banner

Blog

Oct 15, 2024

Generation of long-chain fatty acids by hydrogen-driven bicarbonate reduction in ancient alkaline hydrothermal vents | Communications Earth & Environment

Communications Earth & Environment volume 5, Article number: 30 (2024) Cite this article

7043 Accesses

7 Citations

462 Altmetric

Metrics details

The origin of life required membrane-bound compartments to allow the separation and concentration of internal biochemistry from the external environment and establish energy-harnessing ion gradients. Long-chain amphiphilic molecules, such as fatty acids, appear strong candidates to have formed the first cell membranes although how they were first generated remains unclear. Here we show that the reaction of dissolved hydrogen and bicarbonate with the iron-rich mineral magnetite under conditions of continuous flow, alkaline pH and relatively low temperatures (90 °C) generate a range of functionalised long-chain aliphatic compounds, including mixed fatty acids up to 18 carbon atoms in length. Readily generated membrane-forming amphiphilic organic molecules in the first cellular life may have been driven by similar chemistry generated from the mixing of bicarbonate-rich water (equilibrated with a carbon dioxide-enriched atmosphere) with alkaline hydrogen-rich fluids fed by the serpentinisation of the Earth’s iron-rich early crust.

How inorganic geochemical substrates and catalysts ultimately resulted in the first set of organic molecular systems capable of energy harvesting and self-replication is central to understanding the origins of cellular life on the Hadean-Archaean Earth, >3.5 billion years ago. Cell membranes form boundaries, creating internal microenvironments that host metabolic activity and can sustain ion gradients driving CO2 fixation and energy metabolism, a functionality so deeply conserved that it suggests a role for membranes at the origin of life1,2,3,4,5,6,7. The production of some sort of lipid membrane was a pivotal step towards the emergence of cellular life and occurred early in the path from geochemistry to cellular biology8,9. Single-chain amphiphilic monocarboxylic acids (such as fatty acids) with ≥8 carbon atoms can spontaneously self-assemble into vesicles10,11 in aqueous solutions producing semi-permeable barriers that are strong candidates to have been selected as the primitive membranes of the first protocells8,12,13. Despite their importance, the source of the amphiphilic constituents of protocell membranes remains unresolved11,14,15,16,17,18,19.

One compelling potential environment for the origin of life is the mixing zone between bicarbonate and CO2-enriched seawater (and even potentially freshwater)19,20 and the upwelling hydrogen-rich fluids from alkaline hydrothermal vents (AHV)21,22,23,24, that are found today on the deep ocean floor17,25,26 shallow seas20,27, freshwater-tidal mixing zones18,20, and in terrestrial environments28,29,30 but were likely to have been more common on the Hadean-Archaean Earth31,32,33,34. The H2-rich fluids are primarily derived from the serpentinisation of ultramafic rocks, which on mixing with seawater generate far-from-equilibrium conditions conducive to inorganic carbon reduction35 to produce organic compounds36,37,38,39,40. Potentially, the organic compounds generated, including amphiphilic subunits of protocellular membranes, were selected by the precursor biochemistry on Haden-Archaean Earth6,41,42.

Organic compounds in modern AHV fluids of the Lost City hydrothermal vent field are dominated by C9–C14 aliphatic hydrocarbons, C6–C16 aromatic compounds and C8–C18 carboxylic acids, a fraction of which may have been produced abiotically35,43,44,45. Due to the oxygen-free CO2-rich atmosphere, and subsequent abundance of potentially catalytic Fe/Ni-minerals, organic molecular generation at Hadean-Archaean AHV counterparts may have been more efficient38,39,46. Prior electrochemical experiments using greigite (Fe3S4) at room temperature47 and higher temperature batch experiments in the absence of O2 using greigite, magnetite (Fe3O4) and awaruite (Ni3Fe) at temperatures ≥100 °C48,49 and Fe0, Ni0 and Co0 between 30 and 100 °C50 have shown that CO2 can be reduced to C1–C3 organic compounds, some including monocarboxylic acids. However, these experiments used fixed volumes of reactants, thus, the disequilibrium required for CO2 reduction would decay as the reaction proceeded. Continuous flow, bench-top electrochemical reactors51 and oceanic simulant fluids in microfluidic reactors generating pH gradients across Fe(Ni)S precipitates52 have also resulted in the production of formate. Additionally, the reduction of bicarbonate, assisted by cobalt oxides, promoted a carbon−carbon coupling process that generated hydrocarbons up to C24 long at ∼300 °C and 30 MPa53. However, no experiments have yet demonstrated the generation of ≥C8 long-chain fatty acids that are necessary for self-assembling protocell vesicles2,8 from H2 and HCO3− under AHV conditions at <100 °C.

We hypothesised that having both continuous flow and pressurised hydrogen would enable the generation of long carbon chain fatty acids on Fe-minerals. Therefore, we built a pressurised continuous flow reactor to more accurately mimic key elements of the chemistry of upwelling alkaline hydrothermal vent fluid and Hadean-Archaean surface water mixing. An HCO3− solution was mixed with dissolved H2 at concentrations comparable to modern and putative ancient AHV, kept in disequilibrium at 90 °C and 16 bar and passed over magnetite (Fe2O3), a mineral associated with AHV54 (Detailed method in Supplementary Methods). We used a range of complementary bulk and mineral surface-sensitive techniques to analyse the organic molecules generated in experiments, alongside monitoring changes in the inorganic chemistry.

