suites. In Fig. 3, an extrapolation of DSEA WATER from present day to approximately 3.8 Ga based on the average values from Table 2 is shown as a blue line (SI Text). If Archaean DSEA WATER were lower; e.g., approximately 60 as is proposed by an alternative study based on the 3.4 Ga Buck Reef Cherts in South Africa (3), the slope of this line would be much steeper; if DSEA WATER had remained constant with time, the line would be horizontal. The average isotope composition of serpentine that would be in equilibrium with our modeled DSEA WATER is denoted by the parallel purple line, which is closely correlated to the maximum measured (and thus the bestpreserved) D values of serpentinite over geologic time. Although an ambitious study, the Buck Reef Cherts analyzed by Hren et al. (3) (also shown in Fig. 5 as scenario D) are a less well-suited proxy for seawater D than serpentine, resulting in the discrepancy between our two datasets. D of chert is measured from small concentrations of water (typically less than 1.4 wt% compared to up to 14 wt% in serpentine) that are susceptible to contamination with organic matter, clay minerals, or secondary fluids (40), and the temperature dependence of HD fractionation between chert and water is not straightforward (3, 41). Another important consideration is the relationship between the D and mass of Archaean oceans and the extent to which methane photolysis controlled oxidation of Earths surface (36). DSEA WATER of 60 (3) would result in early oceans that were 25 to 67% larger than their present size (Fig. 5), thus releasing 4 to 10 times as much oxygen by H escape from photolysis of biogenically produced methane than in Catling et al. (36)s calculation for oxidation of the crust, summarized by the reaction: CO2 2H2 O CH4 2O2 CO2 O2 4H space: [1] An additional oxygen sink, such as oxidation of the mantle via subduction, would be necessary to accommodate the oxidation associated with such an extreme hydrogen flux to space over Earth history. In contrast, the amount of oxygen released due to hydrogen escape from our models in Table 2 is 2.5 1.9 1021 mol O2 , a range consistent with the estimated 2.0 to 2.9 1021 mol O2 equivalent necessary to oxidize the Earths crust before the rise of oxygen at ca. 2.4 Ga (36).
Pope et al.

4 of 6

www.pnas.org/cgi/doi/10.1073/pnas.1115705109

Table 2. Ocean volume at approximately 3.8 Ga based on DSERPENTINE 54.5 1.5


Archaean DSEA WATER 20 % Continents Formed 0 15 70 100 0 15 70 100 H lost via escape (mol H)* 2.40 10 3.43 1021 7.18 1021 9.23 1021 11.93 1021 12.82 1021 16.11 1021 17.91 1021
21

O2 Equivalent (mol O2 ) 0.60 10 0.86 1021 1.80 1021 2.31 1021 2.98 1021 3.21 1021 4.03 1021 4.48 1021
21

Ocean size: (mol H2 O) 9.29 10 9.16 1022 8.71 1022 8.46 1022 9.76 1022 9.63 1022 9.16 1022 8.90 1022
22

(kg H2 O) 1.67 10 1.65 1021 1.57 1021 1.52 1021 1.76 1021 1.73 1021 1.65 1021 1.60 1021
21

(% modern) 119 118 112 109 126 124 118 114

30

*Elemental hydrogen.

