Next Article in Journal
Equations of State of Ca-Silicates and Phase Diagram of the CaSiO3 System under Upper Mantle Conditions
Next Article in Special Issue
Thermodynamic Constraints on Smectite and Iron Oxide Formation at Gale Crater, Mars: Insights into Potential Free Energy from Aerobic Fe Oxidation in Lake Water–Groundwater Mixing Zone
Previous Article in Journal
Potential of Tailing Deposits in Chile for the Sequestration of Carbon Dioxide Produced by Power Plants Using Ex-Situ Mineral Carbonation
Previous Article in Special Issue
Thermodynamic Impact of Mineral Surfaces on Amino Acid Polymerization: Aspartate Dimerization on Two-Line Ferrihydrite, Anatase, and γ-Alumina
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stable Abiotic Production of Ammonia from Nitrate in Komatiite-Hosted Hydrothermal Systems in the Hadean and Archean Oceans

1
Institute for Extra-Cutting-Edge Science and Technology Avant-Garde Research (X-Star), Super-Cutting-Edge Grand and Advanced Research (SUGAR) Program, Japan Agency for Marine-Earth Science and Technology (JAMSTEC), Yokosuka 237-0061, Japan
2
Department of Subsurface Geobiological Analysis and Research (D-SUGAR), JAMSTEC, Yokosuka 237-0061, Japan
3
Research and Development (R&D) Center for Submarine Resources, JAMSTEC, Yokosuka 237-0061, Japan
4
Project Team for Development of New-Generation Research Protocol for Submarine Resources, JAMSTEC, Yokosuka 237-0061, Japan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Minerals 2021, 11(3), 321; https://doi.org/10.3390/min11030321
Submission received: 8 February 2021 / Revised: 6 March 2021 / Accepted: 16 March 2021 / Published: 19 March 2021

Abstract

:
Abiotic fixation of atmospheric dinitrogen to ammonia is important in prebiotic chemistry and biological evolution in the Hadean and Archean oceans. Though it is widely accepted that nitrate (NO3) was generated in the early atmospheres, the stable pathways of ammonia production from nitrate deposited in the early oceans remain unknown. This paper reports results of the first experiments simulating high-temperature, high-pressure reactions between nitrate and komatiite to find probable chemical pathways to deliver ammonia to the vent–ocean interface of komatiite-hosted hydrothermal systems and the global ocean on geological timescales. The fluid chemistry and mineralogy of the komatiite–H2O–NO3 system show iron-mediated production of ammonia from nitrate with yields of 10% at 250 °C and 350 °C, 500 bars. The komatiite–H2O–NO3 system also generated H2-rich and alkaline fluids, well-known prerequisites for prebiotic and primordial metabolisms, at lower temperatures than the komatiite–H2O–CO2 system. We estimate the ammonia flux from the komatiite-hosted systems to be 105–1010 mol/y in the early oceans. If the nitrate concentration in the early oceans was greater than 10 μmol/kg, the long-term production of ammonia through thermochemical nitrate reduction for the first billion years might have allowed the subsequent development of an early biosphere in the global surface ocean. Our results imply that komatiite-hosted systems might have impacted not only H2-based chemosynthetic ecosystems at the vent-ocean interface but also photosynthetic ecosystems on the early Earth.

1. Introduction

The reduction of abundant dinitrogen (N2) to ammonia (NH3) is important for pre-biotic chemical evolution and the early development of life on Earth. Ammonia participates in the synthesis of nitrogen-bearing organic compounds that are essential for cellular functions and self-reproduction (e.g., amino acids and nucleotides). Thus, the production of ammonia is assumed to be a prerequisite for origin-of-life scenarios, e.g., [1,2], and the development of microbial ecosystems in the Hadean and Archean oceans, e.g., [3,4,5]. Theoretical and observational assessments indicate that the main components of the early atmosphere are N2 and CO2, e.g., [6,7,8,9,10,11]. Thus, high energy reactions are required to reduce chemically stable N2. Previous studies have suggested that nitric oxide (NO) was produced by electrical discharges in the early atmosphere [3,12]. The resulting NO was likely converted into nitric acid (HNO3) or nitrous acid (HNO2) due to photochemical and aqueous-phase reactions, and it was eventually deposited in the ocean through rain [13]. This suggests ammonia was produced by the abiotic reduction of nitrate (NO3) and nitrite (NO2) in the early oceans because the oldest traces of early life were reported from 3.9–3.8 Ga sedimentary rocks of marine origin, e.g., [14,15,16,17,18,19,20], c.f. [21,22,23].
Iron-bearing materials are potential reductants for ammonia production from nitrate and nitrite in the early anoxic ocean. Accordingly, chemical reactions of nitrate and nitrite in the water column or marine sediments have been tested experimentally using ferrous ions and sulfate green rust (FeII4FeIII2(OH)12SO4·yH2O) at around 25 °C [24,25,26,27]. These experiments have provided the following three findings. (1) At 22 °C, Nitrite reduction by ferrous ions produces ammonia at pH ≥ 7.5, but ammonia production is suppressed in CO2-saturated water with respect to FeCO3. (2) Nitrate reduction by ferrous ions produces ammonia at pH > 8 in the presence of Cu2+. (3) Sulfate green rust reduces nitrite and nitrate to ammonia at pH 7–8.3. These findings suggest that the reduction of nitrite and nitrate at around 25 °C depends highly on the composition of the aqueous solution and hardly occurs at acidic pH. Because the early oceans are thought to be CO2-rich and sulfate-poor, with an acidic to slightly alkaline pH (ca. 5.5–7.5) [8,9,11,28,29,30], the reduction of nitrite and nitrate may have been inhibited in the water column or marine sediments in these oceans at around 25 °C.
Seafloor hydrothermal environments are another potential site of pre-biotic ammonia production. Abiotic reduction of nitrate and nitrite has been examined under simulated hydrothermal conditions with transition metal sulfides (Fe1−xS, FeS2, NiS, Cu2S, CuFeS2), oxides (Fe3O4, NiO, Cu2O), and iron-nickel alloys [31,32,33,34]. The results of these experiments have shown that some minerals participate in the reduction of nitrate and/or nitrite to ammonia, while others participate in the reduction of nitrate and/or nitrite to other forms of nitrogen (putatively N2 or N2O). Major anions (e.g., chloride, phosphate, sulfate, and dissolved carbonate species) further suppress nitrate reduction by Fe1−xS at room temperature due to competing adsorption onto Fe1−xS [31]. Given that concentrations of transition metal sulfides are typically found in hydrothermal mounds where high-temperature fluids mix with cold seawater, the reactions examined in previous studies [31,32,33,34] occur in nature under fluctuations and steep gradients of temperature, pH, and redox potential. These fluctuations and the associated unstable environment in hydrothermal mounds should result in a lower conversion to ammonia than in the laboratory experiments, with the exception of the relatively robust reduction of nitrite to ammonia catalyzed by MoS with geoelectrical current [35]. These results imply that other mechanisms for the sustainable supply of ammonia to the early oceans are favorable both for the pre-biotic chemical evolution in hydrothermal vent environments and the accumulation of ammonia in the global ocean.
This study focused on the reaction of nitrate with oceanic crust under high temperature and high-pressure conditions. Such reactions probably occur in the recharge and reaction zones of hydrothermal systems (i.e., sub-seafloor environments) and have affected the composition of vent fluids and the budget of nitrogen species in the early oceans. At present, the ammonia concentration in high-temperature hydrothermal fluids in sediment-free systems is typically similar to that of ambient deep-sea water (1 μm or less), although occasionally it is as high as 15 μm at certain locations [36,37]. These observations suggest the possibility of abiotic reduction of nitrate in seawater (40 μm) within oceanic crust. However, the simultaneous complex processing of nitrogen compounds by thermophilic microorganisms within hot oceanic crusts can obscure the signal of abiotic nitrate chemistry, as observed for nitrogen dynamics in geothermal environments [38]. Thus, an experimental approach is required to understand the abiotic chemistry of nitrate within hot oceanic crusts and the resultant vent fluid composition.
This paper reports the first experimental results of a simulation of high-temperature, high-pressure reactions between nitrate and komatiite to find chemical pathways that may have delivered ammonia to the vent–ocean interface and the global ocean on geological timescales. Komatiite is an ultramafic lava (MgO > 18%) produced by exceptionally hot melting in mantle upwellings, e.g., [39]. Most komatiite eruptions occurred between 3.5 and 1.5 Ga [40], and komatiite is believed to be the predominant ultramafic rock on the Hadean–Archean seafloor. The young Earth was hotter, and a thick layer of oceanic crust (more than three times greater than the present-day equivalents) probably restricted the eruption of mantle peridotites to the seafloor [41,42]. Therefore, rather than peridotite-hosted hydrothermal systems at slow-spreading mid-oceanic ridges, komatiite-hosted hydrothermal systems on oceanic islands and/or plateaus has been hypothesized to generate H2-rich and alkaline fluids [43], which are prerequisites for pre-biotic and primordial metabolisms at the vent–ocean interface, e.g., [43,44,45,46]. On the basis of the findings of the present study, the roles of komatiite-hosted hydrothermal systems on pre-biotic chemistry and implications for biological nitrogen metabolism and the global nitrogen cycle on the early Earth were explored.

2. Materials and Methods

Synthesis of komatiite, hydrothermal reaction experiments and chemical/mineralogical analyses of the reactants and products were conducted at Japan Agency for Marine-Earth Science and Technology (JAMSTEC). Details were provided in previous publications [47,48,49,50,51,52].

2.1. Starting Materials

Komatiite used in the experiments was synthesized from a mixture of 12 reagents (i.e., SiO2, TiO2, Al2O3, Fe2O3, MnO, MgO, CaCO3, Na2CO3, K2CO3, P2O5, NiO, and Cr2O3) by stepwise heating and cooling at oxygen fugacity of the quartz–fayalite–magnetite buffer. Mixing ratios of the reagents were adjusted to make Al-depleted (Barberton-type) komatiite more representative of ultramafic volcanism in Hadean oceanic islands and/or plateaus with a hotter mantle upwelling (plume) compared with Al-undepleted (Munro-type) komatiite [49]. The synthetic komatiite was primarily composed of olivine and glass. The komatiite was crushed using a tungsten mortar and an agate mill and then sieved to obtain particles smaller than 90 μm. Approximately 30 g of the powdered komatiite was ultrasonically washed with acetone and pure water (18.4 MΩ cm at 25 °C) several times and freeze-dried in a vacuum to remove possible contamination by organic matter during sample preparation. Specifically, carbon and nitrogen contents of the resultant komatiite sample were 210 ppm and 0.5 ppm, respectively [52].
Simulated aqueous solutions of Hadean seawater containing nitrate were prepared with the following compositions: 1000 mmol/kg NaCl and 10 mmol/kg NaNO3 for the experiment at 250 °C; and 1000 mmol/kg NaCl and 0.94 mmol/kg KNO3 for the experiment at 350 °C. Prior to use, the NaCl reagent was baked at 450 °C for 5 h under atmosphere to combust any organic contaminants.
We also conducted hydrothermal experiments without komatiite to test nitrate reduction in the H2O–NO3 system at 250 °C and 350 °C under 500 bars (experiments N250C and N350C). The initial aqueous solutions were prepared with the following composition: 1000 mmol/kg NaCl and 10 mmol/kg KNO3.

2.2. Hydrothermal Experiments

The hydrothermal reaction experiments were conducted in a flexible-cell hydrothermal apparatus consisting of a flexible gold bag with a titanium head (reaction cell) enclosed in an Inconel® alloy autoclave (modified after [53]). Prior to the experiments, the titanium head attached to the gold bag was completely oxidized to TiO2 to eliminate possible H2 generation by a reaction between metallic titanium and water at the experimental conditions. All materials in contact with the reacting fluid (i.e., gold bag, titanium head, and gold-lined sampling tube) were baked at 450 °C for 5 h under atmosphere in a muffle furnace to remove any organic contamination.
Approximately 10–12 g of the komatiite powder was initially reacted with 64–72 g of the simulated Hadean seawater in the reaction cell at 250 °C and 350 °C, under 500 bars (experiments N250 and N350). Changes in fluid composition were monitored by collecting 3–5 g of fluid samples multiple times during the ongoing experiment from the reaction cell through the gold-lined sampling tube while maintaining the temperature at 250 °C or 350 °C and the pressure at 500 bars. Consequently, the water–rock mass ratios decreased from 7 to 4 during experiment N250 (4870 h) and from 5 to 3 during experiment N350 (2515 h). These experimental conditions mimic the high-temperature regions in sub-seafloor hydrothermal systems where the water–rock mass ratios are generally below 5 [54]. After the experiments, solid reaction products were freeze-dried in a vacuum and then stored in a vacuum desiccator until mineral compositions could be analyzed.

2.3. Chemical and Mineralogical Analyses

2.3.1. Aqueous Fluids

For pH analysis, 1.0 mL of fluid was collected in a vial and the pH was measured by a pH meter (LAQUAtwin; HORIBA Ltd., Kyoto, Japan) at a room temperature under atmospheric conditions. The estimated precision of pH measurement was ±0.2 units of the reported value. For H2 analysis, 0.5 mL of fluid was introduced to an Ar-purged vial. The vial was left for 30 min to be reached in equilibrium at a room temperature. The H2 and N2O concentrations in the headspace of the vial was measured by gas chromatography with dielectric- barrier discharge ionization detector (GC-BID) (Nexis GC-2030, SHIMADZU Corp., Kyoto, Japan). The analytical precision (1SD) of replicate measurements was typically better than 5% for H2 and N2O at 100 ppm.
For cation analyses, 0.2 mL of fluid was sampled in a vial and diluted with 10 mL of pure water. The final pH of the diluted solution was adjusted to 2 by adding 0.2 mL of 1N HNO3 to avoid mineral precipitation. For anion analyses, 0.2 mL of fluid was collected and diluted with 10 mL of pure water. The final pH of the diluted solution was adjusted to 12 by adding 0.2 mL of 1N NaOH was added to avoid the precipitation of amorphous silica. The concentrations of Cl, and Na were analyzed by ion chromatography (Dionex ICS-1600/2100; Thermo Fisher Scientific, Waltham, MA, USA), while those of K, Ca, Mg, Fe, Mn, and Si were analyzed by inductively coupled plasma optical emission spectrometry (ICP-OES) (Spectro ARCOS; AMETEK, Berwyn, PA, USA). The analytical precision (1SD) of replicate measurements was typically better than 2% for Cl, Na, and K, and typically better than 5–10% for the other elements depending on their concentrations.
The nitrate and nitrite concentrations were measured using ion chromatography (IC-Pac A25S column; Thermo Fisher Scientific, Waltham, MA, USA) with UV detection at 220 nm (GL-7451; GL Science, Tokyo, Japan) (the detection limit was 0.1 μmol/kg and the reproducibility was better than 3%) [51]. The ammonium concentration in the reacted fluid was determined with fluorescence photometry after reacting at 50 °C with a 20 mmol/kg orthophthaldialdehyde solution and a 200 mmol/kg sodium borate solution containing 2 mmol/kg of sodium sulfite [52]. The analytical precision (1SD) of replicate measurements was typically better than 1%, 4%, and 3% for nitrate, nitrite and ammonium, respectively.

