Biogeosciences, 15, 1535–1548, 2018 https://doi.org/10.5194/bg-15-1535-2018 © Author(s) 2018. This work is distributed under the Creative Commons Attribution 4.0 License. Ideas and perspectives: hydrothermally driven redistribution and sequestration of early Archaean biomass – the “hydrothermal pump hypothesis” Jan-Peter Duda1,2, Volker Thiel1, Thorsten Bauersachs3, Helge Mißbach1,4, Manuel Reinhardt1,4, Nadine Schäfer1,2, Martin J. Van Kranendonk5,6,7, and Joachim Reitner1,2 1Department of Geobiology, Geoscience Centre, Georg-August-Universität Göttingen, Goldschmidtstraße 3, 37077 Göttingen, Germany 2“Origin of Life” Group, Göttingen Academy of Sciences and Humanities, Theaterstraße 7, 37073 Göttingen, Germany 3Department of Organic Geochemistry, Institute of Geosciences, Christian-Albrechts-Universität Kiel, Ludewig-Meyn-Straße 10, 24118 Kiel, Germany 4Department Planets and Comets, Max Planck Institute for Solar System Research, Justus-von-Liebig-Weg 3, 37077 Göttingen, Germany 5Australian Centre for Astrobiology, University of New South Wales, Kensington, New South Wales 2052, Australia 6School of Biological, Earth and Environmental Sciences, University of New South Wales, Kensington, New South Wales 2052, Australia 7Australian Research Council Centre of Excellence for Core to Crust Fluid Systems, School of Biological, Earth and Environmental Sciences, University of New South Wales, Kensington, New South Wales 2052, Australia Correspondence: Jan-Peter Duda (jan-peter.duda@geo.uni-goettingen.de) Received: 1 December 2017 – Discussion started: 5 December 2017 Revised: 6 February 2018 – Accepted: 8 February 2018 – Published: 15 March 2018 Abstract. Archaean hydrothermal chert veins commonly contain abundant organic carbon of uncertain origin (abi- otic vs. biotic). In this study, we analysed kerogen contained in a hydrothermal chert vein from the ca. 3.5 Ga Dresser Formation (Pilbara Craton, Western Australia). Catalytic hy- dropyrolysis (HyPy) of this kerogen yielded n-alkanes up to n-C22, with a sharp decrease in abundance beyond n-C18. This distribution (≤ n-C18) is very similar to that observed in HyPy products of recent bacterial biomass, which was used as reference material, whereas it differs markedly from the unimodal distribution of abiotic compounds experimen- tally formed via Fischer–Tropsch-type synthesis. We there- fore propose that the organic matter in the Archaean chert veins has a primarily microbial origin. The microbially de- rived organic matter accumulated in anoxic aquatic (sur- face and/or subsurface) environments and was then assimi- lated, redistributed and sequestered by the hydrothermal flu- ids (“hydrothermal pump hypothesis”). 1 Introduction Extensive hydrothermal chert vein systems containing abun- dant organic carbon are a unique phenomenon of early Ar- chaean successions worldwide (Lindsay et al., 2005; Van Kranendonk, 2006; Hofmann, 2011). A dense stockwork of several hundred kerogen-rich hydrothermal chert veins that penetrate footwall pillowed komatiitic basalts of the ca. 3.5 Ga Dresser Formation (Pilbara Craton, Western Aus- tralia; Fig. 1) are up to 2 km deep by 25 m wide (Hickman, 1973, 1983; Nijman et al., 1999; Van Kranendonk and Pira- jno, 2004; Lindsay et al., 2005; Van Kranendonk, 2006; Van Kranendonk et al., 2008) (Fig. S1 in the Supplement). De- pleted stable carbon isotope signatures (δ13C) of bulk kero- gens (−38.1 to −24.3 ‰) and of organic microstructures (−33.6 to −25.7 ‰) in these hydrothermal chert veins, as well as remnants of what appear to be microbial remains, are consistent with a biological origin of the organic matter (Ueno et al., 2001, 2004; Glikson et al., 2008; Pinti et al., 2009; Morag et al., 2016). Problematically, however, simi- Published by Copernicus Publications on behalf of the European Geosciences Union. 1536 J.-P. Duda et al.: The “hydrothermal pump hypothesis” larly depleted δ13C values (partly down to ca.−36 ‰ relative to the initial substrate) can also be formed through abiotic processes, for instance via Fischer–Tropsch-type synthesis (McCollom et al., 1999; McCollom and Seewald, 2006), and putative microbial remains are not always reliable (Schopf, 1993; Brasier et al., 2002, 2005; Schopf et al., 2002; Bower et al., 2016). Organic biomarkers can help to trace life and biological processes through deep time and add important information on the origin of the organic matter, even in very old sedi- mentary rocks (Brocks and Summons, 2003; Summons and Hallmann, 2014). In Archaean rocks, however, molecular fin- gerprints in the conventionally analysed bitumen (i.e. the ex- tractable portion of organic matter) are commonly blurred by thermal maturation and/or ancient or modern contamina- tion (Brocks et al., 2008; Gérard et al., 2009; Brocks, 2011; French et al., 2015). In contrast, the non-extractable portion of organic matter, known as kerogen, tends to be less affected by thermal maturation and contamination and is considered to be syngenetic with the host rock (Love et al., 1995; Brocks et al., 2003b; Marshall et al., 2007; Lockhart et al., 2008). Catalytic hydropyrolysis (HyPy) is a powerful tool for sen- sitively releasing kerogen-bound hydrocarbon moieties with little structural alteration (Love et al., 1995). HyPy has been successfully applied to Archaean kerogens in rocks from the Pilbara Craton, liberating syngenetic organic compounds consistent with a biogenic origin (Brocks et al., 2003b; Mar- shall et al., 2007). However, this technique has not yet been used on kerogens contained in hydrothermal chert veins of this age. Here, we present the results of analyses of kerogen em- bedded in a freshly exposed hydrothermal chert vein of the ca. 3.5 Ga Dresser Formation. Our analyses include field and petrographic observations, Raman spectroscopy and organic geochemistry (δ13CTOC; kerogen-bound molecules via HyPy followed by gas chromatography – mass spectrometry (GC- MS) and gas chromatography – combustion – isotope ratio mass spectrometry (GC-C-IRMS)). To further constrain po- tential sources of the Dresser kerogen, we additionally ap- plied HyPy on (i) excessively pre-extracted cyanobacterial biomass and (ii) produced abiotic organic matter via Fischer– Tropsch-type synthesis using a hydrothermal reactor. Results of these investigations suggest that the Dresser kerogen has a primarily microbial origin. We hypothesize that biomass- derived organic compounds accumulated in anoxic aquatic environments and were then redistributed and sequestered by subsurface hydrothermal fluids (“hydrothermal pump hy- pothesis”). 