The magnetite and quartz composition before and after the reactions were obtained with X-ray diffraction (Supplementary Note 1 and Supplementary Data 1), which indicated that these minerals were unaffected by the reactions in the AHV simulator. Methane levels above background levels were not observed in the AHV simulator gas trap (Supplementary Methods) in the test experiment using magnetite following the HCO3− and H2 hydrothermal reaction or in the control experiments where HCO3− was absent, or where quartz replaced magnetite (Supplementary Note 2). In contrast, low molecular weight organic acids (C2–C4) were detected (µM concentrations) in the aqueous phase of the test experiment, but not in control experiments (Supplementary Note 3 and Supplementary Data 2) and the presence of such material in the effluent was indicated using 1H nuclear magnetic resonance (Supplementary Note 4). Importantly, after 16 h, even under our low temperature hydrothermal conditions (90 °C, 16 bar total pressure, 5.5 mM H2 and 110 mM HCO3−, pH 8.5 to 9.3) our thermal desorption/volatilisation-gas chromatography-mass spectrometry (TD-GC/MS) analyses demonstrated that a range of functionalised straight and branched long-chain (C10-C20) aliphatic organic compounds were generated on the magnetite after the reaction with H2 and HCO3− (Fig. 1a). This included methyl, hydroxyl, ketones, and ester groups. Triplicate experiments generated molecules with these functional groups, but each replicate resulted in a range of chain lengths and stereochemistry (Supplementary Note 5 and Supplementary Data 3). Importantly, saturated, and unsaturated long-chain (up to C18) monocarboxylic acids (fatty acids) were detected (Fig. 1a). In contrast, these functional groups were undetected in controls with magnetite and H2 but without HCO3− and where quartz replaced magnetite, giving confidence that these organic molecules were generated in the experiments from H2 and HCO3− in the presence of magnetite. The TD-GC/MS peak assignments were confirmed with Attenuated Total Reflectance-Fourier Transform Infra-red Spectroscopy (ATR-FTIR; Fig. 2a–d), and X-ray Photoelectron Spectroscopy (XPS; Table 1, and Supplementary Data 4–6 and Supplementary Note 6). Additionally, bulk Thermogravimetric-differential scanning calorimetry coupled with quadrupole mass spectrometry (TG-DSC-QMS) provided insights into hydrocarbon chain lengths in the samples (Fig. 3a–d) and the mass of organic material generated.

A Comparison of a representative of one of the three replicate total ion chromatograms obtained from reacting magnetite with HCO3− and H2 at 90 °C and 16 bar total pressure for 16 h alongside controls. (a) Organic molecules generated using magnetite with H2 and HCO3−. (b) Control sample using magnetite with H2 (c) Control sample using quartz with H2 and HCO3− (d) Control sample using quartz with H2. B Magnification of the chromatogram region bounded by the broken line box in (A). Peaks with signal: noise ratio <5:1 or with a reverse match factor <65.0 to the NIST20 library are not labelled (Supplementary Data 3 for peak identities obtained from the NIST20 library).

(a) Magnetite with H2 and HCO3−. (b) Control sample using magnetite with H2 (c) Control sample using quartz with H2 and HCO3− (d) Control sample using quartz with H2. A A representative of three ATR-FTIR replicate spectra annotated with candidate functional groups. B Enlarged 3150 cm−1 to 3000 cm−1 region showing carbon-hydrogen bonding, and (C) Enlarged 2000 cm−1–1600 cm−1 region showing carbon-oxygen bonding, the identified peaks are annotated and highlighted by the shaded areas.

CO2 (m/z 44) and H2O (m/z 18) traces resulting from the thermal decomposition of organic material between 35 to 500 °C. (a) Magnetite with H2 and HCO3−. (b) Control sample using magnetite with H2. (c) Control sample using quartz with H2 and HCO3−. (d) Control sample using quartz with H2. Desorption = atmospheric water released from mineral surfaces; Carbonate = water produced from the combustion of hydrogen in the carbonate; vug = water escaping from vugs. Hydrocarbon = components consistent with the combustion of hydrogen in hydrocarbons; Aromatic = component consistent with the combustion of aromatics50,51.

ATR-FTIR indicated that a range of alkane, alkene chains, alcohol and carboxylic acid functional groups were generated on the magnetite after reaction with HCO3− and H2 (Fig. 2); although signal peaks were low, they were consistent with the functional groups in the molecules identified with TD-GC/MS (Fig. 1) and the carbon bonding environment additionally measured by XPS. These peaks were not detected in the magnetite controls without HCO3−, or where quartz replaced magnetite. The chemistry detected by XPS analyses was also consistent with the presence of these functional groups on the magnetite reacted with HCO3− and H2, and not present in magnetite or quartz controls (Table 1). Aromatic functional groups were identified in the magnetite with H2 and HCO3− using ATR-FTIR and the XPS spectra including the control samples (Table 1). Aromatics were not identified in the TD-GC/MS data (Fig. 1a), due to the 350 °C thermal desorption/volatilisation temperature which was likely too low to desorb/volatilise aromatic compounds.

Results from the TG-DSC-QMS (Fig. 3a–d) indicated carbonaceous material with a range of resistance to thermal decomposition was detected on the magnetite following reaction with HCO3− and H2 (Fig. 3a). Increasing resistance to thermal decomposition in TG-DSC-QMS analysis can be interpreted as organic material with increasing molecular weight and complexity55,56. The TG-DSC-QMS m/z 18 traces indicated water was produced from four sources, 1. Trace moisture desorbed from the mineral surface (Fig. 3a–d; annotated Desorption), 2. The thermal decomposition of carbonate (Fig. 3a, c; annotated Carbonate) 3. Water escaping from vugs in the quartz samples (Fig. 3c, d; vug) 4. The combustion of hydrogen (Fig. 3a; hydrocarbons). The combustion of hydrogen, alongside the production of CO2 (m/z 44), was consistent with hydrocarbon-based compounds. Given the range of combustion temperatures, we suggest that the TG-DSC-QMS data indicates hydrocarbons with a range of molecular weights corresponding to the aliphatic compounds with a range of chain lengths, detected using TD-GC/MS (Fig. 1a) and the presence of aromatised organic material50,51, which was supported by the ATR-FTIR (Fig. 2d) and XPS (Table 1) data. The increase in CO2 (m/z 44) as a function of temperature may reflect either higher frequencies of long-chain hydrocarbon or higher carbon number molecules and the decrease in water (m/z 18) suggests a decrease in the carbon:hydrogen ratio in the putative hydrocarbons. An additional experiment using magnetite with H2 and 13C labelled NaHCO3− showed that 13C labelled organic compounds were generated (Fig. 4a), providing confidence that the organic material observed in the analysis of the magnetite with H2 and HCO3− was indeed derived from in situ CO2 reduction. The concurrent evolution of water (m/z 18) similarly implied the production of 13C hydrocarbons that also included a component consistent with aromatic compounds, similar to the m/z 44 trace in Fig. 3d. The TG-DSC-QMS analysis indicated an average (n = 3) of 1.6 mg g C-1 was generated. This was supported by total organic carbon combustion analysis indicating an average (n = 3) of 1.5 mg C g-1 dry weight organic carbon was added to the magnetite (Supplementary Note 7; Supplementary Data 7). In contrast, no fresh carbonaceous material was detected on the control magnetite without HCO3− (Fig. 3b) and where quartz replaced magnetite (Fig. 3c, d).