2.3.2. Solids

Initial characterization of the morphology and chemical composition of the solid reaction products was performed after the experiments using scanning electron microscopes (SEM) equipped with energy dispersive X-ray spectrometers (EDX) (a Hitachi High-Tech Miniscope® TM3000 and a FEI Quanta 450 FEG). Data on the mineral assemblages of reaction products were obtained by X-ray diffraction (XRD) and Raman spectroscopy using a Rigaku MiniFlex II and a Nanophoton RAMANtouch. Quantitative compositional data of the reaction products were obtained through electron probe micro analysis (EPMA) with wavelength dispersive X-ray spectroscopy using a JEOL JXA–8500F. The whole-rock compositions of 10 major elements (i.e., Si, Ti, Al, Fe, Mn, Mg, Ca, Na, K, and P) of the synthetic komatiite were analyzed by X-ray Fluorescence (XRF) using a Rigaku ZSX Primus II.
Three types of samples were prepared from the solid reaction products for the morphological, mineralogical and compositional analyses. The first group comprises fragments mounted in epoxy resin and polished by silicon carbide powder and diamond paste to determine the cross sections of the solid products. Samples in the second group were crushed using a stainless spatula, and the resulting sub-millimeter size grains were mounted on a glass slide with carbon tape or double-sided tape to examine 3D structures. Half of the second group samples were washed further with ethanol and purified water in an ultrasonic cleaner to remove the suspended fraction. The procedure was repeated until the ethanol and water were no longer cloudy. The washed samples were dried in a vacuum at 40 °C and then mounted on a glass slide. After examination using optical microscope and by Raman spectroscopy, the polished and crushed samples were coated with osmium or carbon particles for the analyses by SEM/EDX or EPMA, respectively. The third group of fragment samples was ground with agar mortar for XRD and XRF analyses. The resulting powder samples were processed further for XRF analysis to prepare glass beads using Li2B4O7 reagent.

2.3.3. Thermodynamic Calculation

The pHin-situ of the sampled fluids was calculated using the Geochemists’ Workbench computer code [55], based on pH at room temperature and the concentrations of dissolved species. Saturation index (SI) was calculated from the relation SI = log(Q/K), where Q is the activity quotient of the species involved in the reaction and K is the equilibrium constant. The required thermodynamic database was generated using the SUPCRT92 computer program [56] with thermodynamic data for aqueous species and complexes obtained from [57,58,59,60]. The B-dot activity model was used in the calculation of sampled fluids [61,62], and the activity coefficients of neutral species were assumed to be unity.

3. Results

3.1. Fluid Chemistry

Table 1 and Figure 1 present the concentrations of selected species measured for aqueous fluids in the komatiite–H2O–NO3 system as a function of time at 250 °C and 350 °C under 500 bars. The pH values of the initial solutions of experiments N250 and N350 after the pressurization to 500 bars at room temperature were approximately 9. During the high temperature water-rock reactions, the pH measured at room temperature (pH25°C) decreased from the initial values of 9.6 and 7.2 to the steady-state values of 8.6 (alkaline) and 5.7 (acidic) at 250 °C and 350 °C, respectively. “Steady state” here refers to the condition under which the concentration of a given dissolved species changes little with time [49,50]. The concentrations of almost all major species in the experiments (with the exception of H2 and Ca) reached a steady state within 300 h at both 250 °C and 300 °C.
The calculated in situ pH (pHin-situ) decreased from 7.7 and 7.1 to the steady-state values of 5.6 and 5.0 at 250 °C and 350 °C, respectively. As neutral pH values at the experimental conditions are 5.46 at 250 °C and 5.56 at 350 °C, the komatiite–H2O–NO3 system generated near neutral and acidic fluids at 250 °C and 350 °C, respectively.
The concentration of Cl remained relatively constant throughout the experiments because chlorine-containing minerals were rare in the altered komatiites. In contrast, the concentration of other species changed as the fluid/rock reactions progressed. Na concentrations steadily decreased by 30–40 mmol/kg from 990–1000 mmol/kg both at 250 °C and 350 °C, whereas K concentrations steadily increased by 1.5 and 1.1 mmol/kg from 0.3 and 1.0 mmol/kg at 250 °C and 350 °C, respectively. Ca and Si concentrations reached their maximum values and then decreased to the steady-state values (Si: 0.7 mmol/kg at 250 °C and 2.8 mmol/kg at 350 °C; Ca: 30 mmol/kg at 250 °C and 23 mmol/kg at 350 °C). The Si concentration started to decrease to the steady-state value when the Ca concentration increased to the maximum value (44–264 h at 250 °C, 17–71 h at 350 °C; Table 1).
Nitrate concentrations decreased from 11 mmol/kg to approximately 0.01 mmol/kg after 264 h at 250 °C, and decreased from 1 mmol/kg to approximately 0.005 mmol/kg after 71 h at 350 °C. Over the same time interval, ammonium concentration increased from 0.02 mmol/kg to 1.14 mmol/kg at 250 °C and 0.08 mmol/kg to 0.18 mmol/kg at 350 °C. Nitrite concentration increased from less than 0.002 mmol/kg to 0.011 mmol/kg in the first 18 h and then decreased to less than 0.002 mmol/kg at 250 °C, whereas it was less than 0.002 mmol/kg throughout the experiment at 350 °C.
The time-dependent changes of nitrogen compounds suggest that a multi-step pathway exists for the reduction of nitrate to ammonia via nitrite at 250 °C. This is comparable to the biological process referred to as dissimilatory nitrate reduction to ammonia. The mol ratio of ammonia produced to nitrate reduced (i.e., the product yield) was estimated to be 10% at both 250 °C and 350 °C. Thus, the other 90% of nitrate initially present in the fluid is expected to be reduced to other nitrogen species, most likely N2 via nitrous oxide (Section 4.2). The decrease in nitrate concentration appeared to be coupled with the increase in calcium concentration and with the decrease in pHin-situ at both 250 °C and 350 °C.
By comparison, experiments without komatiite show less reduction of nitrate at both 250 °C and 350 °C. In these experiments, nitrate concentrations were constant at 10 mmol/kg within the range of analytical uncertainty (±2%) and decreased slightly from 10 mmol/kg to 9 mmol/kg within 860 h and 264 h under 500 bars at 250 °C and 350 °C, respectively (Table 2, Figure 2). Over the same time intervals, nitrite concentration increased to 1.1 mmol/kg at 350 °C, whereas it was below detection limit at 250 °C. Nitrous oxide concentrations increased up to 0.01 mmol/kg and 0.05 mmol/kg at 250 °C and 350 °C, respectively. Ammonium concentrations were below the detection limit at both 250 °C and 350 °C. This indicates nitrate reduction to nitrite in the H2O–NO3 system at 350 °C. However, the rate of nitrate reduction is more than 40 times slower than that in the komatiite–H2O–NO3 system at 350 °C. Hence, komatiite causes or stimulates nitrate reduction and ammonia production in aqueous fluids at temperature above 250 °C and pressure of 500 bars.
The dissolved H2 concentration increased to steady-state values of 2.3 and 1.7 mmol/kg at 250 °C and 350 °C, respectively. By comparison, steady-state concentrations of the dissolved H2 concentration in the komatiite–H2O–CO2 system were reported to be 0.02 mmol/kg and 2 mmol/kg at 250 °C and 350 °C, respectively (Figure 3) [50]. The steady-state concentration of dissolved H2 in the komatiite–H2O system was reported to be 23 mmol/kg at 300 °C [49]. Table 3 compares the steady-state composition of fluids in the komatiite–H2O–NO3 system with that in the komatiite–H2O–CO2 system, as reported by [50]. Besides H2, the remarkable differences between the aqueous fluids in the two systems are the steady-state concentrations of Mg, and Fe and fluid pH at 250 °C. The dissolved Mg and Fe concentrations in the komatiite–H2O–NO3 system are 2–3 orders of magnitude lower than those in the komatiite–H2O–CO2 system. The pH measured at room temperature in the komatiite–H2O–NO3 system is more alkaline than that in the komatiite–H2O–CO2 system.

3.2. Alteration Products

The characterization of the solid products of the komatiite alteration experiments at 250 °C and 350 °C by X-ray diffraction, Raman spectroscopy, EPMA and SEM/EDX indicates that the products were dominated by a mixture of smectite, chlorite, and serpentine, and/or mixed layer clays such as chlorite/smectite and chlorite/serpentine (Figure 4 and Figure 5 and Table 4 and Table 5). Andradite and clinopyroxene were observed as characteristic major components at 250 °C and 350 °C, respectively. The reaction products exhibited the following characteristics:
  • The reaction products at 250 °C and 350 °C were mainly composed of 10–50-μm-size clusters of mineral plates/flakes. The plates/flakes developed an edge-to-face pattern and/or a rosette-like pattern (Figure 6e and Figure 7d,g). Such patterns are observed for the plates/flakes of chlorite, smectite and corrensite (mixed layered chlorite-smectite mineral) coated on a variety of silicate minerals in natural and experimental systems, e.g., [63,64,65]. Raman spectra of the clusters separated as the yellowish green or greenish grains from the bulk products showed peaks attributable to smectite, chlorite, serpentine and vermiculite (mixed layered chlorite-smectite mineral) (Figure 5h-1,h-3). The clusters also showed intermediate compositions between smectite (saponite), chlorite (clinochlore), and serpentine (Table 4 and Table 5). Collectively, we interpreted that the mineralogy of the clusters was a mixture of smectite, chlorite, and serpentine, and/or mixed layer clays such as chlorite/smectite and chlorite/serpentine (Table 4).
  • The clusters at both 250 °C and 350 °C had overgrowth rims extending outward from the center (Figure 6a,b and Figure 7a,b). The rims were composed of slightly curved flakes that developed honeycomb patterns with pore of 1–2 μm and 2–3 μm in size at 250 °C and 350 °C, respectively (Figure 6c and Figure 7b). Moreover, overgrowth rims with frost column textures partly covered the surface of the cluster at 250 °C (Figure 6e). There were not marked differences in elemental ratios between the rims and the phyllosilicate clusters at both 250 °C and 350 °C, suggesting that the rims had the same mineral composition as the clusters (Table 5).
  • Spherical (polygonal) minerals (ca. 1–3 μm in size) were found in the reaction products at 250 °C (Figure 6e and Figure 7a,b,f). They existed both in the cluster and the overgrowth rims. These minerals were identified as andradite (Ca3FeIII2Si3O12) based on their Raman spectra (Figure 5h-2). Results of the semi-quantitative analysis by SEM/EDX of the spherical minerals indicate elemental ratios that were consistent with those of andradite (Table 5).
  • Needle-shaped minerals occurred with the overgrowth rims in the reaction products at 350 °C (Figure 7a,b,f,h). The needles stuck together and formed an irregular-shaped, opaque grain that was separated as an unsuspended fraction after sonication (Figure 5d,g and Figure 7f). Raman spectra of these opaque grains indicated that the needles were clinopyroxene (Figure 5h-4). Semi-quantitative analysis by SEM/EDX of the needles showed elemental ratios that were consistent with clinopyroxene (Table 5).
  • Fibrous minerals were observed in the reaction products at 350 °C (Figure 7c,d). Regarding their morphology, they could be chrysotile fibers.
  • Needle-shaped calcium silicate occasionally occurred with andradite in the reaction products at 250 °C (Figure 6h). SEM/EDX analysis shows Ca/Si over 1 with Mg + Fe/Si smaller than 0.5 (Table 5). Clinopyroxene was also found among the reaction products at 250 °C as a minor component.
Based on the petrographic observations described above, clinopyroxene deposited after the formation of smectite–serpentine–chlorite core assemblage at 350 °C.

4. Discussion

4.1. Water-Rock Interactions in the Komatiite–H2O–NO3 System

Fluid composition and secondary mineral assemblages provide information on the water-rock interactions that include hydrogen generation and nitrate reduction. To aid the interpretation of reactions occurring during the experiments, saturation indexes of solid materials were calculated based on the measured fluid composition (Table 6). A list of potential water-rock reactions is also shown in Table 7.
In experiment N250, the concentration of dissolved Si (SiO2(aq)) increased to the maximum value (6 mmol/kg) during the first 8 h at 250 °C and then, remained at the maximum value from 8 h to 44 h (Figure A1). The maximum concentration of SiO2(aq) was 15% of the SiO2(aq) concentration predicted from the solubility of amorphous silica (Table 6), but it was ~100 times higher than the SiO2(aq) concentration under olivine dissolution at temperatures from 230 °C to 320 °C and apHin-situ of 8 [66]. Moreover, the positive SI values of forsterite and fayalite after the first 44 h (i.e., supersaturation with respect to forsterite) indicates that olivine dissolution had been inhibited in the first 44 h (Table 6). Therefore, the water-rock interactions that mainly occurred from 8 h to 44 h are predicted to be the dissolution of komatiite glass and the precipitation of secondary silicates. Given the low solubility of Al2O3 [67] and the fact that the sampled fluids had low concentrations of Mg and Fe, the precipitation of secondary silicates may have largely occurred through reactions involving solid phases of komatiite glass (rxns. 1–3 in Table 7). This inference is supported by the stable pHin-situ value in the first 44 h because reactions 1 to 3 did not affect pHin-situ much as compared with silicate formation from dissolved species (rxns. 6–8).
The time interval between 44 h to 264 h was characterized by the decrease in SiO2(aq) concentrations from the maximum value to the steady state value and the decrease in pHin-situ from 7.7 to 6.2. The decrease in the SiO2(aq) concentration suggests the completion of the dissolution of the SiO2 component in komatiite glass and the precipitation of secondary silicates. In particular, the decrease in pHin-situ requires the precipitation of dissolved Ca as Ca-bearing silicates (saponite, andradite, Ca-pyroxene) (rxns. 1, 4, 5) because the dissolved Ca concentration was high at the milli-molal level, while the dissolved concentrations of Mg and Fe were low over the same time interval. The precipitation of Ca-bearing silicates was consistent with the positive SI values of andradite and diopside (CaMgSi2O6) over the same time interval (Table 6). On the other hand, this was apparently inconsistent with the increase in the dissolved Ca concentration from 6 to 36 mmol/kg. This discrepancy can be resolved if the decrease in pHin-situ promoted the dissolution of the CaO component in glass (rxn. 15). Because andradite precipitation involves iron oxidation and H2 generation (rxn. 4), this reaction probably induced nitrate reduction between 44 h to 264 h (Section 4.2). The oxidation of the FeO component in komatiite glass also possibly contributed to H2 generation and nitrate reduction between 44 h to 264 h (rxns. 13–14).
The last interval between 264 h to 4871 h was characterized by the decrease in the Ca concentration to the steady state value (30 mol/kg; 266 h to 1054 h) and the increase in the H2 concentration to 2 mmol/g. The decrease in the Ca concentration was attributed to the formation of andradite because the SI value of andradite was positive until 1054 h, whereas the SI value of diopside became negative after 264 h (Table 6). This suggests that andradite formation had produced H2 until 1054 h. On the other hand, dissolution of fayalite, as inferred from the saturation of fluid with respect to fayalite since 1054 h (SI = 0), may have produced H2 since 1054 h by the oxidation of the FeO component in olivine (FeOol) with water to form magnetite and/or FeIII-bearing serpentine (rxns. 13–14) [68].
In experiment N350, the SiO2(aq) concentration in the first 17 h at 350 °C had the maximum value (20 mmol/kg), that is 63% of the SiO2(aq) concentration predicted from the solubility of amorphous silica (Figure A1). Over the same time interval, pHin-situ decreased from 7.1 to 5.8, whereas the Ca concentration increased to 10 mmol/kg. As discussed above, this suggests the dissolution of komatiite glass coupled with the formation of secondary silicates, which includes the precipitation of dissolved Ca as Ca-bearing silicates (saponite, Ca-pyroxene) (rxns. 1–5).
Subsequently, the SiO2(aq) concentration decreased from the maximum value to the steady state value (1 mmol/kg) from 17 h to 71 h. The decrease in the SiO2(aq) concentration was accompanied by a decrease in pHin-situ. Thus, this time interval was probably characterized by the complete consumption of the SiO2 component in komatiite glass and the precipitation of secondary silicates, including Ca-bearing silicates, which is consistent with the positive SI value of diopside at 17 h.
A main feature of the time interval between 71 h to 2515 h was the increase in the H2 concentration to 3 mmol/kg. The SI values of secondary minerals predicted that H2 generation during the entire interval of N350 can be driven by the oxidation of dissolved and solid phases of ferrous iron (Fe2+(aq), FeOrock = FeOglass, FeOol) to form magnetite and/or FeIII-bearing serpentine (rxns. 13–14). Because nitrate reduction to ammonia was completed by 71 h, H2 produced from iron oxidation may have been completely and/or partially consumed to reduce nitrate before 71 h.