Finally, our model indicates that hydrogen escape is limited to at most 18 1021 mol H, releasing approximately 4.5 1021 mol of free O2 to the Earth system (Table 2). If this much oxygen was released over approximately 1.4 billion years (3.8 to 2.4 Ga), then atmospheric CH4 at 3.8 Ga must have been at most 480 ppmv (64 to 480 ppmv for the range of O2 released in Table 2; SI Text) (36). Haqq-Misra et al., (42) has shown that at CH4 CO2 ratios above 1.0 or 0.2 (depending on atmospheric model), a climatically cooling organic haze would form, lowering surface temperatures. For an atmospheric mixing ratio of CH4 in the Archaean of approximately 64 to 480 ppmv, the pCO2 needed to maintain a haze-free atmosphere would be between 104.2 and 102.6 bar, or at a minimum approximately 0.2 to 6 times present atmospheric levels. These values are in the pCH4 and pCO2 ranges suggested by Precambrian paleosols, methanogenic metabolic constraints, and magnetite/siderite stability in early Archaean banded iron formations (9, 43, 44) (Fig. S7). 18 O of Eoarchaean Seawater The 18 OSEA WATER of Archaean oceans we calculate from ISB serpentinites (0.8 to 3.8; Fig. 1) is significantly higher than estimates based on carbonate and chert analyses that assume a temperate Archaean climate, which are as low as 13.3 (2, 3). Our values are, however, consistent with observations made from other Archaean volcanic rocks (13), including those from nearby pillow basalts in the ISB (14), from biogenic phosphates preserved in the 3.5 to 3.2 Ga Barberton Greenstone Belt (4), and from the ophiolite record of the past approximately 3.5 Ga (10). The progressive increase in 18 O of chemical sediments over geologic time to their modern average value of approximately 0 [PDB] we attribute to postdepositional exchange with hydrothermal or pore water fluids on the seafloor in accord with conclusions drawn from the Barberton phosphate data (4). In modern oceans, sediments are on average approximately 5001;000 m thick (45), and the near-surface geothermal gradient is about 25 Ckm. If the Earths heat flow in the early Archaean was approximately three times greater than at present (46, 47), then a thick package of seafloor sediments on the Archaean ocean floor could isotopically exchange with seawater-derived pore waters at temperatures approaching 70 C, as on average, the temperature gradient of the Earths crust will scale linearly with heat flow. A decline in Earths heat flow with geologic time could explain the progressive increase in 18 O of cherts and carbonates, although we note that such a decrease in heat flow over geologic time has been disputed (48, 49). Conclusions Petrogenetic and geochemical evidence preserved in ISB serpentinites constrains a minimum DSEA WATER of early Archaean oceans to 25 5 relative to present-day values (VSMOW). In the context of the metasomatic history of the ISB, the structural setting of the samples considered, the secular trend in hydrogen isotope compositions of serpentinites over geologic time, and the corroboration with independent models of surface oxidation and atmospheric methane levels in the Eoarchaean, these data appear to genuinely reflect seawater-oceanic crust
Pope et al.