(a) 13CO2 (m/z 45), (b) 12CO2 (m/z 44) and (c) H2O (m/z 18). Desorption = atmospheric water released from mineral surfaces; Carbonate = water produced from the combustion of hydrogen in the carbonate; Hydrocarbon = components consistent with the combustion of hydrogen in hydrocarbons. Aromatic = component consistent with the combustion of aromatics50,51.

The average (n = 3) high-resolution Fe2p XPS analyses showed a ca. 1–10 nm layer of FeOOH covering the magnetite Fe2O3 surface (Supplementary Data 6). The ratio of Fe2OOH: Fe2O3 increased by an average of 8.9% on the test magnetite with H2 and HCO3− but not in the control magnetite without HCO3−. This suggested that the net oxidation of the magnetite surface was related to the fixation of HCO3− during the hydrothermal reactions, thus serving as a reactant rather than a catalyst.

Aliphatic organic material was readily synthesised in our continuous flow reactor mimicking H2-rich upwelling AHV fluids mixing with HCO3−-rich ocean or surface water equilibrated with a CO2-rich Hadean-Archaean atmosphere (Supplementary Methods). Long-chain saturated and unsaturated fatty acids and associated amphiphilic molecules such as carboxylic acids and alcohols were synthesised on a magnetite surface at low temperatures (90 °C) in the presence of H2 and HCO3−.

Vesicle formation requires amphiphilic molecules with chains ≥C8 long10. Earlier experiments mimicking aspects of AHV vent fluid with HCO3− have only generated these longer chain fatty acids at higher temperatures (optimally 175 °C) in the presence of high concentrations of formate or oxalate and long-chain (up to C24) hydrocarbon generation used temperatures and pressures of ∼300 °C and 30 MPa53. Our study is the first to demonstrate the facile generation of complex organic molecules including long-chain fatty acids under mild hydrothermal conditions (<100 °C) using realistic reactants (H2 and HCO3−) on a magnetite mineral surface. Our success in generating organic molecules >C3 may partially reflect that the organic material measured on mineral surfaces detected in earlier studies50,57 were not characterised.

We suspect that using both continuous flow and elevated H2 pressures may have favoured continuous long-chain aliphatic molecule synthesis, due to increased H2 solubility58 and continuous disequilibrium59. This constant H2 flux will have provided a source of reducing electrons for continual CO2 reduction and potentially promoting carbon−carbon coupling processes53 or a Fisher-Tropsch type (FTT) reaction, commonly invoked to explain geochemically driven organic synthesis60,61. Alongside sequential carbon-carbon addition, the presence of other groups suggests other non-sequential carbon-carbon bond formation may have taken place (Figs. 1, 2). We speculate that these non-sequential reactions may involve addition reactions involving the products of prior reduction reactions (e.g., formaldehyde or formate) producing hydroxyl, ketone and ether groups (Figs. 1, 2 and Table 1).

We suggest that the functionalised aliphatics, including the long-chain fatty acids, are formed through a reduction of inorganic carbon on a Fe surface48,50 generating an initial carboxyl compound, followed by sequential addition of C1 compounds49 or FTT synthesis56,57. We speculate that mineral surface topography and defects could influence this generation resulting in the random chain lengths and stereochemistry observed in the replicate experiments using TD-GC/MS (Supplementary Data 3). The high-resolution Fe2p XPS analysis (Supplementary Data 6) showed that the surface of the magnetite with H2 and HCO3− samples was oxidised. This suggests that e- donation from the magnetite surface promoted reduction during the C1 addition or FTT reaction, potentially associated with the transfer of oxygen from the organic compounds to the Fe mineral surface. Hence, the addition reaction is partly through a surface-promoted alteration of the magnetite substrate, rather than the magnetite acting as a true catalyst. This implies that oxygen is transferred to the surface rather than being reduced to water, thus no strict dehydration reaction may occur in the FTT reaction.

Our results showed the length of the core hydrocarbon chain moieties varied and there was a complex range of methyl and alkyl branches, keto, hydroxyl and importantly carboxyl functional groups associated with the core carbon chain. Flaws on the magnetite surface could likely result in non-linear bond formation, producing alkenes or methyl groups observed in the study. This implies the vesicles generated from such synthesis would plausibly consist of a consortium of different molecules that are reported to produce more stable vesicles62. Additionally, biological cell membranes require several types of lipid subunits to accomplish their physiological role63,64. A steady output of vesicles comprised of random amphiphilic subunits with a range of properties may have permitted prebiotic life to utilise protocellular vesicles that were most conducive to energy harvesting and cellular division.

Notably, if the mineral surface-bound organic molecules generated in our experiments were rapidly released into the 3.4 mL effluent of our reactor tube (e.g., by an elevation in pH causing electrostatic repulsion of the deprotonated carboxylic acids with a negatively charged mineral surface)65 would generate >100 mM concentrations of organic molecules compared to 1.3 µM to mM concentrations of mixed fatty acids required to form self-assembling vesicles62. Such vesicles may have served as a primitive membrane that preceded more complex modern counterparts66,67,68,69.

Inorganic membranes such as iron monosulphide precipitates4,6, mineral gels70 or the interlayers of clay minerals71,72 have been proposed as pre-cellular inorganic membranes allowing the evolution of internal biochemistry and producing membrane-bound ion gradients capable of harnessing energy, before the development of modern organic membranes4. In the oceans of the early Earth, inorganic membranes may have allowed the development of chemical disequilibria, with higher molecular concentrations and resultant chemical gradients relative to the external environment, permitting CO2 reduction by H24,73. However, they are typically a thousand times thicker than the few nm thicknesses of lipid membranes. The synthesised amphiphilic molecules observed in our investigation suggest that organic vesicles more similar to the membranes of extant biological cells may have been readily available from AHV mixing chemistry.

We conclude that the mixing of H2-rich fluid upwelling from AHV with HCO3−-rich water over the Fe-mineral magnetite on the early Earth has the potential to be the source of primitive amphiphilic membranes of the first protocells at the dawn of life. We suggest that this synthesis was a single carbon addition or a Fischer-Tropsch type process, synthesising organic compounds that may be conducive to the formation of a range of bilayer membranes, some of which could have been selected as cell membranes in the origin of life. We speculate that the mechanism described here may also have contributed to the generation of some of the carboxylic acids and fatty acids detected in the protoplanetary precursors to meteorites74. We further speculate that low temperature amphiphilic membrane generation may still be active today in analogous reactions in the subsurface of icy moons in our solar system75,76, contributing to the possibility of other origins of life in these worlds.