4.2. Mechanisms of Nitrate Reduction to Ammonia and Denitrification in the Komatiite–H2O–NO3 System

This study strongly suggests that reactions in komatiite cause the complete reduction of nitrate to N2 and NH3 with a ratio of N2(aq)/NH3(aq) of 4.5 at both 250 °C and 350 °C, under pressures of 500 bars (Figure 8). The thermodynamic equilibrium of dissolved nitrogen species in the N–H–O system was evaluated first as a possible explanation for the N2(aq)/NH3(aq) ratio to characterize the mechanism of nitrate reduction in hydrothermal environments.
Figure 9a,b show the pHin-situ and aH2(aq) (or aO2(aq)) states of water with theoretically predicted stability fields of dissolved nitrogen species in thermodynamic equilibrium at 250 °C and 350 °C, 500 bars, respectively. The stability fields are drawn based on the activity of NH4+ in the final sampled fluid (aNH3(aq) = 10−3.228 and 10−3.98 at 250 °C and 350 °C, respectively). Bold red lines in the diagrams indicate the pHin-situ and aH2(aq) values at which the activity ratios of two nitrogen species (NH3(aq)–NH4+, NH3(aq)–N2(aq), NH4+–N2(aq), and N2(aq)–NO3) equal one in thermodynamic equilibrium. The pHin-situ and aH2(aq) values were plotted for all sampled fluids in the komatiite–H2O–NO3 and H2O–NO3 systems in these diagrams to compare the dominant nitrogen species in the experimental products with those from the thermodynamic prediction. Given that the very low H2 concentrations of sampled fluids in the H2O–NO3 systems were difficult to measure, the presumed aH2(aq) values were plotted in the diagrams assuming that sub-milli molal O2 in the initial solutions was not consumed during the experiments at 250 °C and 350 °C (e.g., dissolved O2 concentration is 0.2 mmol/kg in an air saturated seawater).
In the H2O–NO3 system, sampled fluids at 250 °C and 350 °C were predicted to locate the stability fields of nitrate and N2, respectively (purple and black boxes in Figure 9). In these experiments, nitrate at 250 °C was unreacted for 862 h, whereas nitrate at 350 °C was partially reduced for 264 h to nitrite and N2O, which were expected to be reduced to N2 via chemo-denitrification. The nitrate chemistry observed in the experiments is consistent with that predicted from the thermodynamic equilibrium of dissolved nitrogen species in the N–H–O system.
In the komatiite–H2O–NO3 systems, sampled fluids at 250 °C and 350 °C evolved to have higher aH2(aq) values than that in the H2O–NO3 systems, which are more favorable for ammonium production (green and yellow circles in Figure 9). The N2(aq)/NH3(aq) ratios in thermodynamic equilibrium at the experimental aH2(aq) and pHin-situ conditions are estimated to be less than 50 for fluid samples taken during the interval of nitrate reduction at 250 °C (the first 264 h; N250-1–N250-6), whereas they increase from less than 5,000,000 to 50 for fluid samples taken during the first 264 h at 350 °C (N350-1–N350-3). Therefore, theoretical prediction assuming thermodynamic equilibrium of nitrogen species in N–H–O system underestimates the amount of NH3(aq) produced from nitrate in the komatiite–H2O–NO3 systems by a factor of 10 and 10–1,000,000.
This suggests that nitrate reduction proceeds at the site of H2 generation where local aH2(aq) is higher than aH2(aq) of bulk fluids (e.g., heterogeneous reaction at the solid surface). It is also possible that nitrate reduction is mainly driven by a reductant other than H2.
Ferrous iron components in komatiite (FeOrock) are potential candidates for reaction sites to reduce nitrate. A general H2 generation reaction during serpentinization is written as iron oxidation reactions 14 and 15 in Table 7 [68]. In the komatiite–H2O–NO3 systems at 250 °C, the H2 generation reaction also occurs during andradite formation as discussed in Section 4.1 (rxn. 4). As discussed in Section 4.1, nitrate reduction in experiments N250 and N350 occurred during the time interval that was thermodynamically favorable to andradite formation (at 250 °C) and iron oxidation to form magnetite and/or FeIII-bearing serpentine (at both 250 °C and 350 °C). Thus, we suggest the following reactions as nitrate reduction in the komatiite–H2O–NO3 system (see Appendix A for details).
  • Nitrate reduction during serpentinization (T = 250 °C and 350°C, P= 500 bars):
2NO3 + 2H+ + 15FeOrock = N2(aq) + 5Fe3O4 + H2O
2NO3 + 2H+ + 24FeOrock + 2H2O = 2NH3(aq) + 8Fe3O4
2NO3 + 2H+ + 10FeOrock + 10SiO2(aq) + 9H2O = N2(aq) + 5FeIII2Si2O5(OH)4
2NO3 + 2H+ + 32FeOrock + 32SiO2(aq) + 40H2O = 2NH3(aq) + 16FeIII2Si2O5(OH)4
  • Nitrate reduction coupled with andradite formation (T = 250 °C, P = 500 bars):
2NO3 + 15Ca2+ + 10FeOrock + 15SiO2(aq) + 14H2O = N2(aq) + 5Ca3FeIII2Si3O12 + 28H+
2NO3 + 24Ca2+ + 16FeOrock + 24SiO2(aq) + 26H2O = 2NH3(aq) + 8Ca3FeIII2Si3O12 + 46H+
The reactions 19 to 24 represent nitrate reduction by both H2 and FeII at the site of serpentinization and andradite formation. They probably contribute to a lower N2(aq)/NH3(aq) ratio in the komatiite–H2O–NO3 system than expected due to the thermodynamic equilibrium of dissolved nitrogen species in N–H–O system at 250 °C and 350 °C, 500 bars.

4.3. H2 Generation Ability of Komatiite-Hosted Hydrothermal System in Early Earth

Previous simulations of komatiite hydrothermal alteration in the early ocean [49,50] typically considered two types of experiments, i.e., those under CO2-bearing and CO2-free conditions. These contrasting conditions not only reflect geological records (such as strongly carbonated komatiites and carbonate-free, serpentinized komatiites) [69], but also the two types of fluid circulation paths. The first type addresses the direct reactions with Hadean–Archean seawater containing high concentration of ΣCO2 [50]; this is similar to sulfate in modern seawater introduced directly through faults into reaction zones at the base of seafloor hydrothermal systems, forming sulfate minerals that strongly affect the redox state of hydrothermal fluid, e.g., [70,71].
The second type assumes reactions between komatiite and CO2-free fluid in a reaction zone where CO2 in downwelling seawater has been mostly precipitated as carbonate minerals in the surrounding rocks (e.g., recharge zone) before the fluid reaches reaction zones [49]. Previous studies revealed that the formation of carbonate minerals in the reaction zone suppressed H2 generation given that most of iron originally contained in igneous phases was incorporated directly into carbonate minerals as Fe(II) without going through oxidation and reduction processes in water that generate H2 [50]. The experiments in this study were a simulation of the CO2-free hydrothermal alteration of komatiite in a reaction zone, which has higher reduction abilities than those in CO2-bearing systems.
In this case, a certain amount of H2 generation in the fluids was expected in the CO2-free experiments. However, the H2 concentrations in fluid in N350 is 1.7 mmol/kg, which is almost identical to or lower than that in CO2-bearing experiments at 350 °C. On the contrary, the H2 concentration in N250 is 100 times higher than that of the CO2-bearing experiments at 250 °C (Table 3). This indicates that the reduction ability of komatiite was partially used for the reduction of nitrate during the experiments, as discussed in Section 4.2. Accordingly, the potential H2 concentrations were calculated for the case where nitrate was not included in the initial solutions to assess the maximum H2 generation ability of komatiite during its alteration at 250 °C and 350 °C. Using the N2(aq)/NH3(aq) ratio of 4.5 in experiments N250 and N350, the presumed mass balance reaction of nitrate reduction by H2(aq) that is generated from the reaction of ferrous iron with water can be expressed as
20NO3 + 20H+ + 53H2(aq) = 9N2(aq) + 2NH3(aq) + 60H2O.
This reaction implies that the reduction of nitrate in the initial fluid requires 2.6 times H2(aq), suggesting that an additional 26 mmol/kg and 2.6 mmol/kg of H2(aq) were potentially generated in N250 and N350, respectively. Certainly, these values should be upper limits for H2 concentrations given than nitrate can be reduced by ferrous iron (Section 4.2.). However, the potential H2 concentrations in N250 and N350 and the results of the previous CO2-free experiment at 300 °C (H2 = 23 mmol/kg) [49] suggests that the optimal temperature range for H2 generation is 250–300 °C during the serpentinization of Al-depleted komatiite in CO2-free conditions. This temperature range is comparable to the optimal temperature range for H2 generation during the serpentinization of peridotites inferred from thermodynamic constraints [68,72].

4.4. Roles of Komatiite-Hosted Hydrothermal System on Pre-Biotic Chemistry

Komatiite-hosted hydrothermal systems have influenced pre-biotic chemistry and affected the sustainability of primitive microbial communities through the production of H2, alkalinization of fluids, and ammonia production in the komatiite–H2O–NO3 system.
H2 is a key molecule for pre-biotic chemistry and energy metabolism. Theoretical predictions for the thermodynamic state of pre-biotic chemical evolution have highlighted that an H2-rich hydrothermal environment was energetically advantageous for the synthesis and preservation of amino and fatty acids [73]. Further, the most plausible energy metabolisms to support the emergence and early evolution of life are considered to be H2-driven hydrogenotrophic methanogenesis/acetogenesis and/or methanotrophic acetogenesis, e.g., [43,45,46]. On the basis of fluid chemistry and mineralogy in the komatiite–H2O–CO2 system, Ueda et al. (2016) have suggested that H2-rich fluid might not have been produced by the hydrothermal alteration of komatiite under CO2-bearing conditions at temperatures ≤ 250 °C [50]. This implies that hydrothermal systems on the deep seafloor (more than 1000 m below the sea surface) were favorable for the sustainability of primitive microbial communities as high hydrostatic pressure was necessary to create high-temperature fluids. In contrast, the steady-state H2 concentrations in the komatiite–H2O–NO3 system at 250 °C were 2.3 mmol/kg for experiment N250 and could be as high as 26 mmol/kg if the initial nitrate concentration is lower than 10 mmol/kg (Section 4.3). Such hydrogen concentrations are comparable to those in high-temperature hydrothermal fluids venting from a modern, peridotite-hosted system that harbors H2-based chemosynthetic microbial communities, including methanogens (H2 = 3–16 mmol/kg) (e.g., Kairei and Rainbow Fields) [74,75]. Thus, H2-based, primitive microbial communities in the Hadean ocean may have developed in komatiite-hosted hydrothermal systems of lower fluid temperature (down to 250 °C) than previously thought.
Alkaline fluids were necessary for primordial metabolisms driven by chemiosmosis and electrochemistry at the vent–ocean interface (i.e., hydrothermal mounds), where hydrothermal fluids mixed with acidic and/or neutral, CO2-rich Hadean seawater [76,77,78]. For example, the selective synthesis of alanine from pyruvate and ammonia (>30% in yield) occurs in water with pH of 9–11 at 70 °C in the presence of iron oxyhydroxides [76]. Certain steps in the reductive tricarboxylic acid cycle are promoted by FeS-Fe0 assemblages that can be produced electrochemically in alkaline water with pH of 10 at 100 °C in the presence of 1 mmol/kg H2 [78]. This study shows that the pH of fluid in the komatiite–H2O–NO3 system becomes more alkaline than that in the komatiite–H2O–CO2 system at 250 °C (pH25°C = 8.7 and 6.0, respectively) (Table 3). Thus, hydrothermal fluids generated in the komatiite–H2O–NO3 system may have been more favorable to certain types of primordial metabolisms at the vent–ocean interface where the reaction zone temperature was around 250 °C. In contrast, when the temperature of the reaction zone rises over 350 °C, the hydrothermal fluids in the komatiite–H2O–CO2 system become more favorable for primordial metabolisms as the destabilization of carbonate minerals above 350 °C stimulates the alkalinization effect by ΣCO2 in the aqueous fluid, producing an alkaline pH in the CO2-bearing fluid [48,50,79].
Ammonia is produced from nitrate at a high yield (10%) in the komatiite–H2O–NO3 system both at 250 °C and 350 °C (Table 1). Ammonia is a key reactant for synthesizing amino and nucleic acids, and previous research successfully produced organic nitrogen compounds (amides, amino acids, and nitriles) by reacting ammonia with a variety of organic compounds (e.g., methane, formic acid, oxalic acid, and pyruvate) under limited hydrothermal conditions with and without minerals [76,80,81,82,83]. As the fluid residence time from the onset of the high-temperature (>200 °C) reaction to venting at the seafloor (several years for present-day hydrothermal systems) [84,85] is much longer than the time required for complete reduction of nitrate at ≥250 °C (<300 h) (Table 1), a stable supply of ammonia to hydrothermal environments could be expected for the komatiite–H2O–NO3 system in the Hadean ocean. Although the actual concentration of nitrate in the early oceans (thus, ammonia concentration in hydrothermal fluids) may have been lower than that used in our experiments, the stable supply of ammonia may have allowed abiotic synthesis of organic nitrogen compounds at least in localized environments that can concentrate ammonia such as pore spaces of clay mineral clusters in altered komatiite crusts, and interiors of iron oxyhydroxides in hydrothermal mounds (e.g., Figure 6g). However, further studies are necessary to explore pathways for the synthesis of organic nitrogen compounds that are resistant to the fluctuation of environmental conditions (pH, temperature and coexisting mineral type) that the hydrothermal vent environments probably encountered.