interaction and are the best constraint thus far for D of early Archaean seawater. Mass balance considerations demonstrate that low-deuterium waters were primarily sequestered in continental crust during its progressive growth since the Hadean, and later also into groundwater and glacial reservoirs. At most, hydrogen lost to space is limited to 18 1021 mol H, releasing 4.5 1021 mol of free O2 to the Earth system, and constraining atmospheric methane concentrations at 3.8 Ga to 480 ppmv. This supports the argument that the combined greenhouse effect of atmospheric CO2 and CH4 cannot independently reconcile the faint early sun paradox. Additional forcing, such as a lower Earth-albedo (9), is necessary to maintain temperate conditions in the early Archaean. Oxygen isotope compositions of ISB serpentines suggest that they formed in the presence of seawater with a 18 O similar to modern oceans, consistent with oxygen isotope studies of Archaean biogenic phosphates (4), volcanic rocks from other Archaean greenstone Belts (13, 43), and mafic pillow lavas and sheeted dikes in the ISB (14). We therefore suggest that the low 18 O of Archaean chemical sediments is a result of postdepositional exchange with shallow ground- or pore waters. Analytical Methods Mineral separates were hand-crushed or powdered using a microdrill for hydrogen and XRD analyses; rock samples were crushed and sieved to a size of approximately 0.5 mm, then handpicked using a stereo microscope for oxygen analysis. Mineral purity was visually assessed, and samples with better than 95% purity were used for analysis. Hydrogen and oxygen isotope analyses of mineral separates of serpentine, talc, tremolite, anthophyllite, hornblende, biotite, and muscovite from ultramafic lenses and adjacent amphibolite and felsic rocks within all major lithologies in the ISB were performed at the Stanford University Stable Isotope Biogeochemistry Lab, following the methods of Sharp (50) and Sharp et al. (51). For hydrogen isotope analysis, 15 mg of powdered mineral separate were dried at vacuum for >24 h, then dropped into a Finnigan high temperature conversion elemental analyzer (TC-EA) using an autosampler flushed with helium. H2 gas produced from sample combustion in a 1,450 C carbon reduction furnace was introduced into a Finnigan DeltaPlusXL mass spectrometer in a helium gas stream. For oxygen isotope analysis, approximately 1 mg of mineral separates were dried at vacuum for >1 h in a nickel sample holder. Samples were heated using a CO2 -infrared laser in a vacuum chamber containing the oxidizing agent BrF5 . The oxygen gas was directly analyzed using a dual-inlet Finnigan MAT 252 mass spectrometer. Isotope compositions of the samples were corrected relative to National Bureau of Standards samples as well as laboratory standards, and are correct within 3.4 for hydrogen and 0.2 for oxygen of accepted values. Results of these analyses are presented in the standard delta notation as parts per thousand (), relative to the VSMOW standard. X-ray diffraction was performed on serpentine samples to distinguish antigorite, lizardite, and chrysotile polymorphs, at the Geballe Laboratory at Stanford University.
PNAS Early Edition 5 of 6