The AHV simulator was assembled according to the Supplementary Methods. All experiments used Type 1 18.2 MΩ.cm ultrapure water and all glassware and metalware were heat treated to 120 °C for 120–180 min before use and magnetite and quartz were confirmed using XRD (Supplementary Note 1 and Supplementary Data 1). Components were connected using 1/16-inch PTFE tubing (Darwin Microfluidics, Paris, France). A 130 mM aqueous solution of either NaHCO3 (Sigma-Aldrich, UK; CAS: 144-55-8; Product No. S6014) or 13C labelled NaHCO3 (Sigma-Aldrich, UK; CAS: 87081-58-1; Product No. 372382) was prepared. The pH was adjusted to pH 7.0 using 100 mM HCl. We calculate that the final HCO3− was 110 mM due to CO2 degassing and that 20 mM NaCl was present in the solution resulting from the addition of HCl to the NaHCO3 or 13C NaHCO3 solution. The solution was then purged with N2 for a minimum of 1 h before use. In experiments, the pH changed from pH 7.0 to pH 8.5-9.3 within 3 h (Supplementary Data 8). The solution was pumped under pressure using a Gilson 305 HPLC pump with a Gilson 805 monomatic module (Gilson Inc. Middleton WI, USA). The reactant pressure was monitored using the integrated HPLC pump manometer. The reactor pressure was brought to 16 bar using a flow rate of 2 mL min−1 then the flow rate was set to 0.1 mL min-1, producing a fluid velocity of 500 mm.h-1. H2 was delivered via a ‘Gastropod’ tube-in-tube gas exchange system77 (Cambridge Reactor Designs Ltd, Cambridgeshire, UK) at 7.0 bar from an H2 generator (Peak Scientific Ltd, Inchinnan, UK). This provided a dissolved H2 equivalent of approximately 5 mM H2. An HPLC column serving as a heat exchanger (Thames Restek UK Ltd, Bucks, UK) heated fluids to 90 °C using an HPLC column block heater (Jones Chromatography Ltd, Hengoed, UK). Fluid was passed over a magnetite (CAS No. 215-169-2; Inoxia Ltd, Cranleigh, UK) or quartz (CAS No. 14808-60-7; Analar grade, Sigma-Aldrich Ltd, UK; Product code S6014-500g) bed at 90 °C. 16 bar was maintained with a back pressure valve, then gas was collected in a gas trap and the effluent was collected.

1 mL aliquots were taken from the concentrated control and test effluents and filtered using a 0.22 μm polyethylene syringe filter (VWR International, UK). The supernatants were acidified using neat (37%) hydrochloric acid until all NaHCO3 was decomposed to CO2. C1 to C5 carboxylic acids were selected for analysis (Supplementary Note 3). Analysis was conducted using the ‘Dionex Aquion’ system. 25 µL of the sample was injected into liquid Ion Chromatography (Dionex ICS-1000) equipped with a Dionex IonPac ICE_AS1 4 × 250 mm column (Thermo Scientific, UK) and a Dionex ACRS-ICE500 4 mm suppressor with 1 mM heptafluorobutyric acid as the eluent with an isocratic flow rate 0.16 mL min−1 for 35 mins and using 5 mM tetrabutylammonium hydroxide as the regenerant.

Freeze-dried control and test mineral samples were analysed using infra-red spectroscopy. Liquid effluent was pipetted onto the ATR platform and dried in air for 6–8 h before analysis immediately followed by a background deduction. Analysis was conducted in the mid to far-infra-red range, using a Nicolet 4100 Fourier transform infra-red (FTIR) spectrometer (Thermo Fisher Scientific, MA, U.S.A.), with a diamond platform attenuated total reflectance (ATR) module (Specac Ltd. London, UK) attached. Background scans were taken within 30 min of each analytical scan. Broad scan spectra were obtained from 128 replicate scans conducted between 4000 cm−1 to 550 cm−1, in 4 cm−1 steps. Narrow scan spectra were obtained from 128 replicate scans conducted between 3150 cm−1 to 3000 cm−1 and 2000 cm−1 to 1610 cm−1 in 1 cm steps. Spectra were analysed using Omnic 8.2.0.387 software (Thermo Fisher Scientific, MA, U.S.A.). Band identification was conducted using the Sigma Aldrich FTIR table (www.sigmaaldrich.com/technical-documents/articles/biology/ir-spectrum-table.html).

20 mg of freeze-dried mineral samples were placed in stainless steel pyrolysis cups and then mounted on Frontier Labs 3030S Single shot pyrolyser unit, set to 350 °C and linked to the GC/MS via a CDS 1500 valved interface that was maintained at 300 °C. The products passed into an HP6890 gas chromatograph (GC) with a 60 m HP-5MS column (0.25 mm internal diameter, 0.25 µm film thickness; J&W Scientific, USA). A helium carrier gas was used at a flow rate of 1 mL min-1. The GC was held at 50 °C for 1 min, and then the temperature was ramped from 50 to 310 °C at 5 °C min−1 and held at the final temperature for 12 min, for a total run time of 60 min. Product detection was carried out using an HP5972 series mass selective detector in full scan mode (50–650 amu). Data acquisition was controlled by an HP kayak xa ChemStation computer. The chromatograms were analysed using Openchrom software (www.openchrom.net). Traces were smoothed using a Savitzky–Golay filter then sensitive nonlinear iterative peak baselines were obtained and then deducted from the chromatogram traces. The peaks with a signal-to-noise ratio >5:1 were acquired. Peak identities with reverse match factors >70% were obtained using the NIST20 mass spectra library and compound classes assigned to the compounds that could be identified78.

A 1 mm deep layer of freeze-dried mineral was immobilised by compaction into Al discs according to Purvis et al. (2019)78. The discs were held in place on SEM stubs using double-sided carbon tape. This 1 mm deep sample layer was in excess of the 1–10 nm depth of analysis. The surface of the sample was decontaminated according to Purvis et al. (2017)79 using Ar gas clusters centred around 1000 atoms per cluster at 4 keV, for 90 s. XPS analysis was conducted using a Thermo Scientific K-alpha X-ray Photoelectron Spectrometer™ (Thermo Scientific, East Grinstead, UK). Survey spectra (broad energy, multiple elements) were used to obtain the quantitative elemental data from the scans between -5.0 eV and 1350.0 eV; a dwell time of 50 ms; a step size of 1.0 eV and a pass energy of 200.0 eV.