4.5. Implications for Biological Nitrogen Metabolism, Phytosynthetic Activity, and the Global Nitrogen Cycle in the Early Earth

The stable and high-yield production of ammonia in the komatiite–H2O–NO3 system is an important process not only for pre-biotic chemistry in the hydrothermal environments but also for controlling ammonia concentrations in the early global ocean that probably influenced biological nitrogen metabolism, the development of photosynthetic ecosystems, and global nitrogen cycle in the early Earth. Approximations of the ammonia production rate in the komatiite-hosted hydrothermal systems in the early ocean, FNH3,ko (mol/y), suggest that biological activity had started in hydrothermally influenced local environments, but was not likely widespread throughout the global surface ocean. The FNH3,ko (mol/y) value can be estimated from the following relation:
FNH3,ko = YNH3,ko FH2O,ko CNO3,
where YNH3,ko is the yield of ammonia by thermochemical nitrate reduction in the komatiite-hosted hydrothermal systems, FH2O is the high-temperature (≥250 °C) fluid flux from the komatiite-hosted hydrothermal systems (kg/y), and CNO3 is the nitrate concentration in seawater (mol/kg). Our experiments suggest that YNH3,ko value is 0.1, whereas numerical modeling by Laneuville et al. (2018) indicated that CNO3 in early seawater was dependent upon atmospheric nitrate production rate and had a range from 10−6 to 10−2 mol/kg [86], cf. [87,88]. The FH2O value is initially approximated below, followed by the FNH3 value.
In the modern ocean, hydrothermal fluid flux is considered proportional to the annual production rate of oceanic crust at fast-spreading ridges [89]. Numerical modeling also suggests that plate velocity has not changed much throughout Earth’s history [90,91]. Thus, the high-temperature fluid flux from the komatiite-hosted hydrothermal systems can be approximated as follows:
FH2O,ko = k Pko,
where Pko is the annual komatiite production rate (km3/y) and k is a coefficient that relates high-temperature fluid flux to Pko (kg/km3). Field observations suggest that the thickness of the Archean oceanic crust was approximately three times greater than modern equivalents owing to the higher degree of partial melting [41,42]. Further, numerical modeling suggests that the global spreading rate of Archean oceanic crust (the product of plate velocity and plate boundary length: km2/year) was approximately three times higher than that at present [90]. Thus, the annual production rate of Archean oceanic crust was estimated to be 180 km3/y, that is, nine times greater than at present [92]. If the production ratio of oceanic plateau to oceanic crust at mid-oceanic ridges in the Archean was identical to that at present (0.05), the Pko value was estimated to be 10 km3/y in the Archean, which is nine times greater than that of the present oceanic plateau (1.1 km3/year) [93]. Given that high-temperature (>250 °C) fluid flux in the modern ocean is estimated to be 0.2–3 × 1013 kg/y, e.g., [94,95,96], k is calculated to be 1011–1012 kg/km3. Collectively, the FH2O,ko and FNH3,ko values of the Archean ocean are estimated to be 1012–1013 kg/y and 105–1010 mol/y, respectively (Table 8). Given higher heat flux from Earth’s interior in the Hadean than Archean, the FH2O,ko and FNH3,ko values of the Hadean ocean are considered to be greater than the Archean ocean.
The FNH3,ko value provide insights into the evolution of biological nitrogen metabolism. It has to be noted that our estimate of the FNH3,ko value is based on simple equations and several assumptions, thus allowing the FNH3,ko value to have a wide range from 105 to 1010 mol/y. Nonetheless, our estimate indicates that the ammonia concentration in the early oceans had increased to 0.01–1000 μmol/kg from 4.4 Ga to 3.4 Ga if ammonia consumption rate in the ocean (e.g., biological assimilation) was much lower than the ammonia production rate. Microorganism can assimilate ammonia at the micro-to-milli molal levels for growth. For example, a half saturation constant for ammonia uptake has been reported to be 4–74 μmol/kg for bacteria isolated from different habitats [97]. Thus, if nitrate concentration in the early oceans was greater than 10 μmol/kg, the long-term production of ammonia through thermochemical nitrate reduction for the first billion years might have allowed the subsequent development of an early biosphere in the global surface ocean. In contrast, if nitrate concentration in the early oceans was less than 1 μmol/kg, komatiite-hosted hydrothermal activity would not have contributed to the accumulation of ammonia in the global ocean. In this case, the onset of biological nitrogen fixation (N2 → NH3) was required for the development of early biosphere, consistent with genomic, physiologic, and geological records that are compatible with nitrogen fixation by the last universal common ancestor or methanogen in hydrothermal environments at 3.5 Ga [4,5,98,99], cf. [100]. To determine which scenario is more appropriate for the evolutional history of nitrogen metabolism, future work is required to estimate nitrate concentration in early seawater from geological records. Nitrogen isotopic characterization of abiotically produced ammonia is also important to determine the source of nitrogen isotopic compositions in Archean sedimentary rocks of marine origin, which have been interpreted as evidence for biological nitrogen fixation in surface oceans at 3.2 Ga [101] on the basis of nitrogen isotopic similarity to N2 in the Archean atmosphere (ca. 0‰) [10,102,103,104].
Results of our experiments also suggest that 90% of nitrate nitrogen introduced to the komatiite-hosted hydrothermal system had probably been converted to N2, resulting in a larger input of N2 to the early ocean than that of ammonia. The higher input of N2 may have stabilized the partial pressure of N2 (pN2) in the atmosphere and climate in an era of low solar luminosity as the increase in pN2 would have increased the warming effect of existing greenhouse gases in an anoxic atmosphere (CO2, CH4) [105]. The approximate global N2 flux from thermochemical nitrate reduction in early Archean seafloor hydrothermal systems, including both komatiite- and basalt-hosted systems, is FN2 (mol/y), which can be compared to the N2 fluxes from other processes to develop a quantitative assessment of the contribution of thermochemical nitrate reduction to pN2 in the early Archean.
The estimated FN2 value in the early Archean includes the estimate of the N2 production rate in the komatiite-hosted hydrothermal system in the early Archean, FN2,ko (mol/y) using the following equation:
FN2,ko = YN2,ko FH2O,ko CNO3,
where YN2,ko is the yield of N2 by thermochemical nitrate reduction in the komatiite-hosted hydrothermal systems. By substituting 0.45, 1012–1013 kg/y, and 10−6–10–2 mol/kg for YN2,ko, FH2O,ko, and CNO3 in Equation (28), respectively, FN2,ko is calculated to be 106–1011 mol/y. If the production ratio of oceanic plateau to oceanic crust at mid-oceanic ridges in the Archean was identical to that at present (0.05), high-temperature fluid flux from the basalt-hosted hydrothermal system at Archean mid-oceanic ridges is one order magnitude larger than FH2O,ko. Accordingly, FN2 is estimated to be 107–1012 mol/y (i.e., 108–1013 gN/y) in the early Archean, if the yield of N2 by thermochemical nitrate reduction in the basalt-hosted system YN2,mo had been equal or greater than YN2,ko, which is an appropriate assumption due to lower potentials of H2 and NH3 generation in basalt-hosted system than komatiite-hosted system.
When comparing N2 fluxes from major processes, the FN2 value in the early Archean is quantitatively important. Major processes that supply N2 to the present atmosphere are biological denitrification in the global ocean (NO3 → N2) (1–3 × 1014 gN/y) [106], degassing from the MORB source mantle (3 × 1010 gN/y) [107] and degassing by arc volcanism (3 × 1011 gN/y) [108]. In the early Archean, geological records suggest that biological denitrification was not yet underway, whereas degassing fluxes of N2 from the MORB source mantle and via arc volcanism could have been one order of magnitude greater than today as they change proportionally with the annual production rate of oceanic crust. Therefore, N2 flux by thermochemical nitrate reduction in the global hydrothermal systems would have been greater than that from the Earth’s interior if nitrate concentration in Archean seawater had been greater than 1000 μmol/kg.
In summary, the approximate N2 fluxes to the atmosphere in the early Archean from hydrothermal systems by thermochemical nitrate reduction and from the Earth’s interior were 108–1013 gN/y and 1012 gN/y, respectively. Meanwhile, fluid inclusions in the Archean rock samples indicate that the abundance of N2 in the Archean atmosphere was 2–12 × 1021 gN, which is 0.5–3 times the present atmospheric N2 [10,103,104,109]. Therefore, the residence time of atmospheric N2 in the early Archean is calculated to be 108 y or greater. For comparison, the residence time of atmospheric N2 at present is 107 years due to the high rates of biological nitrogen fixation and biological denitrification in the global ocean (1–3 × 1014 gN/y) [106]. Although pN2 and its secular variation during Earth history have not yet been estimated precisely, our simple estimate implies that pN2 is stable for ≥108 years in the early Archean if the global nitrogen cycle was driven primarily by abiotic nitrogen processes in the atmosphere and seafloor hydrothermal systems. Together with the emission of hydrogen and methane gases during serpentinization, this point should be considered when evaluating the role of hydrothermal activity in the long-term climate of the early Earth.

5. Conclusions

We conducted hydrothermal experiments simulating the high-temperature, high-pressure reactions between nitrate in seawater to find probable chemical pathways to deliver ammonia to the vent–ocean interface of komatiite-hosted hydrothermal systems and the global ocean on geological timescales. Our experiments provide the following key results and conclusions.
  • In the komatiite–H2O–NO3 system, ammonia was produced from nitrate with yields of 10% at 250 °C and 350 °C in a CO2-free condition. The fluid chemistry and mineralogy suggest that nitrate reduction co-occurred with serpentinization and andradite formation at 250 °C, whereas it co-occurred with serpentinization at 350 °C. This represents a case where CO2 in downwelling fluid precipitated as carbonate minerals in surrounding rocks before reaching the high-temperature reaction zones.
  • Nitrate reduction by both H2 and FeIII at the site of serpentinization and andradite formation probably contributed to a lower N2(aq)/NH3(aq) ratio (4.5) than expected due to the thermodynamic equilibrium of dissolved nitrogen species in N–H–O system at 250 °C and 350 °C, 500 bars.
  • H2-rich and alkaline fluids, which are prerequisites for both prebiotic and primordial metabolisms, were generated in the komatiite–H2O–NO3 system at lower temperatures (down to 250 °C) than previously thought under CO2-rich conditions. The komatiite–H2O–NO3 system also provided a stable supply of ammonia to the hydrothermal environments for biomolecule synthesis. Given that the temperature of hydrothermal fluids depends on hydrostatic pressure, the komatiite–H2O–NO3 system probably reduces the minimum water depth of seafloor hydro-thermal systems to about 1000 m, which may have favored the start of pre-biotic chemistry and the sustainable development of H2-based microbial ecosystems at the vent-ocean interface of the Hadean ocean.
  • The ammonia production rate in the komatiite-hosted hydrothermal systems depends on nitrate concentration in seawater and is estimated to be 105–1010 mol/y in the Archean ocean. The rate in the Hadean ocean was likely to be greater than the Archean ocean given higher heat flux in the Hadean ocean. If nitrate concentrations in the early oceans were greater than 10 μmol/kg, the long-term production of ammonia by thermochemical nitrate reduction for the first billion years might have allowed the subsequent development of an early biosphere in the global surface ocean through biological ammonia assimilation.
  • The N2 production rate in the komatiite-hosted hydrothermal systems is estimated to be 106–1011 mol/y in the Archean ocean. The estimate implies that pN2 in the Archean atmosphere is stable at least for 108 years if abiotic nitrogen processes in the atmosphere and seafloor hydrothermal systems mainly drove the global nitrogen cycle.
Collectively, these findings imply that komatiite-hosted systems might have impacted not only H2-based chemosynthetic ecosystems at the vent-ocean interface but also photosynthetic ecosystems on the early Earth.

Author Contributions

Conceptualization, M.N., T.S. (Takazo Shibuya) and K.T.; experiments, M.N., A.M., T.S. (Takuya Saito) and M.S.; modeling, H.U. and T.S. (Takazo Shibuya); data curation, A.M.; writing—original draft preparation, M.N., H.U., T.S. (Takuya Saito) and T.S. (Takazo Shibuya); writing—review and editing, K.T.; project administration, M.N.; funding acquisition, M.N., M.S., T.S. (Takazo Shibuya) and K.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Grants-in-aid for Scientific Research from the Japan Society for Promotion of Science (KAKENHI JSPS), grant numbers JP15H03740, JP15H03741, JP17H06455, and JP17H01176, and the Mitsubishi Foundation, grant number 28120. The APC was funded by JP17H06454.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The Authors thank M. Uematsu, T. Arai and T. Sato for their technical supports of SEM/EDS and XRF analyses.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Derivation of Reactions 19–24

Nitrate reduction to N2(aq) or NH3(aq) by H2(aq) is written as follows:
2NO3 + 2H+ + 5H2(aq) = N2(aq) + 6H2O,
2NO3 + 2H+ + 8H2(aq) = 2NH3(aq) + 6H2O.
Combining reactions A1 and A2 with reactions 1 and 2, reactions 19 to 24 are derived.

Appendix B. Operational Settings to Measure Mineralogical and Chemical Compositions of Solid Samples

  • The accelerating voltages for the secondary electron imaging and semi-quantitative chemical analyses by SEM/EDX were 5 kV and 15 kV, respectively.
  • EPMA parameters for the quantitative analyses were 15 kV of accelerating voltage, and 12 nA of specimen current.
  • XRD analyses were conducted at 30 kV and 15 mA using Cu–Kα radiation. XRD peaks were assigned with the aid of PDXL software using the Inorganic Crystal Structure Database.
  • XRF analyses were conducted at 50 kV and 60 mA using an Rh target.
  • Raman spectra of the solid reaction products were obtained with 532 nm excitation. The incident laser power density was typically 100–700 kW/cm2 for 100× (NA, 0.90) objective lenses. The final laser power at the sample surface was estimated to be 1–3 mW, and the spatial resolution of the Raman analysis was finer than 1 μm. Each point typically was scanned three times for 10–90 s from 100 to 2600 cm–1 at a spectral resolution of 2.3 cm–1 with a grating containing 600 grooves/mm. The Raman shift was calibrated using a silicon wafer prior to the analyses.