2 Material and methods 2.1 Sample preparation A fresh decimetre-sized sample of a Dresser chert vein was obtained from a recent cut wall of the abandoned Dresser Mine in the Pilbara Craton, Western Australia (GPS: 21◦09′04.13′′ S; 119◦26′15.21′′ E; Figs. 1, S1d; for geologi- cal maps, see Hickman, 1983; Van Kranendonk, 1999; Hick- man and Van Kranendonk, 2012). The external surfaces (ca. 1–2 cm) of the sample block were removed using an acetone-cleaned rock saw and then used for the prepara- tion of thin sections. The surfaces of the resulting inner block were extensively rinsed with acetone and then removed (ca. 1–2 cm) with an acetone-cleaned high precision saw (Buehler; Isomet 1000, Germany). The surfaces of the result- ing sample were again rinsed with acetone and then crushed and powdered using a carefully acetone-cleaned pebble mill (Retsch MM 301, Germany). 2.2 Petrography and Raman spectroscopy Petrographic analysis was conducted using a Zeiss SteREO Discovery.V8 stereomicroscope (transmitted and reflected light) linked to an AxioCam MRc5 5-megapixel camera. Raman spectra were recorded using a Horiba Jobin Yvon LabRam-HR 800 UV spectrometer (focal length of 800 mm) attached to an Olympus BX41 microscope. For excitation an Argon ion laser (Melles Griot IMA 106020B0S) with a laser strength of 20 mW was used. The laser beam was fo- cused onto the sample using an Olympus MPlane 100× ob- jective with a numerical aperture of 0.9 and dispersed by a 600 Lmm−1 grating on a charge-couple device (CCD) de- tector with 1024× 256 pixels. This yielded a spectral reso- lution of < 2 cm−1 per pixel. Data were acquired over 10 to 30 s for a spectral range of 100–4000 cm−1. The spectrom- eter was calibrated by using a silicon standard with a major peak at 520.4 cm−1. All spectra were recorded and processed using the LabSpec™ database (version 5.19.17; Jobin Yvon, Villeneuve d’Ascq, France). 2.3 Raman-derived H /C data H /C values were calculated based on integrated peak intensities of Raman spectra using the formula H /C= 0.871 · ID5/(IG+D2)− 0.0508 (Ferralis et al., 2016). The peaks were fitted in the LabSpec™ software (see Sect. 2.2) using the Gauss/Lorentz function. 2.4 Molecular analysis of the Dresser kerogen All materials used for preparation were heated to 500 ◦C for 3 h and/or extensively rinsed with acetone prior to sample contact. A laboratory blank was prepared and analysed in parallel to monitor laboratory contaminations. Biogeosciences, 15, 1535–1548, 2018 www.biogeosciences.net/15/1535/2018/ J.-P. Duda et al.: The “hydrothermal pump hypothesis” 1537 Figure 1. Location of the study site in Western Australia. The hydrothermal chert vein analysed occurs in a recent cut wall of the abandoned Dresser Mine close to Marble Bar. We applied catalytic hydropyrolysis (HyPy) to release molecules from the Dresser kerogen and pre-extracted biomass of the heterocystous cyanobacterium Anabaena cylindrica SAG 1403-2 following previously published pro- tocols (Snape et al., 1989; Love et al., 1995, 2005). HyPy allows the breaking of covalent bonds by progressive heat- ing under high hydrogen pressure (150 bar). The released products are immediately removed from the hot zone by a constant hydrogen flow and trapped downstream on clean, combusted silica powder cooled with dry ice (Meredith et al., 2004). All hydropyrolysates were eluted from the silica trap with high-purity dichloromethane (DCM), desulfurized overnight using activated copper and subjected to gas chro- matography – mass spectrometry (GC-MS). Before all ex- periments, the empty HyPy system was heated (ambient tem- perature to 520 ◦C, held for 30 min) to remove any residual molecules. Isolation of the Dresser kerogen followed standard proce- dures (cf. Brocks et al., 2003b). 57 g of powdered sample was first demineralized with hydrochloric acid (24 h) and then hy- drofluoric acid (11 days). The purified organic matter was then exhaustively extracted using three excess volumes of DCM and n-hexane, respectively (×3), ultrasonic swelling in pyridine (2× 20 min at 80 ◦C), and ultrasonic extraction with methanol, DCM (×3) and DCM /methanol (1/1; v/v). The kerogen was subsequently extracted with n-hexane un- til no more compounds were detected via GC-MS. The pure kerogen was used for HyPy following an experimental proto- col optimized for the analysis of Archaean kerogens (Brocks et al., 2003b; Marshall et al., 2007). In order to monitor potential contamination, we applied HyPy to blanks before and after the kerogen run. The Dresser kerogen (131.06 mg) and blanks were sequentially heated in the presence of a sulfided molybdenum catalyst (10 wt %) and under a constant hydrogen flow (5 dm3 min−1) using a two-step approach (cf. Brocks et al., 2003b; Marshall et al., 2007; Fig. S2). The low-temperature step included heat- ing from ambient temperature to 250 ◦C (300 ◦Cmin−1) and then to 330 ◦C (8 ◦Cmin−1) to release any molecules that were strongly adsorbed to the kerogen and not accessible to solvent extraction. The silica powder was subsequently recovered from the product trap, and the trap was refilled with clean, combusted silica powder for the following high- temperature step. This step included heating from ambient temperature to 520 ◦C (8 ◦Cmin−1) to release molecules co- valently bound to the kerogen. 2.5 Molecular analysis of pre-extracted cyanobacterial biomass (Anabaena cylindrica) We used cell material of the heterocystous cyanobacterium Anabaena cylindrica strain SAG 1403-2 as it fulfils all criteria for reference material (availability, well character- ized, etc.). The material was obtained from the Culture Collection of Algae at the Georg-August-Universität Göt- tingen (Germany) and grown at the Christian-Albrechts- Universität Kiel (Germany). The axenic batch culture was maintained in 250 mL of BG110 medium free of com- bined nitrogen sources (Rippka et al., 1979) at 29 ◦C. A light : dark regime of 14 h : 10 h with a photon flux density of 135 µmol photonsm−2 s−1 was provided by a white flu- orescent light bulb. At the end of the logarithmic growth phase, cells were harvested by centrifugation, and they were lyophilized thereafter. Subsequently, an aliquot of the freeze-dried biomass was exhaustively extracted following the methodology described by Bauersachs et al. (2014). HyPy was performed following an experimental proto- col optimized for biomass applications (Love et al., 2005). Briefly, the material was heated in the presence of a sul- fided molybdenum catalyst (1.5 times the weight of the bac- terial extraction residue) and under a constant hydrogen flow (6 dm3 min−1). The HyPy program included heating from ambient temperature to 260 ◦C (300 ◦Cmin−1) and then to 500 ◦C (8 ◦Cmin−1). 