High-resolution (narrow scan, single element) carbon 1 s (C1s), and iron (Fe2p 1/2) spectra were collected at 40 eV pass energy with a 0.1 eV step size. Spectra were acquired using a monochromatic Al Kα X-ray source with an output energy of 1486.6 eV with a maximum X-ray beam spot size of 400 μm. Surface charge compensation was obtained with a low-energy dual-beam electron/ion flood gun operated at 40 eV. Spectral analysis was carried out using CasaXPS software (CasaXPS Ltd. Teignmouth, UK) with major peaks selected for element identification using the Handbook of X-ray photoelectron spectroscopy80. Any energy shifts were calibrated using the Na1s binding energy at 1072.0 eV. C1s synthetic components were fitted according to Purvis et al. (2019)78. Shirley backgrounds were applied, and then Gaussian synthetic components for each chemical bond were added to the spectra, centring on their known binding energy80. The components were fitted to the C1s peak automatically using the Levenberg–Marquardt algorithm within CasaXPS. The unknown N1s peaks were assigned using the data from the NIST XPS (srdata.nist.gov/xps/Default.aspx) and La Surface online databases (www.lasurface.com).

Thermal analysis was carried out to determine the relative proportions of different C components according to Lopez-Capel et al (2005)55 extended by the determination of evolved gases. A Netzsch STA 449C Jupiter TG-DSC system connected by a heated capillary (200 °C) to a Netzsch Aeolos quadrupole mass spectrometer system was used to determine the following parameters: (1) mass change, (2) energy gained or lost from the sample, (3) composition of evolved gases (m/z range 10–300). The freeze-dried control and test quartz or magnetite samples (typically 15 mg) were placed within an alumina crucible, using an identical empty reference crucible, and were heated from ambient temperature to 990 °C under flowing an 80:20 He/O2 gas mixture. Intensities for m/z 44 (CO2 + N2O) and m/z 18 (H2O) were recorded. No correction was made to m/z 44 for evolved N2O, which was assumed to be negligible compared with CO2 given the failure to detect measurable m/z 30 (NO). Digitised evolved gas traces were smoothed using a Savitzky–Golay filter and then interpreted using GRAMS/AI peak fitting software (www.AdeptScience.co.uk) that permits investigation of the contribution of several individual peaks to an overall trace. By defining initial conditions (including a presumed Gaussian peak shape), the range of potential values for each peak centre (125 °C, 125–450 °C) (labelled i-vi) and the number of iterations to perform (in this case 50), the software fits individual aliphatic hydrocarbon carbon peaks56 such that their combined shapes and areas are the best possible fit for the overall CO2 spectrum.

Headspace gas was removed from the reactor gas trap and 8 mL was transferred to pre-evacuated Exetainers (Labco, Lampeter, Wales). 50 µL were removed and injected into an Agilent 8890 GC coupled to an Agilent 5977B GC/MSD. This was operated in a He stream in continuous flow mode with a split injector (10:1 split ratio) and a flow rate of 1.5 mL.min−1 at 6.470 psi with a column temperature of 100 °C. The mass spectrum was operated at 250 °C. Chromatograms were analysed using Openchrom software (www.openchrom.net) and the total ion chromatogram was filtered for m/z 16.

The TOC was measured using a Leco RC612 multiphase carbon and hydrogen/moisture determinator (Leco Corp, St Joseph, MI, USA). Approximately 150 mg of the freeze-dried magnetite or quartz control and test samples were added to ceramic crucibles. The samples were then combusted at 1100 °C in an O2 environment. An afterburner furnace (850 °C) and a secondary oxidation catalyst ensured full combustion of all evolved components, which were then detected using IR.

Raw data for all analyses for all samples is available at from Newcastle University Library data repository https://doi.org/10.25405/data.ncl.12721631.v1 For all samples: XPS spectra is in the native.vms file format; TG-DCS-QMS spectra are digitised and stored in a.xls file format; XRD and 1HNMR spectra are stored as images. ATR-FTIR data is stored as .spa files, GC/MS and TD-GC/MS are stored as .ocb files.

Mulkidjanian, A. Y., Galperin, M. Y. & Koonin, E. V. Co-evolution of primordial membranes and membrane proteins. Trends Biochem. Sci. 34, 206–215 (2009).

Article CAS Google Scholar

Gill, S. & Forterre, P. Origin of life: LUCA and extracellular membrane vesicles (EMVs). Int. J. Astrobiol. 15, 7–15 (2016).

Article CAS Google Scholar

Deamer, D. Membranes and the origin of life: a century of conjecture. J. Mol. Evol. 83, 159–168 (2016).

Article CAS Google Scholar

Lane, N. Proton gradients at the origin of life. BioEssays 39, 1600217–1600217 (2017).

Article Google Scholar

Lane, N. Why are cells powered by proton gradients? Nat. Educ. 3, 2 (2010).

Google Scholar

Lane, N., Allen, J. F. & Martin, W. How did LUCA make a living? Chemiosmosis in the origin of life. BioEssays 32, 271–280 (2010).

Article CAS Google Scholar

Lane, N. & Martin, W. F. The origin of membrane bioenergetics. Cell 151, 1406–1416 (2012).

Article CAS Google Scholar

Segré, D., Ben-Eli, D., Deamer, D. W. & Lancet, D. The lipid world. Orig. Life Evol. Biosph. 31, 119–145 (2001).

Article Google Scholar

Chen, I. A. & Walde, P. From self-assembled vesicles to protocells. Cold Spring Harb. Perspect. Biol. 2, a002170 (2010).

Article Google Scholar

Apel, C. L., Deamer, D. W. & Mautner, M. N. Self-assembled vesicles of monocarboxylic acids and alcohols: conditions for stability and for the encapsulation of biopolymers. Biochim. Biophys. Acta 1559, 1–9 (2002).

Article CAS Google Scholar

Cohen, Z. R., Todd, Z. R., Wogan, N., Black, R. A., Keller, S. L. & Catling, D. C. Plausible sources of membrane-forming fatty acids on the early earth: a review of the literature and an estimation of amounts. ACS Earth Space Chem. 22, 11–27 (2022).

Google Scholar

Luisi, P. L., Walde, P. & Oberholzer, T. Lipid vesicles as possible intermediates in the origin of life. Curr. Opin. Colloid Interface Sci. 4, 33–39 (1999).