Appendix C

Table A1. Abbreviations and compositions used in this study.
Table A1. Abbreviations and compositions used in this study.
Act = Actinolite, Ca2(Mg,Fe)5Si8O22(OH)2
Ad = Andradite, Ca3FeIII2Si3O12
Aug = Augite, (Ca,Mg,Fe)2Si2O6
Chl = Chlorite
Chry = Chrysotile, Mg3Si2O5(OH)4
Clin = Clinochlore, (Mg5Al)(Si3Al)O10(OH)8
Cpx = Clinopyroxene
Diop = Diopside, CaMgSi2O6
For = Forsterite, Mg2SiO4
Faya = Fayalite, Fe2SiO4
Hal = Halite, NaCl
Mgt = Magnetite, Fe3O4
Ol = Olivine, (Mg,Fe)2SiO4
Sap = Saponite, Ca0.25(Mg2.5Fe0.5) (Si3.5Al0.5)O10(OH)2∙4H2O
Sme = Smectite
SiO2 (am) = Amorphous silica, SiO2
Srp = Serpentine, (Mg,Fe)3Si2O5(OH)4
Vrm = Vermiculite, (Mg,FeII,Al)3(Si,Al)4O10(OH)2∙4H2O
Wol = Wollastonite, CaSiO3
Figure A1. Time trends for saturation index, pHin-situ, and dissolved species concentrations of fluids from experiments in the komatiite–H2O–NO3 system at 250 °C (a) and 350 °C (b).
Figure A1. Time trends for saturation index, pHin-situ, and dissolved species concentrations of fluids from experiments in the komatiite–H2O–NO3 system at 250 °C (a) and 350 °C (b).
Minerals 11 00321 g0a1