2.6 Fischer–Tropsch-type synthesis of organic matter under hydrothermal conditions Fischer–Tropsch-type reactions were carried out based on McCollom et al. (1999). A mixture of 2.5 g oxalic acid (dihy- drate, suprapur®, Merck KGaA), 1 g montmorillonite (K10, Sigma Aldrich; pre-extracted with DCM; ×4) and ca. 11 mL ultrapure water was heated to 175 ◦C for 66 to 74 h in a sealed Morey-type stainless steel autoclave. The autoclave www.biogeosciences.net/15/1535/2018/ Biogeosciences, 15, 1535–1548, 2018 1538 J.-P. Duda et al.: The “hydrothermal pump hypothesis” was rapidly (≤ 15 min) cooled to room temperature with compressed air. Fluid and solid phases were collected and extracted with DCM (×4). The montmorillonite was removed by centrifu- gation and filtration with silica powder and sea sand (which were combusted prior to use). The obtained extracts were then concentrated by rotary evaporation (40 ◦C, 670 mbar) and subjected to GC-MS analysis. Analytical blank exper- iments were carried out with all reactants to keep track of contamination. 2.7 Gas chromatography – mass spectrometry GC-MS analysis of HyPy products was carried out with a Thermo Scientific Trace 1310 GC coupled to a Thermo Sci- entific Quantum XLS Ultra MS. The GC instrument was equipped with a capillary column (Phenomenex Zebron ZB- 5, 30 m, 0.25 µm film thickness, 0.25 mm inner diameter). Fractions were injected into a splitless injector and trans- ferred to the GC column at 270 ◦C. He was used as carrier gas with a constant flow rate of 1.5 mLmin−1. The GC oven temperature was held isothermal at 80 ◦C for 1 min and then ramped to 310 ◦C at 5 ◦Cmin−1, at which it was kept for 20 min. Electron ionization mass spectra were recorded in full-scan mode at an electron energy of 70 eV with a mass range of m/z 50–600 and scan time of 0.42 s. 2.8 Polyaromatic hydrocarbon ratios Polyaromatic hydrocarbons (PAHs) are organic compounds consisting of multiple fused benzene rings. Methylation and isomerization characteristics of various GC-amenable PAHs are altered during maturation and thus provide a measure for assessing the thermal maturity of organic matter (Killops and Killops, 2005; Peters et al., 2005). The following PAH matu- rity parameters were used in this study: 1. methylnaphthalene ratio (MNR)= 2-MN / 1-MN (Radke et al., 1984), 2. methylphenanthrene index (MPI-I)= 1.5 · (2-MP+ 3- MP) / (P+ 1-MP+ 9-MP) (Radke and Welte, 1983), and 3. computed vitrinite reflectance [Rc (MPI-I)]= 0.7 ·MPI- 1+ 0.22 (according to P /MP> 1; Boreham et al., 1988). MN stands for methylnaphthalene, P for phenanthrene and MP for methylphenanthrene. 2.9 Total organic carbon (TOC) and δ13C analyses (TOC and compound specific) TOC, δ13CTOC and compound-specific δ13C analyses were conducted at the Centre for Stable Isotope Research and Analysis (KOSI) at the Georg-August-Universität Göttingen, Germany. Stable carbon isotope data are expressed as delta values relative to the Vienna Pee Dee Belemnite (VPDB) ref- erence standard. The TOC content and δ13CTOC values were determined in duplicate using an elemental analyser (NA-2500 CE- Instruments) coupled to an isotope ratio mass spectrometer (Finnigan MAT Delta plus). Ca. 100 mg of powdered and homogenized whole rock material were analysed in each run. For internal calibration an acetanilide standard was used (δ13C=−29.6 ‰; SD= 0.1 ‰). TOC content measurements showed a mean deviation of 0.1 wt%. The average δ13CTOC value had a standard deviation of 0.3 ‰. Compound-specific δ13C analyses were conducted with a Trace GC coupled to a Delta Plus isotope ratio mass spec- trometer (IRMS) via a combustion interface (all Thermo Sci- entific). The combustion reactor contained CuO, Ni and Pt and was operated at 940 ◦C. The GC was equipped with two serially coupled capillary columns (Agilent DB-5 and DB- 1; each 30 m, 0.25 µm film thickness, 0.25 mm inner diame- ter). Fractions were injected into a splitless injector and trans- ferred to the GC column at 290 ◦C. The carrier gas was He at a flow rate of 1.2 mLmin−1. The GC oven temperature pro- gram was identical to the one used for GC-MS analysis (see above). CO2 with a known δ13C value was used for inter- nal calibration. Instrument precision was checked using lab- oratory standards. Standard deviations of duplicate measure- ments were better than 1.7 ‰. 3 Results The studied hydrothermal chert vein is hosted in komatiitic pillow basalt that has undergone severe hydrothermal acid– sulfate alteration, producing a kaolinite–illite–quartz mineral assemblage (Van Kranendonk and Pirajno, 2004; Van Kra- nendonk, 2006) (Fig. S1). The sampled vein crops out in a recent cut wall of the abandoned Dresser Mine (Fig. S1) and consists of a dense chert (microquartz) matrix of deep black colour that contains kerogen and local concentrations of fresh (unweathered) pyrite crystals (Fig. 2a, b). There is no field and/or petrographic evidence for fluid-flow events that post- date the initial vein formation (brecciation textures, etc.). Petrographic analysis and Raman spectroscopy revealed that kerogen (D bands at 1353 cm−1, G bands at 1602 cm−1) is embedded in the chert matrix (SiO2 band at 464 cm−1) (see Fig. 2c, for a representative Raman spectrum). The organic matter occurs as small clots (< 50 µm) of variable shape. Ra- man spectra of the organic matter exhibit relatively wide D and G bands (82 and 55 cm−1 width, respectively) (Fig. 2c). The total organic carbon (TOC) content is 0.2 wt%, and the δ13CTOC value is −32.8± 0.3 ‰ (Table 1). Raman-derived H /C ratios range between 0.03 and 0.14. HyPy was applied to the isolated Dresser kerogen, as well as to preceding and subsequent analytical blanks (combusted sea sand). Hydropyrolysates of the preceding blank con- Biogeosciences, 15, 1535–1548, 2018 www.biogeosciences.net/15/1535/2018/ J.