EARTH, ATMOSPHERIC, AND PLANETARY SCIENCES

ACKNOWLEDGMENTS. We thank P. Blisniuk and R. Jones for their assistance in data preparation and collection, and N. Sleep, C.P. Chamberlain, J. Stebbins, J.R. ONeil, and anonymous reviewers for their constructive comments. This research was supported by the Danish National Research Foundation
1. Kasting JF, et al. (2006) Paleoclimates, ocean depth, and the oxygen isotopic composition of seawater. EPSL 252:8293. 2. Jaffrs JBD, Shields GA, Wallmann K (2007) 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-Science Rev 83:83122. 3. Hren MT, Tice MM, Chamberlain CP (2009) Oxygen and hydrogen isotope evidence for a temperate climate 3.42 billion years ago. Nature 462:205208. 4. Blake RE, Chang SJ, Lepland A (2010) Phosphate oxygen isotopic evidence for a temperate and biologically active Archaean ocean. Nature 464:10291032. 5. Knauth LP, Epstein S (1976) Hydrogen and oxygen isotope ratios in nodular and bedded cherts. Geochim Cosmochim Acta 40:10951108. 6. Oskvarek JD, Perry EC, Jr (1976) Temperature limits on the early Archaean ocean from oxygen isotope variations in the Isua supracrustal sequence, West Greenland. Nature 259:192194. 7. Knauth LP, Lowe DR (2003) High Archean climatic temperature inferred from oxygen isotope geochemistry of cherts in the 3.5 Ga Swaziland Supergroup, South Africa. Geol Soc Am Bull 115:566580. 8. Robert F, Chaussidon M (2006) A paleotemperature curve for the Precambrian oceans based on silicon isotopes in cherts. Nature 443:969972. 9. Rosing MT, Bird DK, Sleep NH, Bjerrum CJ (2010) No climatic paradox under the faint early sun. Nature 464:744747. 10. Wenner DB, Taylor HP, Jr (1974) DH and 18 O16 O studies of serpentinization of ultramafic rocks. Geochim Cosmochim Acta 38:12551286. 11. Lcuyer C, Gruau G, Frh-Green GL, Picard C (1996) Hydrogen isotope composition of Early Proterozoic seawater. Geology 24:291294. 12. Kyser TK, OHanley DS, Wicks FJ (1999) The origin of fluids associated with serpentinization processes: Evidence from stable-isotope compositions. Canadian Mineralogist 37:223237. 13. Gregory RT (2003) Ophiolites and global geochemical cycles: Implications for the isotopic evolution of seawater. Geol Soc London Special Pub 218:353368. 14. Furnes H, de Wit M, Staudigel H, Rosing M, Muehlenbachs K (2007) A vestige of Earths oldest ophiolite. Science 315:17041707. 15. Frei R, Rosing MT, Waight TE, Ulfbeck DG (2002) Hydrothermal-metasomatic and tectono-metamorphic processes in the Isua supracrustal belt (West Greenland): A multiisotopic investigation of their effects on the Earths oldest oceanic crustal sequence. Geochim Cosmochim Acta 66:467486. 16. Wenner DB, Taylor HP, Jr (1973) Oxygen and hydrogen isotope studies of the serpentinization of ultramafic rocks in oceanic environments and continental complexes. Amer J Sci 273:207239. 17. Evans B (2004) The serpentinite multisystem revisited: Chrysotile is metastable. Int Geol Rev 46:479506. 18. Shaw AM, Hauri EH, Fischer TP, Hilton DR, Kelley KA (2008) Hydrogen isotopes in Mariana arc melt inclusions: Implications for subduction dehydration and the deep-Earth water cycle. EPSL 275:138145. 19. Sacoccia PJ, Seewald JS, Shanks WC, III (2009) Oxygen and hydrogen isotope fractionation in serpentine-water and talc-water systems from 250 to 450 C. Geochim Cosmochim Acta 73:67896804. 20. Sheppard SMF (1980) Isotopic evidence for the origins of water during metamorphic processes in oceanic crust and ophiolite complexes. Colloques Internationaux du CNRS 272:135147. 21. Barnes JD, Paulick H, Sharp ZD, Bach W, Beaudoin G (2009) Stable isotope (18 O, D, 37 Cl) evidence for multiple fluid histories in mid-Atlantic abyssal peridotites (ODP Leg 209). Lithos 110:8394. 22. Kyser TK, Kerrich R (1991) Stable Isotope Geochemistry: A Tribute to Samuel Epstein pp 409422 The Geochemical Soc. Spec. Pub. No. 3. 23. Alt JC, Shanks WC, III (2006) Stable isotope compositions of serpentinite seamounts in the Mariana forearc: Serpentinization processes, fluid sources and sulfur metasomatism. EPSL 242:272285. 24. Mvel C (2003) Serpentinization of abyssal peridotites at mid-ocean ridges. CR Geoscience 335:825852. 25. Bowers TS, Taylor HP, Jr (1985) An integrated chemical and stable-isotope model of the origin of midocean ridge hot spring systems. J Geophys Res 90:1258312606. 26. Jean-Baptiste P, Charlou J-L, Stievenard M (1997) Oxygen isotope study of mid-ocean ridge hydrothermal fluids: Implication for the oxygen-18 budget of the oceans. Geochim Cosmochim Acta 61:26692677. 27. Burkhard DJM, ONeil JR (1988) Contrasting serpentinization processes in the eastern Central Alps. Cont Min Pet 99:498506. 28. Rose NM, Rosing MT, Bridgwater DB (1996) The origin of metacarbonate rocks in the Archaean Isua supracrustal belt, west Greenland. Am J Sci 296:10041044.