References

  1. Orgel, L.E. Prebiotic Chemistry and the Origin of the RNA World. Crit. Rev. Biochem. Mol. Biol. 2004, 39, 99–123. [Google Scholar] [CrossRef] [Green Version]
  2. Martin, W.; Russell, M.J. On the origin of biochemistry at an alkaline hydrothermal vent. Philos. Trans. R. Soc. B Biol. Sci. 2007, 362, 1887–1926. [Google Scholar] [CrossRef] [PubMed]
  3. Navarro-Gonzãlez, R.; McKay, C.P.; Mvondo, D.N. A possible nitrogen crisis for Archaean life due to reduced nitrogen fixation by lightning. Nature 2001, 412, 61–64. [Google Scholar] [CrossRef]
  4. Raymond, J.; Siefert, J.L.; Staples, C.R.; Blankenship, R.E. The Natural History of Nitrogen Fixation. Mol. Biol. Evol. 2004, 21, 541–554. [Google Scholar] [CrossRef] [Green Version]
  5. Nishizawa, M.; Miyazaki, J.; Makabe, A.; Koba, K.; Takai, K. Physiological and isotopic characteristics of nitrogen fixation by hyperthermophilic methanogens: Key insights into nitrogen anabolism of the microbial communities in Archean hydrothermal systems. Geochim. Cosmochim. Acta 2014, 138, 117–135. [Google Scholar] [CrossRef]
  6. Kasting, J.F.; Eggler, D.H.; Raeburn, S.P. Mantle redox evolution and the oxidation state of the Archean atmosphere. J. Geol. 1993, 101, 245–257. [Google Scholar] [CrossRef] [PubMed]
  7. Rye, R.; Holland, H.D. Paleosols and the evolution of atmospheric oxygen: A critical review. Am. J. Sci. 1998, 298, 621–672. [Google Scholar] [CrossRef] [PubMed]
  8. Kitajima, K.; Maruyama, S.; Utsunomiya, S.; Liou, J.G. Seafloor hydrothermal alteration at an Archaean mid-ocean ridge. J. Metamorph. Geol. 2001, 19, 583–599. [Google Scholar] [CrossRef]
  9. Ohmoto, H.; Watanabe, Y.; Kumazawa, K. Evidence from massive siderite beds for a CO2-rich atmosphere before, 1.8 billion years ago. Nature 2004, 429, 395–399. [Google Scholar] [CrossRef] [PubMed]
  10. Nishizawa, M.; Sano, Y.; Ueno, Y.; Maruyama, S. Speciation and isotope ratios of nitrogen in fluid inclusions from seafloor hydrothermal deposits at similar to 3.5 Ga. Earth Planet. Sci. Lett. 2007, 254, 332–344. [Google Scholar] [CrossRef]
  11. Shibuya, T.; Kitajima, K.; Komiya, T.; Terabayashi, M.; Maruyama, S. Middle Archean ocean ridge hydrothermal metamorphism and alteration recorded in the Cleaverville area, Pilbara Craton, Western Australia. J. Metamorph. Geol. 2007, 25, 751–767. [Google Scholar] [CrossRef]
  12. Navarro-González, R.; Molina, M.J.; Molina, L.T. Nitrogen fixation by volcanic lightning in the early Earth. Geophys. Res. Lett. 1998, 25, 3123–3126. [Google Scholar] [CrossRef]
  13. Summers, D.P.; Khare, B. Nitrogen fixation on early Mars and other terrestrial planets: Experimental demonstration of abiotic fixation reactions to nitrite and nitrate. Astrobiology 2007, 7, 333–341. [Google Scholar] [CrossRef] [PubMed]
  14. Mojzsis, S.J.; Arrhenius, G.; Mckeegan, K.D.; Harrison, T.M.; Nutman, A.P.; Friend, C.R.L. Evidence for life on Earth before 3,800 million years ago. Nature 1996, 384, 55–59. [Google Scholar] [CrossRef] [PubMed]
  15. Rosing, M.T. 13C-depleted carbon microparticles in >3700-Ma sea floor sedimentary rocks from West Greenland. Science 1999, 283, 674–676. [Google Scholar] [CrossRef] [PubMed]
  16. Ueno, Y.; Yurimoto, H.; Yoshioka, H.; Komiya, T.; Maruyama, S. Ion microprobe analysis of graphite from ca. 3.8 Ga metasediments, Isua supracrustal belt, West Greenland: Relationship between metamorphism and carbon isotopic composition. Geochim. Cosmochim. Acta 2002, 66, 1257–1268. [Google Scholar] [CrossRef]
  17. Nishizawa, M.; Takahata, N.; Terada, K.; Komiya, T.; Ueno, Y.; Sano, Y. Rare-earth element, lead, carbon, and nitrogen geochemistry of apatite-bearing metasediments from the similar to 3.8 Ga Isua supracrustal belt, West Greenland. Int. Geol. Rev. 2005, 47, 952–970. [Google Scholar] [CrossRef]
  18. Ohtomo, Y.; Kakegawa, T.; Ishida, A.; Nagase, T.; Rosing, M.T. Evidence for biogenic graphite in early Archaean Isua metasedimentary rocks. Nat. Geosci. 2014, 7, 25–28. [Google Scholar] [CrossRef]
  19. Dodd, M.S.; Papineau, D.; Grenne, T.; Slack, J.F.; Rittner, M.; Pirajno, F.; O’Neil, J.; Little, C.T.S. Evidence for early life in Earth’s oldest hydrothermal vent precipitates. Nature 2017, 543, 60. [Google Scholar] [CrossRef] [PubMed]
  20. Tashiro, T.; Ishida, A.; Hori, M.; Igisu, M.; Koike, M.; Méjean, P.; Takahata, N.; Sano, Y.; Komiya, T. Early trace of life from 3.95 Ga sedimentary rocks in Labrador, Canada. Nature 2017, 549, 516. [Google Scholar] [CrossRef] [PubMed]
  21. Van Zuilen, M.A.; Leopand, A.; Arrhenius, G. Reassessing the evidence for the earliest traces of life. Nature 2002, 418, 627–630. [Google Scholar] [CrossRef]
  22. Papineau, D.; De Gregorio, B.T.; Cody, G.D.; Fries, M.D.; Mojzsis, S.J.; Steele, A.; Stroud, R.M.; Fogel, M.L. Ancient graphite in the Eoarchean quartz–pyroxene rocks from Akilia in southern West Greenland I: Petrographic and spectroscopic characterization. Geochim. Cosmochim. Acta 2010, 74, 5862–5883. [Google Scholar] [CrossRef]
  23. Sano, Y.; Terada, K.; Takahashi, Y.; Nutman, A.P. Origin of life from apatite dating? Nature 1999, 400, 127–128. [Google Scholar] [CrossRef] [PubMed]
  24. Buresh, R.J.; Moraghan, J.T. Chemical Reduction of Nitrate by Ferrous Iron. J. Environ. Qual. 1976, 5, 320–325. [Google Scholar] [CrossRef]
  25. Summers, D.P.; Chang, S. Prebiotic ammonia from reduction of nitrite by iron (II) on the early Earth. Nature 1993, 365, 630–633. [Google Scholar] [CrossRef]
  26. Hansen, H.C.B.; Koch, C.B.; Nancke-Krogh, H.; Borggaard, O.K.; Sørensen, J. Abiotic Nitrate Reduction to Ammonium: Key Role of Green Rust. Environ. Sci. Technol. 1996, 30, 2053–2056. [Google Scholar] [CrossRef]
  27. Hansen, H.C.B. Reduction of nitrate to ammonium by sulphate green rust; activation energy and reaction mechanism. Clay Miner. 1998, 33, 87–101. [Google Scholar] [CrossRef]
  28. Habicht, K.S. Calibration of Sulfate Levels in the Archean Ocean. Science 2002, 298, 2372–2374. [Google Scholar] [CrossRef] [Green Version]
  29. Crowe, S.A.; Paris, G.; Katsev, S.; Jones, C.; Kim, S.-T.; Zerkle, A.L.; Nomosatryo, S.; Fowle, D.A.; Adkins, J.F.; Sessions, A.L.; et al. Sulfate was a trace constituent of Archean seawater. Science 2014, 346, 735–739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Zhelezinskaia, I.; Kaufman, A.J.; Farquhar, J.; Cliff, J. Large sulfur isotope fractionations associated with Neoarchean microbial sulfate reduction. Science 2014, 346, 742–744. [Google Scholar] [CrossRef]
  31. Summers, D.P. Ammonia Formation By The Reduction Of Nitrite/Nitrate By Fes: Ammonia Formation Under Acidic Conditions. Orig. Life Evol. Biosph. 2005, 35, 299–312. [Google Scholar] [CrossRef] [PubMed]
  32. Brandes, J.A.; Hazen, R.M.; Yoder, H.S., Jr. Inorganic nitrogen reduction and stability under simulated hydrothermal conditions. Astrobiology 2008, 8, 1113–1126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Smirnov, A.; Hausner, D.; Laffers, R.; Strongin, D.R.; Schoonen, M.A. Abiotic ammonium formation in the presence of Ni-Fe metals and alloys and its implications for the Hadean nitrogen cycle. Geochem. Trans. 2008, 9, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Singireddy, S.; Gordon, A.D.; Smirnov, A.; Vance, M.A.; Schoonen, M.A.; Szilagyi, R.K.; Strongin, D.R. Reduction of nitrite and nitrate to ammonium on pyrite. Orig. Life Evol. Biosph. 2012, 42, 275–294. [Google Scholar] [CrossRef]
  35. Li, Y.; Yamaguchi, A.; Yamamoto, M.; Takai, K.; Nakamura, R. Molybdenum Sulfide: A Bioinspired Electrocatalyst for Dissimilatory Ammonia Synthesis with Geoelectrical Current. J. Phys. Chem. C 2017, 121, 2154–2164. [Google Scholar] [CrossRef]
  36. German, C.R.; Von Damm, K.L. Hydrothermal processes. In The Treatise on Geochemistry; Turekian, K.K., Holland, H.D., Eds.; Elsevier: New York, NY, USA, 2004; Chapter 6.07; pp. 181–222. [Google Scholar]
  37. 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. Geosystems 2012, 13, Q02T01. [Google Scholar] [CrossRef]
  38. Nishizawa, M.; Koba, K.; Makabe, A.; Yoshida, N.; Kaneko, M.; Hirao, S.; Ishibashi, J.-i.; Yamanaka, T.; Shibuya, T.; Kikuchi, T.; et al. Nitrification-driven forms of nitrogen metabolism in microbial mat communities thriving along an ammonium-enriched subsurface geothermal stream. Geochim. Cosmochim. Acta 2013, 113, 152–173. [Google Scholar] [CrossRef]
  39. Takahashi, E.; Scarfe, C.M. Melting of peridotite to 14 GPa and the genesis of komatiite. Nature 1985, 315, 566–568. [Google Scholar] [CrossRef]
  40. Arndt, N.T.; Lesher, C.M.; Barnes, S.J. Komatiite; Cambridge University Press: Cambridge, UK, 2008. [Google Scholar]
  41. Ohta, H.; Maruyama, S.; Takahashi, E.; Watanabe, Y.; Kato, Y. Field occurrence, geochemistry and petrogenesis of the Archean Mid-Oceanic Ridge Basalts (AMORBs) of the Cleaverville area, Pilbara Craton, Western Australia. Lithos 1996, 37, 199–221. [Google Scholar] [CrossRef]
  42. 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. 2004, 146, 333–367. [Google Scholar] [CrossRef]
  43. Takai, K.; Nakamura, K.; Suzuki, K.; Inagaki, F.; Nealson, K.; Kumagai, H. Ultramafics-Hydrothermalism-Hydrogenesis-HyperSLiME (UltraH3) linkage: A key insight into early microbial ecosystem in the Archean deep-sea hydrothermal systems. Paleontol. Res. 2006, 10, 269–282. [Google Scholar] [CrossRef]
  44. Russell, M.J.; Hall, A.J. The emergence of life from iron monosulphide bubbles at a submarine hydrothermal redox and pH front. J. Geol. Soc. 1997, 154, 377–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Martin, W.; Baross, J.; Kelley, D.; Russell, M.J. Hydrothermal vents and the origin of life. Nat. Rev. Microbiol. 2008, 6, 805–814. [Google Scholar] [CrossRef] [PubMed]
  46. Russell, M.J.; Barge, L.M.; Bhartia, R.; Bocanegra, D.; Bracher, P.J.; Branscomb, E.; Kidd, R.; McGlynn, S.; Meier, D.H.; Nitschke, W.; et al. The drive to life on wet and icy worlds. Astrobiology 2014, 14, 308–343. [Google Scholar] [CrossRef] [PubMed]
  47. Yoshizaki, M.; Shibuya, T.; Suzuki, K.; Shimizu, K.; Nakamura, K.; Takai, K.E.N.; Omori, S.; Maruyama, S. H2 generation by experimental hydrothermal alteration of komatiitic glass at 300°C and 500 bars: A preliminary result from on-going experiment. Geochem. J. 2009, 43, e17–e22. [Google Scholar] [CrossRef] [Green Version]
  48. Shibuya, T.; Yoshizaki, M.; Masaki, Y.; Suzuki, K.; Takai, K.; Russell, M.J. Reactions between basalt and CO2-rich seawater at 250 and 350°C, 500bars: Implications for the CO2 sequestration into the modern oceanic crust and the composition of hydrothermal vent fluid in the CO2-rich early ocean. Chem. Geol. 2013, 359, 1–9. [Google Scholar] [CrossRef]
  49. Shibuya, T.; Yoshizaki, M.; Sato, M.; Shimizu, K.; Nakamura, K.; Omori, S.; Suzuki, K.; Takai, K.; Tsunakawa, H.; Maruyama, S. Hydrogen-rich hydrothermal environments in the Hadean ocean inferred from serpentinization of komatiites at 300 °C and 500 bar. Prog. Earth Planet. Sci. 2015, 2, 1–11. [Google Scholar] [CrossRef] [Green Version]
  50. Ueda, H.; Shibuya, T.; Sawaki, Y.; Saitoh, M.; Takai, K.; Maruyama, S. Reactions between komatiite and CO2-rich seawater at 250 and 350 °C, 500 bars: Implications for hydrogen generation in the Hadean seafloor hydrothermal system. Prog. Earth Planet. Sci. 2016, 3, 1–14. [Google Scholar] [CrossRef] [Green Version]
  51. Nishizawa, M.; Sakai, S.; Konno, U.; Nakahara, N.; Takaki, Y.; Saito, Y.; Imachi, H.; Tasumi, E.; Makabe, A.; Koba, K.; et al. Nitrogen and Oxygen Isotope Effects of Ammonia Oxidation by Thermophilic Thaumarchaeota from a Geothermal Water Stream. Appl. Environ. Microbiol. 2016, 82, 4492–4504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Nishizawa, M.; Saitoh, M.; Matsui, Y.; Usui, Y.; Shibuya, T. Removal of organic contaminants from iron sulfides as a pretreatment for mineral-mediated chemical synthesis under prebiotic hydrothermal conditions. Geochem. J. 2017, 51, 495–505. [Google Scholar] [CrossRef]
  53. Seyfried, W.E., Jr.; Janecky, D.R.; Berndt, M.E. Rocking autoclaves for hydrothermal experiments II. The flexible reaction-cell system. In Hydrothermal Experimental Techniques; Ulmer, G.C., Barns, H.L., Eds.; John Wiley and Sons: Hoboken, NJ, USA, 1987; pp. 216–239. [Google Scholar]
  54. Wetzel, L.R.; Shock, E.L. Distinguishing ultramafic-from basalt-hosted submarine hydrothermal systems by comparing calculated vent fluid compositions. J. Geophys. Res. Solid Earth 2000, 105, 8319–8340. [Google Scholar] [CrossRef]
  55. Bethke, C.M. Geochemical and Biogeochemical Reaction Modeling, 2nd ed.; Cambridge University Press: Cambridge, UK, 2008. [Google Scholar]
  56. Johnson, J.W.; Oelkers, E.H.; Helgeson, H.C. SUPCRT92: A software package for calculating the standard molal thermodynamic properties of minerals, gases, aqueous species, and reactions from 1 to 5000 bar and 0 to 1000 °C. Comput. Geosci. 1992, 18, 899–947. [Google Scholar] [CrossRef]
  57. Shock, E.L.; Helgeson, H.C.; Sverjensky, D.A. Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures: Standard partial molal properties of inorganic neutral species. Geochim. Cosmochim. Acta 1989, 53, 2157–2183. [Google Scholar] [CrossRef]
  58. Shock, E.L.; Sassani, D.C.; Willis, M.; Sverjensky, D.A. Inorganic species in geologic fluids: Correlations among standard molal thermodynamic properties of aqueous ions and hydroxide complexes. Geochim. Cosmochim. Acta 1997, 61, 907–950. [Google Scholar] [CrossRef]
  59. Shock, E.L.; Koretsky, C.M. Metal-organic complexes in geochemical processes: Estimation of standard partial molal thermodynamic properties of aqueous complexes between metal cations and monovalent organic acid ligands at high pressures and temperatures. Geochim. Cosmochim. Acta 1995, 59, 1497–1532. [Google Scholar] [CrossRef]
  60. Sverjensky, D.A.; Shock, E.L.; Helgeson, H.C. Prediction of the thermodynamic properties of aqueous metal complexes to 1000°C and 5 kb. Geochim. Cosmochim. Acta 1997, 61, 1359–1412. [Google Scholar] [CrossRef]
  61. Helgeson, H.C. Thermodynamics of hydrothermal systems at elevated temperatures and pressures. Am. J. Sci. 1969, 267, 729–804. [Google Scholar] [CrossRef]
  62. Helgeson, H.C.; Kirkham, D.H. Theoretical prediction of the thermodynamic behavior of aqueous electrolytes at high pressures and temperatures; II, Debye-Huckel parameters for activity coefficients and relative partial molal properties. Am. J. Sci. 1974, 274, 1199–1261. [Google Scholar] [CrossRef] [Green Version]
  63. Buatier, M.D.; Früh-Green, G.L.; Karpoff, A.M. Mechanisms of Mg-phyllosilicate formation in a hydrothermal system at a sedimented ridge (Middle Valley, Juan de Fuca). Contrib. Mineral. Petrol. 1995, 122, 134–151. [Google Scholar] [CrossRef]
  64. Berger, A.; Gier, S.; Krois, P. Porosity-preserving chlorite cements in shallow-marine volcaniclastic sandstones: Evidence from Cretaceous sandstones of the Sawan gas field, Pakistan. AAPG Bull. 2009, 93, 595–615. [Google Scholar] [CrossRef] [Green Version]
  65. Haile, B.G.; Hellevang, H.; Aagaard, P.; Jahren, J. Experimental nucleation and growth of smectite and chlorite coatings on clean feldspar and quartz grain surfaces. Mar. Pet. Geol. 2015, 68, 664–674. [Google Scholar] [CrossRef] [Green Version]
  66. McCollom, T.M.; Klein, F.; Robbins, M.; Moskowitz, B.; Berquó, T.S.; Jöns, N.; Bach, W.; Templeton, A. Temperature trends for reaction rates, hydrogen generation, and partitioning of iron during experimental serpentinization of olivine. Geochim. Cosmochim. Acta 2016, 181, 175–200. [Google Scholar] [CrossRef] [Green Version]
  67. Seyfried, W.E.; Mottl, M.J. Hydrothermal alteration of basalt by seawater under seawater-dominated conditions. Geochim. Cosmochim. Acta 1982, 46, 985–1002. [Google Scholar] [CrossRef]
  68. Klein, F.; Bach, W.; Jöns, N.; Mccollom, T.; Moskowitz, B.; Berquó, T. Iron partitioning and hydrogen generation during serpentinization of abyssal peridotites from 15°N on the Mid-Atlantic Ridge. Geochim. Cosmochim. Acta 2009, 73, 6868–6893. [Google Scholar] [CrossRef]
  69. Stiegler, M.T.; Lowe, D.R.; Byerly, G.R. The Petrogenesis of Volcaniclastic Komatiites in the Barberton Greenstone Belt, South Africa: A Textural and Geochemical Study. J. Petrol. 2010, 51, 947–972. [Google Scholar] [CrossRef] [Green Version]
  70. Alt, J.C. Subseafloor processes in Mid-ocean ridge hydrothermal systems. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological and Geological Interactions; Humphiris, S.E., Zierenberg, R.A., Mullineaux, L.S., Thomson, R.E., Eds.; Geophysical Monograph; Wiley & Sons: New York, NY, USA, 1995; Volume 91, pp. 85–114. [Google Scholar]
  71. Butterfield, D.A.; Seyfried, W.E.; Lilley, M.D. Composition and evolution of hydrothermal fluids. In Energy and Mass Transfer in Marine Hydrothermal Systems; Halbach, P.E., Tunnicliffe, V., Hein, J.R., Eds.; Dahlem University Press: Berlin, Germany, 2003. [Google Scholar]
  72. Mccollom, T.M.; Bach, W. Thermodynamic constraints on hydrogen generation during serpentinization of ultramafic rocks. Geochim. Cosmochim. Acta 2009, 73, 856–875. [Google Scholar] [CrossRef]
  73. Amend, J.P.; McCollom, T.M. Energetics of Biomolecule Synthesis on Early Earth. In Chemical Evolution II: From the Origins of Life to Modern Society; American Chemical Society: Washington, DC, USA, 2009; Volume 1025, pp. 63–94. [Google Scholar]