-P. Duda et al.: The “hydrothermal pump hypothesis” 1539 Table 1. Stable carbon isotopic composition (δ13C) of the total organic carbon (TOC) and n-alkanes released from high-temperature HyPy of the Dresser kerogen. n-C12 n-C13 n-C14 n-C15 n-C16 n-C17 n-C18 n-C12−18 TOC (mean) δ13C −30.3 −33.3 −32.7 −31.1 −29.4 −31.2 −31.7 −31.4 −32.8 SD 0.5 1.4 0.2 1.7 0.3 0.8 0.1 1.2 0.3 Table 2. Maturity indices (based on aromatic hydrocarbons) of the Dresser kerogen. MP/P: methylphenathrene/phenanthrene ratio; MPI-I: methylphenanthrene index (1.5 · (2-MP+ 3- MP)/(P+ 1-MP+ 9-MP); Radke and Welte, 1983); P/MP: phenanthrene/methylphenathrene ratio; Rc (MPI-I): computed vitrinite reflectance (0.7 ·MPI-1+ 0.22, according to P/MP> 1; Boreham et al., 1988); MNR: methylnaphthalene ratio (2-MN/1- MN; Radke et al., 1984). MP/P MPI-I P/MP Rc MNR (MPI-I) 0.33 0.23 3.01 2.87 2.52 tained a series of n-alkanes ≤ n-C24, with maximum abun- dances at n-C15 (Figs. 3, S2–S4). Sulfur (only after low- temperature HyPy), traces of siloxanes and a phenol (Figs. 3, S2, S4) were also present. The blank runs also contained traces of aromatic hydrocarbons (Fig. S5). Low-temperature HyPy of the Dresser kerogen produced traces of C14−18 n-alkanes, with a maximum at n-C15 (Fig. S3). However, these compounds were significantly less abundant than those released during high-temperature HyPy (see below). Other compounds observed in the low- temperature step pyrolysate were elemental sulfur, phenols, phthalic acid, siloxanes and traces of aromatic hydrocarbons (Figs. S2, S4–S5). High-temperature HyPy of the Dresser kerogen yielded n-alkanes ranging from n-C11 to n-C22 with a no- table decrease (“step”) in the abundance of homologues above n-C18 (Figs. 3b, S2, S3), which remained vir- tually unaffected by blank subtraction (Fig. 3c). Apart from that step, no carbon number preference is evident. The n-alkanes have δ13C values ranging from −29.4 to −33.3 ‰ (mean −31.4± 1.2 ‰; Fig. 3b; Table 1). The hy- dropyrolysate also contained isomeric mixtures of C12−18 monomethylalkanes (Fig. S3) and a variety of aromatic hydrocarbons, including (dimethyl-, methyl-)naphthalene(s), (methyl-)biphenyl(s), (methyl-)acenaphthene(s), dibenzofu- ran and (methyl-)phenanthrene(s) (Figs. 3b, c, S4–S6). Bi- ologically diagnostic hydrocarbons such as hopanoids or steroids were absent. HyPy treatment of excessively pre-extracted biomass of the heterocystous cyanobacterium Anabaena cylindrica SAG 1403-2 yielded a variety of organic compounds, but also in- cluded n-alkanes with a clear restriction in carbon number to homologues ≤ n-C18 (Fig. 3d). In contrast, our experimen- tal synthesis of abiotic n-alkanes through Fischer–Tropsch- type reactions under hydrothermal conditions produced a unimodal distribution of homologues ranging from n-C12 to n-C41 without any carbon number preference (Fig. 3e). 4 Discussion 4.1 Maturity of the Dresser kerogen The organic record of Archaean rocks is commonly af- fected by thermal maturation (Brocks et al., 2008; Brocks, 2011; French et al., 2015). The kerogen within the anal- ysed Dresser sample is thermally mature and structurally disordered, as evidenced by relatively wide D and G bands in the Raman spectra (see Fig. 2c for a representative Ra- man spectrum). This is also supported by the absence of S1 bands at 2450 cm−1, high R1 ratios (0.98–1.05) and low FWHM-D1 values (68.64–76.12), consistent with prehnite– pumpellyite to lower greenschist metamorphism at a temper- ature of ca. 300 ◦C (cf. Yui et al., 1996; Delarue et al., 2016). All of these observations are well in line with published Ra- man spectra of Archaean organic matter from the same re- gion (Ueno et al., 2001; Marshall et al., 2007; Delarue et al., 2016) and the general thermal history of the host rock (regional prehnite–pumpellyite to lower greenschist meta- morphism; Hickman, 1975, 1983, 2012; Terabayashi et al., 2003). Fitting of D5 peaks can be difficult for Raman spec- tra of highly mature organic matter (Ferralis et al., 2016). Therefore, the H /C ratios calculated herein (0.03–0.14) should be treated with caution. However, the Raman-based H /C ratios and the methylphenanthrene / phenanthrene value (MP /P= 0.33; Table 2) of the Dresser kerogen are in good accordance with data reported from more mature kerogens from the same region (ca. 3.4 Ga Strelley Pool Formation: 0.08–0.14 and 0.24–0.37, respectively; Marshall et al., 2007). The low methylphenanthrene index (MPI- I= 0.23) and high phenanthrene /methylphenanthrene index (P /MP= 3.01) result in a computed vitrinite reflectance (Rc (MPI-I)) of 2.87 (Table 2), indicating a thermal maturity far beyond the oil-generative stage (Radke and Welte, 1983; Boreham et al., 1988). However, it has to be considered that the MPI-I is potentially affected by (de-)methylation reac- www.biogeosciences.net/15/1535/2018/ Biogeosciences, 15, 1535–1548, 2018 1540 J.-P. Duda et al.: The “hydrothermal pump hypothesis” Figure 2. Petrographic observations on the hydrothermal chert vein. (a, b) Thin section photographs (a transmitted light; b reflected light) showing kerogen (brownish colours) and pyrite (arrows, black colours in a, bright colours in b) embedded within a fine-grained chert matrix. Note that the pyrite crystals (arrows) are excellently preserved and show no evidence of oxidation. (c) Representative Raman spectrum of kerogen (D band at 1353 cm−1 and G band at 1602 cm−1) present in the hydrothermal chert vein. Note the wide D and G bands (82 and 55 cm−1 width, respectively), pointing to ther- mally mature and structurally disordered kerogen, and the absence of an S1 band (at 2450 cm−1), consistent with prehnite–pumpellyite to lower greenschist metamorphosis and a temperature of ca. 300 ◦C (cf. Yui et al., 1996; Delarue et al., 2016). tions (Brocks et al., 2003a). The calculated methylnaphtha- lene ratio (MNR= 2.52; Table 2) corresponds to a somewhat lower mean vitrinite reflectance of ca. 1.5, which is again in line with a post-oil window maturity (cf. Radke et al., 1984). The mismatches between single aromatic maturity pa- rameters are negligible, as these indices have limited appli- cation for highly mature Archaean organic matter (Brocks et al., 2003a). The putative offset between the indices and the metamorphic overprint indicated by Raman data is most likely due to the protection of kerogen-bound moieties even under elevated thermal stress (Love et al., 1995; Lockhart et al., 2008). Furthermore, it has been shown that kerogen iso- lated from the Strelley Pool Formation also contains larger PAH clusters which are not GC-amenable (> 10–15 rings; Marshall et al., 2007). Consequently, it can be anticipated that the compounds detected in the HyPy pyrolysate of the Dresser kerogen represent only a small fraction of the bulk macromolecular organic matter. The distribution of monomethylalkanes≤ n-C18 (Figs. S3, S4) released from the Dresser kerogen resembles high- temperature HyPy products from the Strelley Pool kerogen (Marshall et al., 2007; their Fig. 15). Such isomeric mixtures are typically formed during thermal cracking of alkyl moi- eties (Kissin, 1987) and are in good agreement with the esti- mated maturity range. Methylated aromatics such as methyl- naphthalenes and methylphenanthrenes have also been ob- served in other hydropyrolysates from Archaean kerogens that experienced low-grade metamorphism (Brocks et al., 2003b; Marshall et al., 2007; French et al., 2015). In all of these cases, the degree of alkylation varied with the exact thermal alteration of the respective kerogens (Marshall et al., 2007; French et al., 2015). 