through the Nordic Center for Earth Evolution, and endowment funds from the Department of Geological and Environmental Sciences at Stanford University. Portions of the research were supported by the Allan C. Cox Professorship to M.T.R.
29. Friend CRL, Nutman AP (2010) Eoarchean ophiolites? New evidence for the debate on the Isua supracrustal belt, southern West Greenland. Am J Sci 310:826861. 30. Rosing MT (1999) 13 C-depleted carbon microparticles in >3700-Ma sea-floor sedimentary rocks from West Greenland. Science 283:674676. 31. Criss RE, Taylor HP, Jr (1986) Stable Isotopes in High Temperature Geological Processes. Rev Mineralogy, eds JW Valley, HP Taylor, Jr, and JR ONeil pp:373424. 32. Craig H (1961) Isotopic variations in meteoric waters. Science 133:17021703. 33. Pons M-L, et al. (2011) Early Archean serpentine mud volcanoes at Isua, Greenland, as a niche for early life. PNAS 108:1763917643. 34. Evans BW (2010) Lizardite versus antigorite serpentinite: Magnetite, hydrogen, and life (?). Geology 38:879882. 35. Lcuyer C, Gillet P, Robert F (1998) The hydrogen isotope composition of seawater and the global water cycle. Chem Geol 145:249261. 36. Catling DC, Zahnle KJ, McKay CP (2001) Biogenic methane, hydrogen escape, and the irreversible oxidation of early Earth. Science 293:839843. 37. Schoell M (1980) The hydrogen and carbon isotopic composition of methane from natural gases of various origins. Geochim Cosmochim Acta 44:649661. 38. Catling DC (2006) Comment on A hydrogen-rich early Earth atmosphere. Science 311:38a. 39. Kasting JF, Holm NG (1992) What determines the volume of the oceans? Earth Planet Sc Lett 109:507515. 40. Knauth LP (1992) Isotopic Signatures and Sedimentary Records, Lecture Notes in Earth Sciences #43, eds N Clauer and S Chaudhuri (Springer-Verlag, Berlin), pp 123152. 41. Kolodny Y, Epstein S (1976) Stable isotope geochemistry of deep sea cherts. Geochim Cosmochim Acta 40:11951209. 42. Haqq-Misra JD, Domagal-Goldman SD, Kasting PJ, Kasting JF (2008) A revised, hazy methane greenhouse for the Archean Earth. Astrobio 8:11271137. 43. Hessler AM, Lowe DR, Jones RL, Bird DK (2004) A lower limit for atmospheric carbon dioxide levels 3.2 billion years ago. Nature 428:736738. 44. Sheldon ND (2006) Precambrian paleosols and atmospheric CO2 levels. Precamb Res 147:148155. 45. Divins DL (2010) NGDC total sediment thickness of the worlds oceans and marginal seas, http://www.ngdc.noaa.gov/mgg/sedthick/sedthick.html. Retrieved 21 May 2010. 46. Bickle MJ (1978) Heat loss from the Earth: A constraint on Archaean tectonics from the relation between geothermal gradients and the rate of plate production. EPSL 40:301315. 47. Abbott DH, Hoffman SE (1984) Archaean plate tectonics revisited 1. Heat flow, spreading rate, and the age of subducting oceanic lithosphere and their effects on the origin and evolution of continents. Tectonics 3:429448. 48. Korenaga J (2006) Archean Geodynamics and Environments, AGU Geophysical Monograph Series, eds K Benn, J-C Mareschal, and K Condie pp:732. 49. Herzberg C, Condie K, Korenaga J (2010) Thermal history of the Earth and its petrological expression. EPSL 292:7988. 50. Sharp ZD (1990) A laser-based microanalytical method for in situ determination of oxygen isotope ratios of silicates and oxides. Geochim Cosmochim Acta 54:13531357. 51. Sharp ZD, Atudorei V, Durakiewicz T (2001) A rapid method for determination of hydrogen and oxygen isotope ratios from water and hydrous minerals. Chem Geo 178:197210. 52. Agrinier P, Hekinian R, Bideau D, Javoy M (1995) O and H stable isotope compositions of oceanic crust and upper mantle rocks exposed in the Hess Deep near the Galapagos Triple Junction. EPSL 136:183196. 53. Sakai R, Kusakabe M, Noto M, Ishii T (1990) Origin of waters responsible for serpentinization of the Izu-Ogasawara-Mariana fore-arc seamounts in view of hydrogen and oxygen isotope ratios. EPSL 100:291303. 54. Bonatti E, Lawrence JR, Morandi N (1984) Serpentinization of ocean peridotites: temperature dependence of mineralogy and boron content. EPSL 70:8894. 55. Skelton ADL, Valley JW (2000) The relative timing of serpentinization and mantle exhumation at the ocean-continent transition, Iberia: constraints from oxygen isotopes. EPSL 178:327338. 56. Ribeiro da Costa I, Barriga FJAS, Viti C, Mellini M, Wicks FJ (2008) Antigorite in deformed serpentinites from the mid-atlantic ridge. Eur J Mineral 20:563572. 57. Magaritz M, Taylor HP, Jr (1974) Oxygen and hydrogen isotope studies of serpentinization in the Troodos ophiolite complex, Cyprus. EPSL 23:814. 58. Yui T, Yeh H, Lee CW (1990) A stable isotope study of serpentinization in the Fengtien ophiolite, Taiwan. Geochim Cosmochim Acta 54:14171426. 59. Liakhovitch VV, Quick JE, Gregory RT (2005) Hydrogen and oxygen isotope constraints on hydrothermal alteration of the Trinity peridotite, Klamath Mountains, California. Int Geol Rev 47:203214.

6 of 6

www.pnas.org/cgi/doi/10.1073/pnas.1115705109

Pope et al.

You might also like