  74. Takai, K.; Gamo, T.; Tsunogai, U.; Nakayama, N.; Hirayama, H.; Nealson, K.H.; Horikoshi, K. Geochemical and microbiological evidence for a hydrogen-based, hyperthermophilic subsurface lithoautotrophic microbial ecosystem (HyperSLiME) beneath an active deep-sea hydrothermal field. Extremophiles 2004, 8, 269–282. [Google Scholar] [CrossRef] [PubMed]
  75. Flores, G.E.; Campbell, J.H.; Kirshtein, J.D.; Meneghin, J.; Podar, M.; Steinberg, J.I.; Seewald, J.S.; Tivey, M.K.; Voytek, M.A.; Yang, Z.K.; et al. Microbial community structure of hydrothermal deposits from geochemically different vent fields along the Mid-Atlantic Ridge. Environ. Microbiol. 2011, 13, 2158–2171. [Google Scholar] [CrossRef] [PubMed]
  76. Barge, L.M.; Flores, E.; Baum, M.M.; VanderVelde, D.G.; Russell, M.J. Redox and pH gradients drive amino acid synthesis in iron oxyhydroxide mineral systems. Proc. Natl. Acad. Sci. USA 2019, 116, 4828–4833. [Google Scholar] [CrossRef] [Green Version]
  77. Kitadai, N.; Nakamura, R.; Yamamoto, M.; Takai, K.; Li, Y.; Yamaguchi, A.; Gilbert, A.; Ueno, Y.; Yoshida, N.; Oono, Y. Geoelectrochemical CO production: Implications for the autotrophic origin of life. Sci. Adv. 2018, 4, eaao7265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Kitadai, N.; Nakamura, R.; Yamamoto, M.; Takai, K.; Yoshida, N.; Oono, Y. Metals likely promoted protometabolism in early ocean alkaline hydrothermal systems. Sci. Adv. 2019, 5, eaav7848. [Google Scholar] [CrossRef] [Green Version]
  79. Shibuya, T.; Komiya, T.; Nakamura, K.; Takai, K.; Maruyama, S. Highly alkaline, high-temperature hydrothermal fluids in the early Archean ocean. Precambrian Res. 2010, 182, 230–238. [Google Scholar] [CrossRef]
  80. Yanagawa, H.; Kobayashi, K. Chapter 8 An experimental approach to chemical evolution in submarine hydrothermal systems. Orig. Life Evol. Biosph. 1992, 22, 147–159. [Google Scholar] [CrossRef]
  81. Marshall, W.L. Hydrothermal synthesis of amino acids. Geochim. Cosmochim. Acta 1994, 58, 2099–2106. [Google Scholar] [CrossRef]
  82. Rushdi, A.I.; Simoneit, B.R.T. Lipid formation by aqueous Fischer-Tropsch-type synthesis over a temperature range of 100 to 400?C. Orig. Life Evol. Biosph. 2001, 31, 103–118. [Google Scholar] [CrossRef] [PubMed]
  83. Novikov, Y.; Copley, S.D. Reactivity landscape of pyruvate under simulated hydrothermal vent conditions. Proc. Natl. Acad. Sci. USA 2013, 110, 13283–13288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Kadko, D.; Butterfield, D.A. The relationship of hydrothermal fluid composition and crustal residence time to maturity of vent fields on the Juan de Fuca Ridge. Geochim. Cosmochim. Acta 1998, 62, 1521–1533. [Google Scholar] [CrossRef]
  85. Kadko, D.; Gronvold, K.; Butterfield, D. Application of radium isotopes to determine crustal residence times of hydrothermal fluids from two sites on the Reykjanes Peninsula, Iceland. Geochim. Cosmochim. Acta 2007, 71, 6019–6029. [Google Scholar] [CrossRef]
  86. Laneuville, M.; Kameya, M.; Cleaves, H.J.I. Earth Without Life: A Systems Model of a Global Abiotic Nitrogen Cycle. Astrobiology 2018, 18, 897–914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Wong, M.L.; Charnay, B.D.; Gao, P.; Yung, Y.L.; Russell, M.J. Nitrogen Oxides in Early Earth’s Atmosphere as Electron Acceptors for Life’s Emergence. Astrobiology 2017, 17, 975–983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Ranjan, S.; Todd, Z.R.; Rimmer, P.B.; Sasselov, D.D.; Babbin, A.R. Nitrogen Oxide Concentrations in Natural Waters on Early Earth. Geochem. Geophys. Geosystems 2019, 20, 2021–2039. [Google Scholar] [CrossRef] [Green Version]
  89. Kawada, Y.; Seama, N.; Okino, K. Review of Oceanic Crustal Structure and Seafloor Hydrothermal Circulation. J. Geogr. 2009, 118, 1037–1063. [Google Scholar] [CrossRef] [Green Version]
  90. Abbott, D.; Menke, W. Length of the global plate boundary at 2.4 Ga. Geology 1990, 18, 58–61. [Google Scholar] [CrossRef]
  91. Korenaga, J. Plate tectonics, flood basalts and the evolution of Earth’s oceans. Terra Nova 2008, 20, 419–439. [Google Scholar] [CrossRef]
  92. Peacock, S.A. Fluid Processes in Subduction Zones. Science 1990, 248, 329–337. [Google Scholar] [CrossRef] [PubMed]
  93. Clift, P.D.; Vannucchi, P.; Morgan, J.P. Crustal redistribution, crust–mantle recycling and Phanerozoic evolution of the continental crust. Earth Sci. Rev. 2009, 97, 80–104. [Google Scholar] [CrossRef] [Green Version]
  94. Elderfield, H.; Schults, A. Mid-ocean ridge hydrothermal fluxes and the chemical composition of the ocean. Ann. Rev. Earth Planet. Sci 1996, 24, 191–224. [Google Scholar] [CrossRef]
  95. Chan, L.-H.; Alt, J.C.; Teagle, D.A.H. Lithium and lithium isotope profiles through the upper oceanic crust: A study of seawater–basalt exchange at ODP Sites 504B and 896A. Earth Planet. Sci. Lett. 2002, 201, 187–201. [Google Scholar] [CrossRef]
  96. Nielsen, S.G.; Rehkämper, M.; Teagle, D.A.H.; Butterfield, D.A.; Alt, J.C.; Halliday, A.N. Hydrothermal fluid fluxes calculated from the isotopic mass balance of thallium in the ocean crust. Earth Planet. Sci. Lett. 2006, 251, 120–133. [Google Scholar] [CrossRef]
  97. Reay, D.S.; Nedwell, D.B.; Priddle, J.; Ellis-Evans, J.C. Temperature dependence of inorganic nitrogen uptake: Reduced affinity for nitrate at suboptimal temperatures in both algae and bacteria. Appl. Environ. Microbiol. 1999, 65, 2577–2584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Ueno, Y.; Yoshioka, H.; Maruyama, S.; Isozaki, Y. Carbon isotopes and petrography of kerogens in ~3.5-Ga hydrothermal silica dikes in the North Pole area, Western Australia. Geochim. Cosmochim. Acta 2004, 68, 573–589. [Google Scholar] [CrossRef]
  99. Weiss, M.C.; Sousa, F.L.; Mrnjavac, N.; Neukirchen, S.; Roettger, M.; Nelson-Sathi, S.; Martin, W.F. The physiology and habitat of the last universal common ancestor. Nat. Microbiol. 2016, 1, 16116. [Google Scholar] [CrossRef] [PubMed]
  100. Boyd, E.S.; Hamilton, T.L.; Peters, J.W. An alternative path for the evolution of biological nitrogen fixation. Front. Microbiol. 2011, 2, 205. [Google Scholar] [CrossRef] [Green Version]
  101. Stueken, E.E.; Buick, R.; Guy, B.M.; Koehler, M.C. Isotopic evidence for biological nitrogen fixation by molybdenum-nitrogenase from 3.2 Gyr. Nature 2015, 520, 666–669. [Google Scholar] [CrossRef] [PubMed]
  102. Sano, Y.; Pillinger, C.T. Nitrogen isotopes and N2/Ar ratios in chests: An attempt to measure time evolution of atmospheric δ15N value. Geochem. J. 1990, 24, 315–325. [Google Scholar] [CrossRef] [Green Version]
  103. Marty, B.; Zimmermann, L.; Pujol, M.; Burgess, R.; Philippot, P. Nitrogen isotopic composition and density of the Archean atmosphere. Science 2013, 342, 101–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Avice, G.; Marty, B.; Burgess, R.; Hofmann, A.; Philippot, P.; Zahnle, K.; Zakharov, D. Evolution of atmospheric xenon and other noble gases inferred from Archean to Paleoproterozoic rocks. Geochim. Cosmochim. Acta 2018, 232, 82–100. [Google Scholar] [CrossRef]
  105. Goldblatt, C.; Claire, M.W.; Lenton, T.M.; Matthews, A.J.; Watson, A.J.; Zahnle, K.J. Nitrogen-enhanced greenhouse warming on early Earth. Nat. Geosci. 2009, 2, 891–896. [Google Scholar] [CrossRef]
  106. Brandes, J.A.; Devol, A.H. A global marine-fixed nitrogen isotopic budget: Implications for Holocene nitrogen cycling. Glob. Biogeochem. Cycles 2002, 16, 1120. [Google Scholar] [CrossRef]
  107. Hilton, D.R.; Fischer, T.P.; Marty, B. Noble gases and volatile recycling at subduction zones. In Noble Gases in Geochemistry and Cosmochemistry, Reviews in Mineralogy and Geochemistry; Porcelli, D., Ballentine, C.J., Wieler, R., Eds.; Mineralogical Society of America: Washington, DC, USA, 2002; Volume 47, pp. 319–370. [Google Scholar]
  108. Sano, Y.; Takahata, N.; Nishio, Y.; Fischer, T.P.; Williams, S.N. Volcanic flux of nitrogen from the Earth. Chem. Geol. 2001, 171, 263–271. [Google Scholar] [CrossRef]
  109. Pinti, D.L.; Hashizume, K.; Matsuda, J.i. Nitrogen and argon signatures in 3.8 to 2.8 Ga metasediments: Clues on the chemical state of the archean ocean and the deep biosphere. Geochim. Cosmochim. Acta 2001, 65, 2301–2315. [Google Scholar] [CrossRef]
Figure 1. Measured concentrations of dissolved species in an aqueous fluid coexisting with komatiite and its alteration products as functions of time at 250 °C (a,c) and 350 °C (b,d), 500 bars. (a,b) Fluid compositions in the komatiite–H2O–NO3 system at 250 °C and 350 °C with the initial water–rock mass ratios of 7 and 5, respectively. (c,d) Fluid compositions in the komatiite–H2O–CO2 system at 250 °C and 350 °C with the initial water–rock mass ratios of 5 and 5, respectively (Ueda et al., 2016). White triangles and light-blue circles with black arrows imply that iron and nitrate concentrations in plot are the maximum estimates (0.02 mmol/kg for iron in N250, 0.003 mmol/kg for nitrate in N350).
Figure 1. Measured concentrations of dissolved species in an aqueous fluid coexisting with komatiite and its alteration products as functions of time at 250 °C (a,c) and 350 °C (b,d), 500 bars. (a,b) Fluid compositions in the komatiite–H2O–NO3 system at 250 °C and 350 °C with the initial water–rock mass ratios of 7 and 5, respectively. (c,d) Fluid compositions in the komatiite–H2O–CO2 system at 250 °C and 350 °C with the initial water–rock mass ratios of 5 and 5, respectively (Ueda et al., 2016). White triangles and light-blue circles with black arrows imply that iron and nitrate concentrations in plot are the maximum estimates (0.02 mmol/kg for iron in N250, 0.003 mmol/kg for nitrate in N350).
Minerals 11 00321 g001
Figure 2. Mole fractions of nitrate, nitrite, nitrous oxide, and ammonia in aqueous fluids without komatiite at (a) 250 °C and (b) 350 °C. Measured concentrations of nitrogen species at a given time are normalized by the initial concentration of nitrate (10 mmol/kg) to estimate the mole fractions. For each plot, Y axis on the left side shows the scale of the mole fractions of nitrate and nitrite, whereas that on the right side shows the scale of the mole fractions of nitrous oxide and ammonia.
Figure 2. Mole fractions of nitrate, nitrite, nitrous oxide, and ammonia in aqueous fluids without komatiite at (a) 250 °C and (b) 350 °C. Measured concentrations of nitrogen species at a given time are normalized by the initial concentration of nitrate (10 mmol/kg) to estimate the mole fractions. For each plot, Y axis on the left side shows the scale of the mole fractions of nitrate and nitrite, whereas that on the right side shows the scale of the mole fractions of nitrous oxide and ammonia.
Minerals 11 00321 g002
Figure 3. Time-dependent changes in dissolved H2 concentrations in aqueous fluids in the systems of komatiite–H2O–NO3 (this study), komatiite–H2O–CO2 [50], and komatiite–H2O [49].
Figure 3. Time-dependent changes in dissolved H2 concentrations in aqueous fluids in the systems of komatiite–H2O–NO3 (this study), komatiite–H2O–CO2 [50], and komatiite–H2O [49].
Minerals 11 00321 g003
Figure 4. Powder XRD patterns (CuKα) of the alteration products after experiment N250 (250 °C, 4877 h) and N350 (350 °C, 2520 h), denoted by red. For comparison, XRD patterns (CuKα) of the synthetic komatiite used in the experiments are presented in green. The mineral identified from the XRD patterns of the synthetic komatiite is ferroan forsterite with a mole (Mg/Fe) ratio of ~9. Abbreviations are the same as in Table A1.
Figure 4. Powder XRD patterns (CuKα) of the alteration products after experiment N250 (250 °C, 4877 h) and N350 (350 °C, 2520 h), denoted by red. For comparison, XRD patterns (CuKα) of the synthetic komatiite used in the experiments are presented in green. The mineral identified from the XRD patterns of the synthetic komatiite is ferroan forsterite with a mole (Mg/Fe) ratio of ~9. Abbreviations are the same as in Table A1.
Minerals 11 00321 g004
Figure 5. Optical microscopy images and Raman spectra of grains in unsuspended fractions separated from the bulk products at 250 °C and 350 °C. (a,c,e) Photomicrographs of an unsuspended fraction in experiments N250 (250 °C, 4877 h) showing yellowish green grains (10–50 μm in size). Small spherical grains (ca. 1 μm in size) are often found on the surface of yellowish green grains. (b,d,f,g) Photomicrographs of an unsuspended fraction in experiment N350 (350 °C, 2520 h) showing two types of grains: rounded greenish grains (20–50 μm in size) and random-shaped opaque grains (20–50 μm in size). Numbers in Figure 5e–g; correspond to those shown in Figure 5h. (h) Representative Raman spectra. Raman spectra of the yellowish green grains of N250 and the greenish grains of N350 (numbers 1 and 3) show peaks attributable to smectite, chlorite, actinolite, and serpentine that have nearly the same wavenumber. Raman spectra of the small spherical grains of N250 (number 2) show peaks attributable to andradite. Raman spectra of the random-shaped opaque grains of N350 (number 4) show the peaks attributable to clinopyroxene. A large peak between 650 cm–1 and 700 cm–1 is common in inosilicate and phyllosilicate candidates. Abbreviations in Figure 5h are the same as in Table A1.
Figure 5. Optical microscopy images and Raman spectra of grains in unsuspended fractions separated from the bulk products at 250 °C and 350 °C. (a,c,e) Photomicrographs of an unsuspended fraction in experiments N250 (250 °C, 4877 h) showing yellowish green grains (10–50 μm in size). Small spherical grains (ca. 1 μm in size) are often found on the surface of yellowish green grains. (b,d,f,g) Photomicrographs of an unsuspended fraction in experiment N350 (350 °C, 2520 h) showing two types of grains: rounded greenish grains (20–50 μm in size) and random-shaped opaque grains (20–50 μm in size). Numbers in Figure 5e–g; correspond to those shown in Figure 5h. (h) Representative Raman spectra. Raman spectra of the yellowish green grains of N250 and the greenish grains of N350 (numbers 1 and 3) show peaks attributable to smectite, chlorite, actinolite, and serpentine that have nearly the same wavenumber. Raman spectra of the small spherical grains of N250 (number 2) show peaks attributable to andradite. Raman spectra of the random-shaped opaque grains of N350 (number 4) show the peaks attributable to clinopyroxene. A large peak between 650 cm–1 and 700 cm–1 is common in inosilicate and phyllosilicate candidates. Abbreviations in Figure 5h are the same as in Table A1.
Minerals 11 00321 g005
Figure 6. Representative SEM images of alteration products after experiment N250 (250 °C, 4877 h). (a,b) Polished sample mounted on epoxy resin showing 10–50 μm-size clusters comprising phyllosilicates (smectite, serpentine, and chlorite: Sme-Srp-Chl) in the core with overgrowth rims of phyllosilicates and bright spherical andradite. The dark portion of these panels is epoxy resin. (c) Aggregate of crushed samples showing various morphologies of individual grains. (dg) SEM images of the unsuspended fraction separated from bulk products by sonication with water and ethanol (d) Clusters of phyllosilicates that are partially covered by phyllosilicates with frost column textures and/or andradite. (e) Cluster of phyllosilicates. The core of the cluster comprises plates/flakes that develops an edge-to-face pattern and/or a rosette-like pattern, whereas the rim shows frost column texture. (f) Enlarged view of a dashed black frame in (d) showing botryoidally aggregated andradite spheres overgrown on the surface of phyllosilicate cluster. (g) Spherical aggregate comprising clinopyroxene needles, andradite spheres, and phyllosilicates. (h) Cross sectional view of phyllosilicate cluster covered by phyllosilicate layers. Spherical andradite crystals occur with needle-like calcium silicate (CS) at the surface of the upper most layer of phyllosilicates. Abbreviations are the same as in Table A1.
Figure 6. Representative SEM images of alteration products after experiment N250 (250 °C, 4877 h). (a,b) Polished sample mounted on epoxy resin showing 10–50 μm-size clusters comprising phyllosilicates (smectite, serpentine, and chlorite: Sme-Srp-Chl) in the core with overgrowth rims of phyllosilicates and bright spherical andradite. The dark portion of these panels is epoxy resin. (c) Aggregate of crushed samples showing various morphologies of individual grains. (dg) SEM images of the unsuspended fraction separated from bulk products by sonication with water and ethanol (d) Clusters of phyllosilicates that are partially covered by phyllosilicates with frost column textures and/or andradite. (e) Cluster of phyllosilicates. The core of the cluster comprises plates/flakes that develops an edge-to-face pattern and/or a rosette-like pattern, whereas the rim shows frost column texture. (f) Enlarged view of a dashed black frame in (d) showing botryoidally aggregated andradite spheres overgrown on the surface of phyllosilicate cluster. (g) Spherical aggregate comprising clinopyroxene needles, andradite spheres, and phyllosilicates. (h) Cross sectional view of phyllosilicate cluster covered by phyllosilicate layers. Spherical andradite crystals occur with needle-like calcium silicate (CS) at the surface of the upper most layer of phyllosilicates. Abbreviations are the same as in Table A1.