4.2 Syngeneity of the Dresser kerogen-derived compounds The kerogen of the Dresser Formation exclusively occurs in the form of fluffy aggregates and clots embedded within a very dense chert matrix that is, once solidified, highly im- permeable to fluids. The depositional age of the formation is constrained to 3481±3.6 Ma (Van Kranendonk et al., 2008), and the investigated chert vein shows no evidence for dis- ruption by post-depositional hydrothermal fluids. This has also been described for other hydrothermal chert veins of the Dresser Formation, where the kerogen has been interpreted as being syngenetic (i.e. formed prior to host rock lithifica- tion; Ueno et al., 2001, 2004; Morag et al., 2016). Further- more, the maturity of the embedded kerogen is in good ac- cordance with the thermal history of the host rock. An intro- duction of solid macromolecular organic matter from strati- graphically younger units in this region during later fluid- flow phases, as proposed for the younger Apex chert (Olcott- Marshall et al., 2014), can therefore be excluded. As the bitumen fractions of Precambrian rocks are eas- ily biased by the incorporation of contaminants during later Biogeosciences, 15, 1535–1548, 2018 www.biogeosciences.net/15/1535/2018/ J.-P. Duda et al.: The “hydrothermal pump hypothesis” 1541 Figure 3. Total ion currents (a–c; on the same scale) and ion chromatograms selective for alkanes (d–e; m/z 85). (a) High-temperature HyPy (330–520 ◦C) product of analytical blank (combusted sea sand) obtained prior to HyPy of the Dresser kerogen. n-Alkanes in the range of n-C14 to n-C24 (maxima at n-C15) represent HyPy background contamination. (b) High-temperature HyPy (330–520 ◦C) product of the Dresser kerogen. Note the sharp decrease in abundance of n-alkanes beyond n-C18 (see arrows). δ13C values of n-C12 to n-C28 show a high similarity to the δ13CTOC value (−32.8± 0.3 ‰), further confirming syngeneity. (c) Blank subtraction (b minus a) showing that contaminants have no major impact on the n-alkane pattern yielded during the high-temperature HyPy step of the Dresser kerogen (b). (d) HyPy products of cell material of the heterocystous cyanobacterium Anabaena cylindrica. Note the sharp decrease in abundance of n-alkanes beyond n-C18 (see arrows), similar to the n-alkane distribution of the Dresser kerogen (b). (e) Products of experimental Fischer– Tropsch-type synthesis under hydrothermal conditions; abiogenic n-alkanes show a unimodal distribution that is distinctly different from the Dresser kerogen. Black dots: n-alkanes (numbers refer to carbon chain lengths); N: naphthalene; MN: methylnaphthalenes; BiPh: 1,1′- biphenyl; DMN: dimethylnaphthalenes; DBF: dibenzofuran; MAN: methylacenaphthene; P: phenanthrene; crosses: siloxanes (GC column or septum bleeding). www.biogeosciences.net/15/1535/2018/ Biogeosciences, 15, 1535–1548, 2018 1542 J.-P. Duda et al.: The “hydrothermal pump hypothesis” stages of rock history (Brocks et al., 2008; Brocks, 2011; French et al., 2015), studies have increasingly focussed on kerogen-bound compounds that are more likely to be syn- genetic to the host rock (Love et al., 1995; Brocks et al., 2003b; Marshall et al., 2007; French et al., 2015). Potential volatile organic contaminants adhering to the kerogen are re- moved through excessive extraction and a thermal desorption step (∼ 350 ◦C) prior to high-temperature HyPy (550 ◦C; cf. Brocks et al., 2003b; Marshall et al., 2007). The recur- rence of few n-alkanes in the range of n-C14 to n-C24 (max- imum at n-C15) in high-temperature HyPy blank runs ob- tained immediately before and after the actual sample run (Figs. 3a, S2, S3) indicates minor background contamination during pyrolysis, with a source most likely within the HyPy system. However, these contaminants do not significantly af- fect the n-alkane pattern yielded by high-temperature HyPy of the Dresser kerogen, as evidenced by blank subtraction (Fig. 3c). Contamination of the sample can be further deciphered by the presence of polar additives or hydrocarbons that are not consistent with the thermal history of the host rock. Plastic- derived branched alkanes with quaternary carbon centres (BAQCs), a common contaminant in Precambrian rock sam- ples (Brocks et al., 2008), have not been detected (Fig. S7). Traces of functionalized plasticizers (phenols and phthalic acid; Figs. S2, S6, S8) are unlikely to survive (or result from) HyPy treatment. These compounds are therefore con- sidered as background contamination introduced during sam- ple preparation and analysis after HyPy. The observed silox- anes (Fig. 3, S2) probably originate from the GC column or septum bleeding and are unlikely to be contained in the sample. All of these compounds occur only in low or trace abundances and can be clearly distinguished from the ancient aliphatic and aromatic hydrocarbons contained in the Dresser kerogen. Contamination by (sub-)recent endoliths can be excluded as sample surfaces have been carefully removed and there is no petrographic indication for borings or fissures con- taining recent organic material (Fig. 2). HyPy of untreated or extracted biomass would yield a variety of acyclic and cyclic biomarkers (cf. Love et al., 2005). However, high- temperature HyPy of the Dresser kerogen almost exclusively yielded n-alkanes, minor amounts of monomethylalkanes and various aromatic hydrocarbons (Figs. 3b, c, S2–S6), while hopanoids or steroids were absent (Figs. S8, S9). This is in good agreement with the maturity of the Dresser kero- gen and previous HyPy studies of Archaean rocks (Brocks et al., 2003b; Marshall et al., 2007; French et al., 2015). Accidental contamination of the kerogen by mono-, di- and triglycerides (e.g. dust, skin surface lipids) can also be ruled out as HyPy treatment of these compounds typ- ically results in n-alkane distributions with a distinct pre- dominance of n-C16 and n-C18 homologues, corresponding to the n-C16 and n-C18 fatty acid precursors (Craig et al., 2004; Love et al., 2005; unpublished data from our own ob- servations). At the same time, the observed distribution of kerogen-derived n-alkanes, with a distinct decrease beyond n-C18 (Fig. 3b, c), is notably similar to high-temperature HyPy products of kerogens isolated from the ca. 3.4 Ga Strel- ley Pool Formation of the Pilbara Craton (Marshall et al., 2007; their Fig. 14). Marshall and co-workers considered these n-alkanes unlikely to be contaminants as they were released only in the high-temperature HyPy step. Further- more, the stable carbon isotopic composition of n-alkanes in the Dresser high-temperature pyrolysate (−29.4 to−33.3 ‰; mean −31.4± 1.2 ‰) is very similar to the δ13CTOC signa- ture of the sample (−32.8± 0.3 ‰; Figs. 3, S10; Table 1), indicating that these compounds were generated from the kerogen. Hence, we consider the compounds released from the Dresser kerogen during high-temperature HyPy (Fig. 3b, c) as syngenetic. High-temperature HyPy of the Dresser kerogen yielded a variety of aromatic hydrocarbons, which are orders of mag- nitudes lower or absent in all other pyrolysates (Figs. 3, S2, S4–S6). It also produced significantly higher amounts of n-alkanes than the low-temperature step, and these fur- ther showed a distinct distribution pattern (i.e. a step ≤ n- C18; Figs. 3, S2, S3). The lack of even low quantities of n-alkenes that typically accompany bond cleavage in kero- gen pyrolysis was also observed by Marshall et al. (2007) in their HyPy analysis of cherts from the Strelley Pool For- mation. As a possible explanation for the lack of n-alkenes, these authors suggested that the n-alkanes were not cracked from the kerogen but were rather trapped in closed microp- ores until the organic host matrix was pyrolytically disrupted. Alternatively, however, unsaturated cleavage products may have been immediately reduced during HyPy by the steadily available hydrogen and catalysts. Indeed, it has been demon- strated that double bonds in linear alkyl chains are efficiently hydrogenated during HyPy, even in the case of pre-extracted microbial biomass (e.g. Love et al., 2005; our Fig. 3d). We consider both as plausible scenarios to explain the absence of n-alkenes in the Dresser chert hydropyrolysate. 4.3 Origin of the Dresser kerogen: hydrothermal vs. biological origin The δ13CTOC value of −32.8± 0.3 ‰ is consistent with car- bon fixation by photo- or chemoautotrophs (cf. Schidlowski, 2001). However, organic compounds exhibiting similar 13C depletions (partly down to ca. −36 ‰ relative to the initial substrate) could also be formed abiotically during the serpen- tinization of ultramafic rocks (Fischer–Tropsch-type synthe- sis; McCollom et al., 1999; McCollom and Seewald, 2006; Proskurowski et al., 2008). A further constraint on the origin of the Dresser kerogen is provided by the distinct decrease in n-alkane abundance beyond n-C18 observed in the high- temperature HyPy pyrolysate (Fig. 3b, c). This distribution resembles HyPy products of pre-extracted recent cyanobac- Biogeosciences, 15, 1535–1548, 2018 www.biogeosciences.net/15/1535/2018/ J.-P. Duda et al.: The “hydrothermal pump hypothesis” 1543 terial biomass, which also shows a very pronounced restric- tion in carbon number to homologues ≤ n-C18 (Fig. 3d). The pre-extracted cyanobacterial cell material and the abi- otically produced n-alkanes studied as reference samples ex- perienced no thermal maturation. It can nevertheless be ex- pected that burial, and thus heating, of Anabaena biomass (Fig. 3d) would initially liberate lower n-alkane homologues from their predominant n-alkyl moieties while, up to a cer- tain point, retaining the distinct step at n-C18. Experimental maturation of an immature kerogen revealed the preservation of distinct alkyl-chain length preferences even after 100 days at 300 ◦C (Mißbach et al., 2016; e.g. a step beyond n-C31 in their Fig. 2). Further maturation would ultimately lead to a unimodal distribution of short-chain n-alkanes and erase the biologically inherited pattern (cf. Mißbach et al., 2016). In contrast, abiotically synthesized extractable organic com- pounds show a unimodal homologue distribution from the beginning (Fig. 3e) and will retain it, while thermal mat- uration would gradually shift the n-alkane pattern towards shorter homologues. It can therefore be expected that or- ganic compounds cleaved from an abiotic “Fischer–Tropsch- kerogen” – whose existence has not been proven yet – would also exhibit a unimodal distribution. Consequently, the dis- tinctive distribution of n-alkanes released from the Dresser kerogen (i.e. the step ≤ n-C18; Figs. 3b, c) can be regarded as a molecular fingerprint relating to a biosynthetic origin of the organic matter. Highly 13C-depleted methane in primary fluid inclu- sions in hydrothermal chert veins of the Dresser Forma- tion (δ13C< 56 ‰) was taken as evidence for biological methanogenesis and thus the presence of Archaea (Ueno et al., 2006). Our δ13CTOC value (−32.8± 0.3 ‰; Table 1; Fig. S10) would generally be consistent with both bacte- rial and archaeal sources (cf. Schidlowski, 2001). Archaea only synthesize isoprene-based compounds (Koga and Morii, 2007; Matsumi et al., 2011). Straight-chain (acetyl-based) hydrocarbon moieties such as fatty acids – the potential precursors of the kerogen-derived n-alkanes – are to cur- rent knowledge being formed only by Bacteria and Eukarya, where they typically function as constituents of membranes or storage lipids (cf. Erwin, 1973; Kaneda, 1991). The forma- tion of these lipids is tightly controlled by different biosyn- thetic pathways resulting in characteristic chain-length dis- tributions. In bacterial lipids, carbon chain lengths typically do not extend above n-C18 (cf. Kaneda, 1991). Consequently, given the lack of convincing evidence for the presence of Eu- karya as early as 3.5 Ga (cf. Parfrey et al., 2011; Knoll, 2014; French et al., 2015), the most plausible source of kerogen- occluded n-alkanes in the Dresser hydrothermal chert is Bac- teria. Strongly reducing conditions during the deposition of the Dresser Formation are indicated by widespread pyrite, Fe-rich carbonates and trace element signatures (Van Kra- nendonk et al., 2003, 2008). Potential microbial sources for the Dresser kerogen therefore may have included anoxygenic photoautotrophic, chemoautotrophic and heterotrophic mi- croorganisms. 