Article CAS Google Scholar

Deamer, D. The role of lipid membranes in life’s origin. Life 7, 5–5 (2017).

Article Google Scholar

Deamer, D., Dworkin, J. P., Sandford, S. A., Bernstein, M. P. & Allamandola, L. J. The first cell membranes. Astrobiology 2, 371–381 (2002).

Article CAS Google Scholar

Dworkin, J. P., Deamer, D. W., Sandford, S. A. & Allamandola, L. J. Self-assembling amphiphilic molecules: Synthesis in simulated interstellar/precometary ices. Proc. Natl. Acad. Sci. 98, 815–819 (2001).

Article CAS Google Scholar

Pohorille, A. & Deamer, D. Self-assembly and function of primitive cell membranes. Res. Microbiol. 160, 449–456 (2009).

Article CAS Google Scholar

Sojo, V., Herschy, B., Whicher, A., Camprubi, E. & Lane, N. The origin of life in alkaline hydrothermal vents. Astrobiology 16, 181–197 (2016).

Article CAS Google Scholar

Monnin, C., Chavagnac, V., Boulart, C., Ménez, B., Gérard, M. & Gérard, E. et al. Fluid chemistry of the low-temperature hyperalkaline hydrothermal system of Prony Bay (New Caledonia). Biogeosciences 11, 5687–5706 (2014).

Article Google Scholar

Monnin, C., Chavagnac, V., Boulart, C., Ménez, B., Gérard, M. & Gérard, E. et al. The low temperature hyperalkaline hydrothermal system of the Prony bay (New Caledonia). Biogeosci. Discuss. 11, 6221–6267 (2014).

Google Scholar

Barge, L. M. & Price, R. E. Diverse geochemical conditions for prebiotic chemistry in shallow-sea alkaline hydrothermal vents. Nat. Geosci. 15, 976–981 (2022).

Article CAS Google Scholar

Russell, M. & Arndt, N. Geodynamic and metabolic cycles in the Hadean. Biogeosciences 2, 97–111 (2005).

Article CAS Google Scholar

Russell, M. J. The alkaline solution to the emergence of life: energy, entropy and early evolution. Acta Biotheor. 55, 133–179 (2007).

Article Google Scholar

Russell, M. J., Daniel, R. M., Hall, A. J. & Sherringham, J. A. A hydrothermally precipitated catalytic iron sulfide membrane as a first step toward life. J. Mol. Evol. 39, 231–243 (1994).

Article CAS Google Scholar

Russell, M. J., Nitschke, W. & Branscomb, E. The inevitable journey to being. Philos. Trans. R. Soc. B Biol. Sci. 368, 20120254 (2013).

Article Google Scholar

Martin, W. & Russell, M. J. On the origin of biochemistry at an alkaline hydrothermal vent. Philos. Trans. R. Soc. B Biol. Sci. 362, 1887–1925 (2007).

Article CAS Google Scholar

Kelley, D. S., Karson, J. A., Blackman, D. K., FruÈh-Green, G. L., Butterfield, D. A. & Lilley, M. D. et al. An off-axis hydrothermal vent field near the Mid-Atlantic Ridge at 30 N. Nature 412, 145–149 (2001).

Article CAS Google Scholar

Marteinsson, V. T. R., Kristjánsson, J. K., Kristmannsdóttir, H., Dahlkvist, M., Sæmundsson, K. N. & Hannington, M. et al. Discovery and description of giant submarine smectite cones on the seafloor in Eyjafjordur, northern Iceland, and a novel thermal microbial habitat. Appl. Environ. Microbiol. 67, 827–833 (2001).

Article CAS Google Scholar

Suda, K., Ueno, Y., Yoshizaki, M., Nakamura, H., Kurokawa, K. & Nishiyama, E. et al. Origin of methane in serpentinite-hosted hydrothermal systems: The CH4–H2–H2O hydrogen isotope systematics of the Hakuba Happo hot spring. Earth. Planetary Sci. Lett. 386, 112–125 (2014).

Article CAS Google Scholar

Morrill, P. L., Kuenen, J. G., Johnson, O. J., Suzuki, S., Rietze, A. & Sessions, A. L. et al. Geochemistry and geobiology of a present-day serpentinization site in California: the Cedars. Geochim. Cosmochim. Acta 109, 222–240 (2013).

Article CAS Google Scholar

Leong, J., Howells, A., Robinson, K., Cox, A., Debes, R. & Fecteau, K. et al. Theoretical predictions versus environmental observations on serpentinization fluids: lessons from the Samail ophiolite in Oman. J. Geophys. Res. Solid Earth 126, e2020JB020756 (2021).

Article Google Scholar

Holm, N. G. & Andersson, E. Hydrothermal simulation experiments as a tool for studies of the origin of life on earth and other terrestrial planets: A review. Astrobiology 5, 444–460 (2005).

Article CAS Google Scholar

Fyfe, W. S. The water inventory of the Earth: fluids and tectonics. Geol. Soc. Lond. Spec. Publ. 78, 1–7 (1994).

Article Google Scholar

Jaffrés, J. B. D., Shields, G. A. & Wallmann, K. The oxygen isotope evolution of seawater: a critical review of a long-standing controversy and an improved geological water cycle model for the past 3.4 billion years. Earth-Sci. Rev. 83, 83–122 (2007).

Article Google Scholar

Shields, G. A. & Kasting, J. F. Evidence for hot early oceans? Nature 447, E1–E1 (2007).

Article CAS Google Scholar

Proskurowski, G., Lilley, M. D., Seewald, J. S., Fruh-Green, G. L., Olson, E. J. & Lupton, J. E. et al. Abiogenic hydrocarbon production at Lost City hydrothermal field. Science 319, 604–607 (2008).

Article CAS Google Scholar

Horita, J. & Berndt, M. E. Abiogenic methane formation and isotopic fractionation under hydrothermal conditions. Science 285, 1055–1057 (1999).

Article CAS Google Scholar

Shock, E. L. & Schulte, M. D. Organic synthesis during fluid mixing in hydrothermal systems. J. Geophys. Res. Planets 103, 28513–28527 (1998).

Article CAS Google Scholar

Charlou, J. L., Donval, J. P., Fouquet, Y., Jean-Baptiste, P. & Holm, N. Geochemistry of high H2 and CH4 vent fluids issuing from ultramafic rocks at the Rainbow hydrothermal field (36 14′ N, MAR). Chem. Geol. 191, 345–359 (2002).