Minerals 11 00321 g006
Figure 7. Representative SEM images of alteration products after experiment N350 (350 °C, 2520 h). (a) A polished sample mounted on epoxy resin showing clusters of phyllosilicates (smectite, serpentine, and chlorite) in the core with overgrowth rims of phyllosilicates and needle-like crystals of clinopyroxene. The dark portion of the panel is epoxy resin. (bd) SEM images of crushed samples. (b) Cluster of phyllosilicates. The core of the cluster comprises plates/flakes that develop an edge-to-face pattern and/or a rosette-like pattern, whereas the rim shows honeycomb texture with 2–3 μm pores. Needle-shaped clinopyroxene occurs with overgrowth rims of phyllosilicates. (c,d) Cluster of phyllosilicates with an unidentified polygonal mineral and a fibrous mineral that is presumed to be chrysotile. (e,f) Unsuspended fraction showing rounded clusters comprising phyllosilicates (e) and irregular-shaped clusters comprising clinopyroxene and phyllosilicates (f). (g) Enlarged view of a phyllosilicate cluster. Plates/flakes develop an edge-to-face pattern. (h) Enlarged view of needle-shaped clinopyroxene. Abbreviations are the same as in Table A1.
Figure 7. Representative SEM images of alteration products after experiment N350 (350 °C, 2520 h). (a) A polished sample mounted on epoxy resin showing clusters of phyllosilicates (smectite, serpentine, and chlorite) in the core with overgrowth rims of phyllosilicates and needle-like crystals of clinopyroxene. The dark portion of the panel is epoxy resin. (bd) SEM images of crushed samples. (b) Cluster of phyllosilicates. The core of the cluster comprises plates/flakes that develop an edge-to-face pattern and/or a rosette-like pattern, whereas the rim shows honeycomb texture with 2–3 μm pores. Needle-shaped clinopyroxene occurs with overgrowth rims of phyllosilicates. (c,d) Cluster of phyllosilicates with an unidentified polygonal mineral and a fibrous mineral that is presumed to be chrysotile. (e,f) Unsuspended fraction showing rounded clusters comprising phyllosilicates (e) and irregular-shaped clusters comprising clinopyroxene and phyllosilicates (f). (g) Enlarged view of a phyllosilicate cluster. Plates/flakes develop an edge-to-face pattern. (h) Enlarged view of needle-shaped clinopyroxene. Abbreviations are the same as in Table A1.
Minerals 11 00321 g007
Figure 8. Model of abiotic nitrate reduction processes. Nitrite and nitrous oxide were detected as intermediate products during abiotic nitrate reduction in the komatiite–H2O–NO3 and H2O–NO3 systems. By contrast, ammonia was detected as an end product of abiotic nitrate reduction in the komatiite–H2O–NO3 system (“chemoammonification”), but not in the H2O–NO3 system. N2 is thought to be another end product of abiotic nitrate reduction via nitrous oxide (“chemodenitrification”). Nitrogen species detected by experiments are shown in square. Experiment IDs in which a given nitrogen species was detected are shown. The fact that nitrite remained undetected in N250C and N350 suggests that the rate of nitrite reduction is much faster than that of nitrate reduction to nitrite. n.m.—not measured.
Figure 8. Model of abiotic nitrate reduction processes. Nitrite and nitrous oxide were detected as intermediate products during abiotic nitrate reduction in the komatiite–H2O–NO3 and H2O–NO3 systems. By contrast, ammonia was detected as an end product of abiotic nitrate reduction in the komatiite–H2O–NO3 system (“chemoammonification”), but not in the H2O–NO3 system. N2 is thought to be another end product of abiotic nitrate reduction via nitrous oxide (“chemodenitrification”). Nitrogen species detected by experiments are shown in square. Experiment IDs in which a given nitrogen species was detected are shown. The fact that nitrite remained undetected in N250C and N350 suggests that the rate of nitrite reduction is much faster than that of nitrate reduction to nitrite. n.m.—not measured.
Minerals 11 00321 g008
Figure 9. Diagrams of pHin-situ and aH2(aq) (or aO2(aq)) with stability fields of nitrogen species in thermodynamic equilibrium at (a) 250 °C and (b) 350 °C, 500 bars. Bold red lines indicate the pHin-situ and aH2(aq) values at which the activity ratios of two nitrogen species follow the relations: aNH4+/aNH3(aq) = 1, aN2(aq)/aNH3(aq) = 0.5, aN2(aq)/aNH4+ = 0.5, or aN2(aq)/aNO3 = 0.5. For the calculation, aNH3(aq) values were used in the final sampled fluids at 250 °C and 350 °C (aNH3(aq) = 10−3.228 and 10−3.98 at 250 °C and 350 °C, respectively). Thin, kinked red lines show the pHin-situ and aH2(aq) values at which the activity ratio of aN2(aq)/(aNH3(aq) + aNH4+) is equal to 0.05, 5, and 50. Green and yellow circles indicate the fluid compositions in the komatiite–H2O–NO3 system at 250 °C and 350 °C, respectively. The circles with black arrows imply that aH2(aq) in the plot is the maximum estimate. Numbers associated with green and yellow circles show sub-name of sampled fluids in experiments N250 and N350, respectively (Table 1). Purple and black boxes denote the fluid compositions in the H2O–NO3 system at 250 °C and 350 °C, respectively (experiments N250C and N350C). Light blue areas in pHin-situaH2(aq) spaces are the stability fields of water at 250 °C and 350 °C, respectively.
Figure 9. Diagrams of pHin-situ and aH2(aq) (or aO2(aq)) with stability fields of nitrogen species in thermodynamic equilibrium at (a) 250 °C and (b) 350 °C, 500 bars. Bold red lines indicate the pHin-situ and aH2(aq) values at which the activity ratios of two nitrogen species follow the relations: aNH4+/aNH3(aq) = 1, aN2(aq)/aNH3(aq) = 0.5, aN2(aq)/aNH4+ = 0.5, or aN2(aq)/aNO3 = 0.5. For the calculation, aNH3(aq) values were used in the final sampled fluids at 250 °C and 350 °C (aNH3(aq) = 10−3.228 and 10−3.98 at 250 °C and 350 °C, respectively). Thin, kinked red lines show the pHin-situ and aH2(aq) values at which the activity ratio of aN2(aq)/(aNH3(aq) + aNH4+) is equal to 0.05, 5, and 50. Green and yellow circles indicate the fluid compositions in the komatiite–H2O–NO3 system at 250 °C and 350 °C, respectively. The circles with black arrows imply that aH2(aq) in the plot is the maximum estimate. Numbers associated with green and yellow circles show sub-name of sampled fluids in experiments N250 and N350, respectively (Table 1). Purple and black boxes denote the fluid compositions in the H2O–NO3 system at 250 °C and 350 °C, respectively (experiments N250C and N350C). Light blue areas in pHin-situaH2(aq) spaces are the stability fields of water at 250 °C and 350 °C, respectively.
Minerals 11 00321 g009
Table 1. Concentration of selected species measured in fluids from experiments in the komatiite–H2O–NO3 system.
Table 1. Concentration of selected species measured in fluids from experiments in the komatiite–H2O–NO3 system.
Time [h]T [°C]pH25 apHin-situ bH2 [m]Cl [m]Na [m]K [m]Mg [m]Ca [m]Si [m]Fe [m]Mn [m]NH4+ [m]NO2 [m]NO3 [m]N2O [m]Fluid c [g]
N250 (soln. = 72.4 g, rock = 9.64 g, P = 500 bars)
0 209 0.030 99010140.280.3390.30.10.02 dn.d.n.d.n.d.10.936n.d.72.4
10.02359.67.70.030 9869910.320.0092.72.50.02 dn.d.0.023n.d.10.748n.d.69.4
242469.47.80.030 9949860.390.0043.94.70.02 dn.d.0.0300.00910.5340.00265.2
382508.47.60.030 9869800.390.0024.25.70.02 dn.d.0.0370.01110.7160.00560.7
4182509.17.80.030 9859910.400.0024.66.00.02 dn.d.0.0560.01110.2860.01156.3
5442508.87.70.030 10079840.530.0026.26.20.02 dn.d.0.1010.00910.1320.00952.2
62642508.96.20.030 10169551.830.01836.01.10.02 dn.d.1.140n.d.0.0150.002 48.9
710542508.85.90.030 10519571.870.02330.80.80.02 dn.d.1.168n.d.0.0050.00245.6
819442508.65.60.58510459611.790.02030.70.70.02 dn.d.1.192n.d.0.007n.d. 42.2
926162508.85.81.410439611.810.02030.40.70.02 dn.d.1.184n.d.0.006n.m.38.9
1048712508.65.62.310459601.740.02029.60.70.02 dn.d.1.174n.d.0.010n.m.35.4
N350 (soln. = 63.93 g, rock = 12.2 g, P = 500 bars)
0 25 8.5 0.031 989 992 0.850.5881.300.010.0010.002n.d.1.063n.d. 63.9
10350 7.27.10.019 1004 1006 1.010.0373.818.10.030.0010.085n.d.0.950n.d.60.9
217350 6.35.80.006 999 973 1.070.01110.420.50.010.0040.132n.d.0.1060.00755.3
371350 6.24.90.668 1004 929 1.730.19737.82.90.180.0870.180n.d.0.005n.d. 50.3
4187350 6.14.91.941 1030 955 1.770.26634.22.80.330.1480.182n.d.0.003 dn.d. 44.8
5521350 6.14.93.400 1020 956 2.040.3923.83.00.270.1040.176n.d.0.003n.d. 39.9
62515350 5.75.01.660 1026 967 2.100.2423.12.80.270.0860.171n.d.n.m.n.d.35.1
Concentrations are in mmol/kg [m]. n.d.—not detected. n.m.—not measured. a Measured pH at 25 °C. b Calculated pHin-situ at experimental temperature and 500 bars. c Estimated amount of fluid in the reaction cell prior to sampling. d Upper limit.
Table 2. Concentration of selected species measured in fluids from experiments in the H2O–NO3 system.
Table 2. Concentration of selected species measured in fluids from experiments in the H2O–NO3 system.
Time [h]T [°C]pH25pHin-situCl [m]Na [m]K [m]NH4+ [m]NO2 [m]NO3 [m]N2O [m]
N250C (P = 500 bars)
0 22 97095810.0n.d.0.00210.132n.m.
1222508.75.6101197010.4m.f.0.00210.3290.001
22132508.85.799995010.0m.f.0.00210.2670.007
3381250106.79879719.9m.f.0.00210.2340.006
48602509.46.299896210.1n.d.0.00210.3610.005
N350C (P = 500 bars)
0 226.3 98596610.0n.d.0.00510.2050.000
103399.96.6 a983n.m.n.m.n.d.0.0029.7330.009
243469.96.6 a983n.m.n.m.n.d.0.1229.9190.013
38349106.7 a980n.m.n.m.n.d.0.1539.9230.012
411350106.7 a982n.m.n.m.n.d.0.1969.8190.010
5213509.86.5 a986n.m.n.m.n.d.0.2679.8540.010
6763509.46.1 a996n.m.n.m.n.d.0.4049.4340.010
72643509.56.2103097410.0n.d.1.1039.1370.005
Concentrations are in mmol/kg [m]. n.d.—not detected. n.m.—not measured. m.f.—measurement failed. a Calculated pHin-situ at experimental temperature and 500 bars assuming that Na and K concentrations in the sampled fluid are the same as those in the initial solution (Na = 966 mmol/kg, K = 10.0 mmol/kg).
Table 3. Steady-state composition of aqueous fluids in the komatiite–H2O–NO3 and komatiite–H2O–CO2 systems.
Table 3. Steady-state composition of aqueous fluids in the komatiite–H2O–NO3 and komatiite–H2O–CO2 systems.
SystemKomatiite-H2O-NO3Komatiite-H2O-CO2
Experiment IDN250N350U250 bU350 b
T [°C]250350250350
P [bar]500500500500
Reaction time [hours]4871251527601530
(water/rock)final a3.72.933
pH25°C8.76.04.53.9
pHin-situ5.85.04.85.7
H2 [m]2.31.70.022.2
ΣCO2 [m]n.m.n.m.33171
Cl [m]104510261175953
Na [m]960967949884
K [m]1.742.100.950.81
Mg [m]≤0.020.2436.41
Ca [m]302348.837.3
Si [m]0.742.813.54.1
Fe [m]≤0.020.270.971.55
Mn [m]≤0.0010.0860.250.6
NH4+ [m]1.1740.171n.m.n.m.
referenceThis studyThis study[50].[50]
Concentrations are in mmol/kg [m]. n.m.—not measured. a Water–rock mass ratio. b Initial aqueous solutions were prepared with the following chemical composition: 1000 mmol/kg Cl, 1000 mmol/kg Na, and 400 mmol/kg ΣCO2. The pH values of the initial solutions were ~4.9 at room temperature.
Table 4. Composition of synthetic komatiite and solid reaction products from the experiments in the komatiite–H2O–NO3 system at 250 °C and 350 °C (values in wt.%).
Table 4. Composition of synthetic komatiite and solid reaction products from the experiments in the komatiite–H2O–NO3 system at 250 °C and 350 °C (values in wt.%).
Al-Depleted Synthetic KomatiiteN250N350Reference Minerals (Ideal Composition)
WROlGlassClst 1Clst 2Clst 1Clst 2SapSrpClin
SiO246.7041.1151.7041.0639.7538.8440.6444.1141.0132.42
TiO20.380.030.500.130.410.080.94
Al2O33.70.045.491.174.035.683.855.35 18.24
Cr2O3n.m. (0.35)0.160.320.560.460.250.21
FeO11.437.0210.9411.1310.5810.218.717.5412.26
MnO0.180.130.220.300.300.140.09
MgO29.8949.8018.8325.5923.1827.6821.3921.1434.4036.25
CaO7.260.239.930.712.830.283.142.94
Na2O0.060.010.25≤0.65≤0.84≤0.47≤0.94
K2On.d. (0.03)0.010.00n.d.0.004n.d.0.017
NiOn.m. (0.19)0.190.070.050.070.080.12
P2O50.04n.d.0.02n.d.0.006n.d.n.d.
Total 99.6398.7098.2381.3582.4783.7080.0681.0787.6787.00
MethodXRFEPMAEPMAEPMAEPMAEPMAEPMAcalc.calc.calc.
Note Ca-poorCa-richCa-poorCa-rich
n.m.—not measured. n.d.—not detected. cal.—calculation. Values in parentheses are target concentrations. Srp = Mg5FeSi4O10(OH)8. Abbreviations are the same as in Table A1 except for WR (=whole rock) and Clst (=Cluster).
Table 5. Representative elemental compositions measured using SEM/EDS (atom number %).
Table 5. Representative elemental compositions measured using SEM/EDS (atom number %).
N250
MorphologyClusterFrost Column aPolygonalNeedleNeedleColumnar
InterpretationSme-Srp-ChlSme-Srp-ChlAndraditeAugiteWollastoniteCalcium Silicate
(Mg + Fe)/Si1.32 1.04 0.72 0.81 0.23 0.42
Fe/Fe + Mg0.12 0.12 0.79 0.19 0.13 0.67
Ca/Si0.07 0.04 1.06 0.40 1.24 3.10
Al/Si0.14 0.10 0.07 0.05 0.02 0.02
N350
MorphologyClusterFlake bNeedleNeedle
InterpretationSme-Srp-ChlSme-Srp-ChlAugiteAugite
(Mg + Fe)/Si1.27 1.14 0.58 0.74
Fe/Fe + Mg0.33 0.11 0.28 0.20
Ca/Si0.04 0.04 0.51 0.34
Al/Si0.12 0.13 0.04 0.04
a Secondary mineral partly covering the surface of the cluster. b Mineral flakes covering the surface of the cluster in honeycomb pattern. Abbreviations are the same as in Table A1.
Table 6. Saturation indexes of selected solid materials co-existing with fluids from experiments in the komatiite–H2O–NO3 system.
Table 6. Saturation indexes of selected solid materials co-existing with fluids from experiments in the komatiite–H2O–NO3 system.
Time [h]T [°C]SiO2 (am)ForFaya aChryMgt aAd aDiopWol
N250
10235−1.12.62.76.56.210.24.50.8
24244−0.92.72.56.65.712.05.31.6
38250−0.71.72.85.26.111.64.61.4
418250−0.82.32.36.05.412.55.41.8
544250−0.82.12.55.75.712.75.31.8
6264250−1.3−1.61.8−0.15.64.10.3−1.0
71054250−1.5−2.80.5−2.03.80.2−1.3−1.9
81944250−1.5<−3−0.5<−31.0<−3−2.5−2.5
92616250−1.5<−30.1−2.91.6−2.6−1.8−2.2
104871250−1.5<−3−0.5<−30.5<−3−2.5−2.5
N350
10350−0.34.82.98.46.711.05.71.2
217350−0.20.2−0.41.62.42.31.4−0.8
371350−1.1−1.6−2.4−1.5−1.4<−3−2.0−2.9
4187350−1.1−1.4−1.9−1.3−1.2<−3−2.0<−3
5521350−1.1−1.0−2.1−0.7−1.7<−3−2.0<−3
62515350−1.1−1.5−2.1−1.4−1.4<−3−2.3<−3
Saturation index calculated from the concentration of dissolved species listed in Table 1. a Saturation index during each time interval in experiment N250 lists the upper limit because Table 1 lists the upper limit of the dissolved iron concentration during each time interval in N250. Abbreviations are the same as in Table A1.
Table 7. List of water-rock reactions.
Table 7. List of water-rock reactions.
Reaction#
1. Silicate formation from solid components MgO, Al2O3, and FeO
0.25Ca2+ + 2.5MgOrock + 0.5FeOrock + 0.25Al2O3rock + 3.5SiO2(aq) + 5.25H2O
= Ca0.25(Mg2.5Fe0.5)(Si3.5Al0.5)O10(OH)2∙4H2O [Saponite] + 0.5H+
(1)
5MgOrock + FeOrock + 4SiO2(aq) + 14H2O = Mg5FeSi4O10(OH)8 [Serpentine](2)
5MgOrock + 0.5Al2O3rock+ 3SiO2(aq) + 23H2O = (Mg5Al)(Si3Al)O10(OH)8 [Clinochlore](3)
3Ca2+ + 2FeOrock + 3SiO2(aq) + 4H2O = Ca3FeIII2Si3O12 [Andradite] + H2(aq) + 6H+(4)
Ca2 + MgOrock + 2SiO2(aq) + H2O = CaMgSi2O6 [Diopside] + 2H+(5)
2. Silicate formation from dissolved species
0.25Ca2+ + 2.5Mg2+ + 0.5Fe2+ + 0.5 Al3+ + 3.5SiO2(aq) + 9H2O = Ca0.25(Mg2.5Fe0.5) (Si3.5Al0.5)O10(OH)2∙4H2O + 8H+(6)
5Mg2+ + Fe2+ + 4SiO2(aq) + 10H2O = Mg5FeSi4O10(OH)8 + 12H+(7)
5Mg2+ + 2Al3+ + 3SiO2(aq) + 12H2O = (Mg5Al)(Si3Al)O10(OH)8 + 16H+(8)
3Ca2+ + 2Fe2+ + 3SiO2(aq) + 6H2O = Ca3FeIII2Si3O12 + 10H+ + H2(aq)(9)
Ca2+ + Mg2+ + 2SiO2(aq) + 2H2O = CaMgSi2O6 + 4H+(10)
Ca2+ + SiO2(aq) + H2O = CaSiO3 [Wollastonite] + 2H+(11)
3. Iron oxidation
Fe2+ + H2O = Fe3+ + OH + 0.5H2(aq)(12)
3FeOrock + H2O = Fe3O4 [Mgt] + H2(aq)(13)
2FeOrock + 2SiO2(aq) + 3H2O = FeIII2Si2O5(OH)4 [FeIII-srp] + H2(aq)(14)
4. Dissolution of solid components
CaOrock + 2H+ = Ca2+ + H2O(15)
FeOrock + 2H+ = Fe2+ + H2O(16)
MgOrock + 2H+ = Mg2+ + H2O(17)
Al2O3rock + 3H2O = 2Al3+ + 6OH(18)
Table 8. Estimates of ammonia and dinitrogen production rates by thermochemical nitrate reduction in the early oceans.
Table 8. Estimates of ammonia and dinitrogen production rates by thermochemical nitrate reduction in the early oceans.
SymbolUnitsEstimate (Hadean to Archaean Eon)
YNH3,ko 0.1
kkg/km31011–1012
Pkokg/km310
FH2O,kokg/y1012–1013
CNO3mol/kg10−6–10–2
FNH3,komol/y105–1010
FN2,komol/y106–1011
FN2mol/y107–1012
See text for the definition of symbol.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nishizawa, M.; Saito, T.; Makabe, A.; Ueda, H.; Saitoh, M.; Shibuya, T.; Takai, K. Stable Abiotic Production of Ammonia from Nitrate in Komatiite-Hosted Hydrothermal Systems in the Hadean and Archean Oceans. Minerals 2021, 11, 321. https://doi.org/10.3390/min11030321

AMA Style

Nishizawa M, Saito T, Makabe A, Ueda H, Saitoh M, Shibuya T, Takai K. Stable Abiotic Production of Ammonia from Nitrate in Komatiite-Hosted Hydrothermal Systems in the Hadean and Archean Oceans. Minerals. 2021; 11(3):321. https://doi.org/10.3390/min11030321

Chicago/Turabian Style

Nishizawa, Manabu, Takuya Saito, Akiko Makabe, Hisahiro Ueda, Masafumi Saitoh, Takazo Shibuya, and Ken Takai. 2021. "Stable Abiotic Production of Ammonia from Nitrate in Komatiite-Hosted Hydrothermal Systems in the Hadean and Archean Oceans" Minerals 11, no. 3: 321. https://doi.org/10.3390/min11030321

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

Article Metrics

Back to TopTop