4.4 The “hydrothermal pump hypothesis” Our results strongly support a biological origin of the kero- gen found in the early Archaean hydrothermal chert veins of the Dresser Formation. We explain this finding by the redistribution and sequestration of microbial organic matter that may have formed in a variety of Dresser environments through hydrothermal circulation (proposed herein as the “hydrothermal pump hypothesis”; Fig. 4). Higher geothermal gradients prior to the onset of modern-type plate tectonics at ca. 3.2–3.0 Ga (Smithies et al., 2005; Shirey and Richard- son, 2011) were possibly important drivers of early Archaean hydrothermal systems. In fact, the Dresser Formation was formed in a volcanic caldera environment affected by strong hydrothermal circulation, where voluminous fluid circulation locally caused intense acid–sulfate alteration of basalts and the formation of a dense hydrothermal vein swarm (Nijman et al., 1999; Van Kranendonk and Pirajno, 2004; Van Kra- nendonk, 2006; Van Kranendonk et al., 2008; Harris et al., 2009). In addition to the high crustal heat flow, the absence of thick sedimentary cover may have facilitated the intrusion of seawater into the hydrothermal system. The associated large- scale assimilation of particulate and dissolved organic mat- ter and its transport and alteration by hydrothermal fluids (Fig. 4) therefore appears to be a plausible mechanism that may, at least partly, explain the high amounts of kerogen in early Archaean hydrothermal veins. The hydrothermal pump hypothesis requires a source of organic matter during the deposition of the Dresser For- mation (Fig. 4). Whereas contributions from extraterres- trial sources, as well as from Fischer–Tropsch-type synthe- sis linked to the serpentinization of ultramafic rocks, cannot be excluded, our results indicate a primarily biological origin for the kerogen contained in the chert veins (Fig. 4). The in- ferred biogenicity is also in line with the consistent δ13C off- set between bulk kerogens (ca.−20 to−30 ‰) and carbonate (ca. ±2 ‰) in Archaean rocks (Hayes, 1983; Schidlowski, 2001). Prokaryotic primary producers and heterotrophs may have flourished in microbial mats (Dresser stromatolites; Walter et al., 1980; Van Kranendonk, 2006, 2011; Philippot et al., 2007; Van Kranendonk et al., 2008), the water column (planktic “marine snow”; Brasier et al., 2006; Blake et al., 2010) and even hot springs on land (Djokic et al., 2017). An- other biological source for the ancient organic matter could have been chemoautotrophs and heterotrophs thriving in sub- surface environments, such as basalts (Banerjee et al., 2007; Furnes et al., 2008) and hydrothermal vent systems (Shen et al., 2001; Ueno et al., 2001, 2004, 2006; Pinti et al., 2009; Morag et al., 2016) (Fig. 4). All of these systems are not mu- tually exclusive and the largely anoxic conditions would have encouraged a high steady-state abundance of organic matter in the aquatic environment (Fig. 4). www.biogeosciences.net/15/1535/2018/ Biogeosciences, 15, 1535–1548, 2018 1544 J.-P. Duda et al.: The “hydrothermal pump hypothesis” Figure 4. The “hydrothermal pump hypothesis”. Organic matter was predominantly biologically produced and heterotrophically processed by Bacteria and, possibly, Archaea. Additionally, Fischer–Tropsch-type synthesis of organic matter linked to the serpentinization of ultramafic rocks (McCollom et al., 1999; McCollom and Seewald, 2006) may have occurred locally. Primary producers (chemoautotrophs, anoxygenic photoautotrophs) and heterotrophs may have flourished in surface waters (planktic “marine snow”), at the water–rock interface (microbial mats and/or biofilms) and in cryptic environments (e.g. within basalts and hydrothermal vent systems). After accumulating in different Dresser environments, the organic matter was redistributed and sequestered in veins by hydrothermal fluids. Dissolved organic matter (DOM) in modern seawater may resist decomposition over millennial timescales (Druffel and Griffin, 2015). In recent hydrothermal fields, however, or- ganic matter becomes thermally altered and redistributed (Si- moneit, 1993; Delacour et al., 2008; Konn et al., 2009). Lab- oratory experiments using marine DOM indicate that ther- mal alteration already occurs at temperatures > 68–100 ◦C and efficient removal of organic molecules at 212–401 ◦C (Hawkes et al., 2015, 2016). It has been argued, however, that such DOM removal may also be due to transformation into immiscible material through, for example, condensation (Castello et al., 2014) and/or defunctionalization reactions (Hawkes et al., 2016). These processes, however, are as yet poorly understood. In the Dresser Formation, hydrothermal temperatures ranged from ca. 300 ◦C at depth to 120 ◦C near the palaeosurface, causing propylitic (ca. 250–350 ◦C) and argillic (including advanced argillic: ca. 100–200 ◦C) alter- ation of the host rocks (Van Kranendonk and Pirajno, 2004; Van Kranendonk et al., 2008; Harris et al., 2009). Given this variety of thermal regimes, and the generally anoxic nature of early Archaean seawater (e.g. Van Kranendonk et al., 2003, 2008; Li et al., 2013), it is likely that some of the organic sub- stances underwent in situ alteration, but not complete oxida- tion, during hydrothermal circulation. The entrained organ- ics would have been trapped in the chert that instantaneously precipitated from the ascending hydrothermal fluids due to subsurface cooling (cf. Van Kranendonk, 2006). In summary, the hydrothermal pump hypothesis proposed here (Fig. 4) includes (i) a net build-up of organic matter in different Dresser environments under largely anoxic condi- tions, (ii) a large-scale assimilation of particulate and dis- solved organic matter from various biological sources and its subsurface transport and alteration by hydrothermal fluids, and (iii) its sequestration within hydrothermal chert veins as kerogen. This model explains the presence of abundant or- ganic carbon in early Archaean hydrothermal veins, as well as its morphological, structural and isotopic variability ob- served in the Dresser hydrothermal chert veins (Ueno et al., 2001, 2004; Pinti et al., 2009; Morag et al., 2016). It does not, however, help to pinpoint the formation pathways of distinct carbonaceous structures, as for instance those preserved in the Dresser Formation (Glikson et al., 2008) or the younger Apex chert (e.g. Schopf, 1993; Brasier et al., 2002, 2005; Schopf et al., 2002; Bower et al., 2016). Further work is nec- essary to test whether consistent molecular and compound- specific isotopic patterns can be generated from a larger set of Archaean kerogens. 5 Conclusions Kerogen embedded in a hydrothermal chert vein from the ca. 3.5 Ga Dresser Formation (Pilbara Craton, Western Aus- tralia) is syngenetic. A biological origin is inferred from the presence of short-chain n-alkanes in high-temperature HyPy pyrolysates, showing a sharp decrease in homologue abundance beyond n-C18. HyPy products of pre-extracted re- cent bacterial biomass exhibited a similar restriction to car- bon chain lengths ≤ n-C18, whereas abiotic compounds ex- perimentally formed via Fischer–Tropsch-type synthesis ex- hibited a unimodal distribution. A biological interpretation for Dresser organics is further consistent with the δ13CTOC value (−32.8± 0.3 ‰) and the stable carbon isotopic com- position of n-alkanes in the Dresser high-temperature py- rolysate (−29.4 to −33.3 ‰; mean −31.4± 1.2 ‰). Based on these observations, we propose that the original organic matter was primarily biologically synthesized. We hypothe- size that microbially derived organic matter accumulating in anoxic aquatic (surface and/or subsurface) environments was Biogeosciences, 15, 1535–1548, 2018 www.biogeosciences.net/15/1535/2018/ J.