Article CAS Google Scholar

Holm, N. G. & Charlou, J. L. Initial indications of abiotic formation of hydrocarbons in the Rainbow ultramafic hydrothermal system, Mid-Atlantic Ridge. Earth. Planet. Sci. Lett. 191, 1–8 (2001).

Article CAS Google Scholar

Shock, E. & Canovas, P. The potential for abiotic organic synthesis and biosynthesis at seafloor hydrothermal systems. Geofluids 10, 161–192 (2010).

Article CAS Google Scholar

Russell, M. J. From geochemistry to biochemistry: chemiosmotic coupling and transition element clusters in the onset of life and photosynthesis. Geochem. News 113, 6–12 (2002).

Google Scholar

Kitadai, N. & Maruyama, S. Origins of building blocks of life: a review. Geosci. Front. 9, 1117–1153 (2018).

Article CAS Google Scholar

Konn, C., Charlou, J.-L., Donval, J.-P., Holm, N. G., Dehairs, F. & Bouillon, S. Hydrocarbons and oxidized organic compounds in hydrothermal fluids from Rainbow and Lost City ultramafic-hosted vents. Chem. Geol. 258, 299–314 (2009).

Article CAS Google Scholar

Longnecker, K., Sievert, S. M., Sylva, S. P., Seewald, J. S. & Kujawinski, E. B. Dissolved organic carbon compounds in deep-sea hydrothermal vent fluids from the East Pacific Rise at 9 50′ N. Org. Geochem. 125, 41–49 (2018).

Article CAS Google Scholar

Proskurowski, G., Lilley, M. D., Kelley, D. S. & Olson, E. J. Low-temperature volatile production at the Lost City Hydrothermal Field, evidence from a hydrogen stable isotope geothermometer. Chem. Geol. 229, 331–343 (2006).

Article CAS Google Scholar

Lang, S. Q., Butterfield, D. A., Schulte, M., Kelley, D. S. & Lilley, M. D. Elevated concentrations of formate, acetate and dissolved organic carbon found at the Lost City hydrothermal field. Geochim. Cosmochim. Acta 74, 941–952 (2010).

Article CAS Google Scholar

áde Leeuw, N. H. Bio-inspired CO 2 conversion by iron sulfide catalysts under sustainable conditions. Chem. Commun. 51, 7501–7504 (2015).

Article Google Scholar

Preiner, M., Igarashi, K., Muchowska, K. B., Yu, M., Varma, S. J. & Kleinermanns, K. et al. A hydrogen-dependent geochemical analogue of primordial carbon and energy metabolism. Nat. Ecol. Evol. 4, 534–542 (2020).

Article Google Scholar

Cody, G. D., Boctor, N. Z., Filley, T. R., Hazen, R. M., Scott, J. H. & Sharma, A. et al. Primordial carbonylated iron-sulfur compounds and the synthesis of pyruvate. Science 289, 1337–1340 (2000).

Article CAS Google Scholar

Varma, S. J., Muchowska, K. B., Chatelain, P. & Moran, J. Native iron reduces CO2 to intermediates and end-products of the acetyl-CoA pathway.Nat. Ecol. Evol. 2, 1019–1024 (2018).

Article Google Scholar

Herschy, B., Whicher, A., Camprubi, E., Watson, C., Dartnell, L. & Ward, J. et al. An origin-of-life reactor to simulate alkaline hydrothermal vents. J. Mol. Evol. 79, 213–227 (2014).

Article CAS Google Scholar

Hudson, R., de Graaf, R., Strandoo Rodin, M., Ohno, A., Lane, N. & McGlynn, S. E. et al. CO2 reduction driven by a pH gradient. Proc. Natl. Acad. Sci. 117, 22873–22879 (2020).

Article CAS Google Scholar

He, D., Wang, X., Yang, Y., He, R., Zhong, H. & Wang, Y. et al. Hydrothermal synthesis of long-chain hydrocarbons up to C24 with NaHCO3− assisted stabilizing cobalt. Proc. Natl. Acad. Sci. 118, e2115059118 (2021).

Article CAS Google Scholar

Russell, M. J. & Ponce, A. Six ‘must-have’minerals for life’s emergence: Olivine, pyrrhotite, bridgmanite, serpentine, fougerite and mackinawite. Life 10, 291–291 (2020).

Article CAS Google Scholar

Lopez-Capel, E., Sohi, S. P., Gaunt, J. L. & Manning, D. A. Use of thermogravimetry–differential scanning calorimetry to characterize modelable soil organic matter fractions. Soil Sci. Soc. Am. J. 69, 136–140 (2005).

Article CAS Google Scholar

Manning, D. A., Renforth, P., Lopez-Capel, E., Robertson, S. & Ghazireh, N. Carbonate precipitation in artificial soils produced from basaltic quarry fines and composts: an opportunity for passive carbon sequestration. Int. J. Greenh. Gas Control 17, 309–317 (2013).

Article CAS Google Scholar

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. 116, 4828–4833 (2019).

Article CAS Google Scholar

Pray, H. A., Schweickert, C. & Minnich, B. H. Solubility of hydrogen, oxygen, nitrogen, and helium in water at elevated temperatures. Ind. Eng. Chem. 44, 1146–1151 (1952).

Article CAS Google Scholar

McCollom, T. M., Ritter, G. & Simoneit, B. R. T. Lipid synthesis under hydrothermal conditions by Fischer-Tropsch-type reactions. Orig. Life Evol Biosph. 29, 153–166 (1999).

Article CAS Google Scholar

McCollom, T. M. Laboratory simulations of abiotic hydrocarbon formation in Earth’s deep subsurface. Carbon Earth 75, 467–494 (2013).

Article CAS Google Scholar

McCollom, T. M. & Seewald, J. S. Abiotic synthesis of organic compounds in deep-sea hydrothermal environments. Chem. Rev. 107, 382–401 (2007).

Article CAS Google Scholar

Jordan, S. F., Rammu, H., Zheludev, I. N., Hartley, A. M., Maréchal, A. & Lane, N. Promotion of protocell self-assembly from mixed amphiphiles at the origin of life. Nat. Ecol. Evol. 3, 1705–1714 (2019).

Article Google Scholar

Lombard, J. Once upon a time the cell membranes: 175 years of cell boundary research. Biol. Direct. 9, 1–35 (2014).

Article Google Scholar

Rawicz, W., Olbrich, K. C., McIntosh, T., Needham, D. & Evans, E. Effect of chain length and unsaturation on elasticity of lipid bilayers. Biophys. J. 79, 328–339 (2000).

Article CAS Google Scholar

Snabe, T., Neves-Petersen, M. T. & Petersen, S. B. Enzymatic lipid removal from surfaces—lipid desorption by a pH-induced “electrostatic explosion”. Chem. Phys. Lipids 133, 37–49 (2005).

Article CAS Google Scholar

Koga, Y. Early evolution of membrane lipids: how did the lipid divide occur? J. Mol. Evol. 72, 274–282 (2011).

Article CAS Google Scholar

Koga, Y., Kyuragi, T., Nishihara, M. & Sone, N. Did archaeal and bacterial cells arise independently from noncellular precursors? A hypothesis stating that the advent of membrane phospholipid with enantiomeric glycerophosphate backbones caused the separation of the two lines of descent. J. Mol. Evol. 46, 54–63 (1998).

Article CAS Google Scholar

Martin, W. & Russell, M. J. On the origins of cells: a hypothesis for the evolutionary transitions from abiotic geochemistry to chemoautotrophic prokaryotes, and from prokaryotes to nucleated cells. Philos. Trans. R. Soc. Lond. Ser B: Biol. Sci. 358, 59–85 (2003).

Article CAS Google Scholar

West, T., Sojo, V., Pomiankowski, A. & Lane, N. The origin of heredity in protocells. Philos. Trans. R. Soc. B: Biol. Sci. 372, 20160419–20160419 (2017).

Article Google Scholar

Westall, F. & Brack, A. The importance of water for life. Space Sci. Rev. 214, 1–23 (2018).

Article Google Scholar

Cairns-Smith A. G., Hartman H. Clay minerals and the origin of life. (CUP Archive 1986).

Kloprogge, J. T. & Hartman, H. Clays and the origin of life: the experiments. Life 12, 259–259 (2022).

Article CAS Google Scholar

Vasiliadou, R., Dimov, N., Szita, N., Jordan, S. F. & Lane, N. Possible mechanisms of CO2 reduction by H2 via prebiotic vectorial electrochemistry. Interface Focus 9, 20190073 (2019).

Article Google Scholar

Lai, J. C.-Y., Pearce, B. K., Pudritz, R. E. & Lee, D. Meteoritic abundances of fatty acids and potential reaction pathways in planetesimals. Icarus 319, 685–700 (2019).

Article CAS Google Scholar

Barge, L. M. & Rodriguez, L. E. Life on Enceladus? It depends on its origin. Nat. Astron. 5, 740–741 (2021).

Article Google Scholar

Russell, M. J., Barge, L. M., Bhartia, R., Bocanegra, D., Bracher, P. J. & Branscomb, E. et al. The drive to life on wet and icy worlds. Astrobiology 14, 308–343 (2014).

Article CAS Google Scholar

O’Brien, M., Taylor, N., Polyzos, A., Baxendale, I. R. & Ley, S. V. Hydrogenation in flow: Homogeneous and heterogeneous catalysis using Teflon AF-2400 to effect gas–liquid contact at elevated pressure. Chem. Sci. 2, 1250–1257 (2011).

Article Google Scholar

Purvis, G., Sano, N., van der Land, C., Barlow, A., Lopez-Capel, E. & Cumpson, P. et al. Combining thermal hydrolysis and methylation-gas chromatography/mass spectrometry with X-ray photoelectron spectroscopy to characterise complex organic assemblages in geological material. MethodsX 6, 2646–2655 (2019).

Purvis, G., Gray, N., Sano, N., Barlow, A., Cockell, C. & Abbott, G. D. et al. Decontamination of geological samples by gas cluster ion beam etching or ultraviolet/ozone. Chem. Geol. 466, 256–262 (2017).

Article CAS Google Scholar

Chastain, KingR. C. J. Handbook of X-ray photoelectron spectroscopy: a reference book of standard spectra for identification and interpretation of XPS data. (Physical Electronics.: Eden Prairie, MN, 1995).

Google Scholar

Download references

This research project has also been funded by a Natural and Environmental Research Council grant NE/W005506/1 (to J.T.) and by a discipline Hopping Grant awarded by Newcastle University (reference: NE/X018148/1) and Authors wish to thank Bill Guthery for his advice on pyrolysis-GC/MS and Casey Dixon for operating the NMR.

School of Natural and Environmental Sciences, Newcastle University, Newcastle upon Tyne, NE1 7RU, UK

Graham Purvis, Archie Crosskey, Jupiter Vincent, Corinne Wills, Cijo Xavier & Jon Telling

NEXUS, School of Engineering, Newcastle University, Newcastle upon Tyne, NE1 7RU, UK

Lidija Šiller & Jake Sheriff

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

J.T. conceived and designed the AHV simulator and project and helped shape the wider implications of the work. G.P. led AHV simulator assembly and experimental operation, prepared samples for XPS, FTIR and TG-DSC-QMS, conducted analysis on FTIR, XPS and TD-GC/MS data and wrote the manuscript and prepared the figures. L.S. and J.S. carried out XPS analysis and L.S. advised with XPS analysis and data interpretation. C.X. conducted TG-DSC-QMS and XRD analysis. C.W. advised and conducted 1H NMR analysis. A.C. assisted with AHV simulator assembly and conducted experimental operation, prepared samples for XPS, FTIR and TG-DSC-QMS and conducted GC/MS and TD-GC/MS analysis. J.V. assisted with AHV simulator assembly, prepared samples for TD-GC/MS, TG-DSC-QMS and XRD contributed the interpretation of FTIR data. All authors contributed to writing the paper and the final manuscript was read and accepted by all contributors.

Correspondence to Jon Telling.

The authors declare no competing interests.

Communications Earth & Environment thanks Eloi Camprubi-Casas and Nick Lane for their contribution to the peer review of this work. Primary Handling Editors: Mojtaba Fakhraee, Joe Aslin and Clare Davis. A peer review file is available.

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

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

Purvis, G., Šiller, L., Crosskey, A. et al. Generation of long-chain fatty acids by hydrogen-driven bicarbonate reduction in ancient alkaline hydrothermal vents. Commun Earth Environ 5, 30 (2024). https://doi.org/10.1038/s43247-023-01196-4

Download citation

Received: 24 July 2023

Accepted: 21 December 2023

Published: 10 January 2024

DOI: https://doi.org/10.1038/s43247-023-01196-4

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Nature (2024)

Communications Earth & Environment (2024)

SHARE