-P. Duda et al.: The “hydrothermal pump hypothesis” 1545 assimilated, redistributed and sequestered by hydrothermal fluids (“hydrothermal pump hypothesis”). Data availability. Data are available upon request. The Supplement related to this article is available online at https://doi.org/10.5194/bg-15-1535-2018-supplement. Author contributions. JPD, JR and MJVK conducted the field work and designed the study. JR conducted petrographic analyses. NS performed Raman spectroscopic analyses. MR and JPD conducted pyrolysis experiments. HM conducted Fischer–Tropsch-type syn- thesis. TB prepared cyanobacterial cell material. JPD, HM, MR and VT performed biomarker analyses. JPD wrote the manuscript. All authors discussed the results and provided input to the manuscript. Competing interests. The authors declare that they have no conflict of interest. Acknowledgements. This work was financially supported by the Deutsche Forschungsgemeinschaft (grant Du 1450/3-1, DFG Priority Programme 1833 “Building a Habitable Earth”, to Jan-Peter Duda and Joachim Reitner; grant Th 713/11-1 to Volker Thiel), the Courant Research Centre of the Georg-August- Universität Göttingen (DFG, German Excellence Program), the Göttingen Academy of Sciences and Humanities (to Jan-Peter Duda and Joachim Reitner), the International Max Planck Research School for Solar System Science at the Georg-August-Universität Göttingen (to Manuel Reinhardt and Helge Mißbach) and the ARC Centre of Excellence for Core to Crust Fluid Systems (Mar- tin J. Van Kranendonk). We thank Martin Blumenberg, Cornelia Conradt, Wolfgang Dröse, Jens Dyckmans, Axel Hackmann, Merve Öztoprak and Burkhard C. Schmidt for scientific and technical sup- port. Josh Rochelmeier is thanked for assistance during sample ex- traction of A. cylindrica. We are indebted to Malcolm Walter and Mark A. van Zuilen for helpful comments on the manuscript and Jack Middelburg for editorial handling. This is publication num- ber 5 of the Early Life Research Group (Department of Geobiol- ogy, Georg-August-Universität Göttingen; Göttingen Academy of Sciences and Humanities) and contribution 980 from the ARC Cen- tre of Excellence for Core to Crust Fluid Systems. We acknowledge support by the German Research Foundation and the Open Access Publication Funds of Georg-August- Universität Göttingen. This open-access publication was funded by the University of Göttingen. Edited by: Jack Middelburg Reviewed by: Mark van Zuilen and Malcolm Walter References Banerjee, N. R., Simonetti, A., Furnes, H., Muehlenbachs, K., Staudigel, H., Heaman, L., and Van Kranendonk, M. J.: Direct dating of Archean microbial ichnofossils, Geology, 35, 487–490, https://doi.org/10.1130/g23534a.1, 2007. Bauersachs, T., Mudimu, O., Schulz, R., and Schwark, L.: Distribu- tion of long chain heterocyst glycolipids in N2-fixing cyanobac- teria of the order Stigonematales, Phytochemistry, 98, 145–150, https://doi.org/10.1016/j.phytochem.2013.11.007, 2014. Blake, R. E., Chang, S. J., and Lepland, A.: Phosphate oxygen isotopic evidence for a temperate and biolog- ically active Archaean ocean, Nature, 464, 1029–1032, https://doi.org/10.1038/nature08952, 2010. Boreham, C. J., Crick, I. H., and Powell, T. G.: Alternative calibration of the Methylphenanthrene Index against vit- rinite reflectance: Application to maturity measurements on oils and sediments, Org. Geochem., 12, 289–294, https://doi.org/10.1016/0146-6380(88)90266-5, 1988. Bower, D. M., Steele, A., Fries, M. D., Green, O. R., and Lindsay, J. F.: Raman Imaging Spectroscopy of a Puta- tive Microfossil from the ∼ 3.46 Ga Apex Chert: Insights from Quartz Grain Orientation, Astrobiology, 16, 169–180, https://doi.org/10.1089/ast.2014.1207, 2016. Brasier, M. D., Green, O. R., Jephcoat, A. P., Kleppe, A. K., Van Kranendonk, M. J., Lindsay, J. F., Steele, A., and Grassineau, N. V.: Questioning the evidence for Earth’s oldest fossils, Nature, 416, 76–81, https://doi.org/10.1038/416076a, 2002. Brasier, M. D., Green, O. R., Lindsay, J. F., McLoughlin, N., Steele, A., and Stoakes, C.: Critical testing of Earth’s oldest pu- tative fossil assemblage from the ∼ 3.5 Ga Apex chert, China- man Creek, Western Australia, Precambrian Res., 140, 55–102, https://doi.org/10.1016/j.precamres.2005.06.008, 2005. Brasier, M., McLoughlin, N., Green, O., and Wacey, D.: A fresh look at the fossil evidence for early Ar- chaean cellular life, Philos. T. R. Soc. B, 361, 887–902, https://doi.org/10.1098/rstb.2006.1835, 2006. Brocks, J. J.: Millimeter-scale concentration gradients of hydro- carbons in Archean shales: Live-oil escape or fingerprint of contamination?, Geochim. Cosmochim. Ac., 75, 3196–3213, https://doi.org/10.1016/j.gca.2011.03.014, 2011. Brocks, J. J. and Summons, R. E.: Sedimentary Hydrocarbons, Biomarkers for Early Life, in: Treatise on Geochemistry, edited by: Holland, H. D. and Turekian, K. K., Pergamon, Amsterdam, the Netherlands, 63–115, 2003. Brocks, J. J., Buick, R., Logan, G. A., and Summons, R. E.: Com- position and syngeneity of molecular fossils from the 2.78 to 2.45 billion-year-old Mount Bruce Supergroup, Pilbara Craton, Western Australia, Geochim. Cosmochim. Ac., 67, 4289–4319, https://doi.org/10.1016/S0016-7037(03)00208-4, 2003a. Brocks, J. J., Love, G. D., Snape, C. E., Logan, G. A., Sum- mons, R. E., and Buick, R.: Release of bound aromatic hy- drocarbons from late Archean and Mesoproterozoic kerogens via hydropyrolysis, Geochim. Cosmochim. Ac., 67, 1521–1530, https://doi.org/10.1016/S0016-7037(02)01302-9, 2003b. Brocks, J. J., Grosjean, E., and Logan, G. A.: Assessing biomarker syngeneity using branched alkanes with quaternary carbon (BAQCs) and other plastic contaminants, Geochim. Cosmochim. Ac., 72, 871–888, https://doi.org/10.1016/j.gca.2007.11.028, 2008. www.biogeosciences.net/15/1535/2018/ Biogeosciences, 15, 1535–1548, 2018 1546 J.-P. Duda et al.: The “hydrothermal pump hypothesis” Castello, D., Kruse, A., and Fiori, L.: Supercritical water gasi- fication of hydrochar, Chem. Eng. Res. Des., 92, 1864–1875, https://doi.org/10.1016/j.cherd.2014.05.024, 2014. Craig, O. E., Love, G. D., Isaksson, S., Taylor, G., and Snape, C. E.: Stable carbon isotopic characterisation of free and bound lipid constituents of archaeological ceramic vessels released by solvent extraction, alkaline hydrolysis and cat- alytic hydropyrolysis, J. Anal. Appl. Pyrol., 71, 613–634, https://doi.org/10.1016/j.jaap.2003.09.001, 2004. Delacour, A., Früh-Green, G. L., Bernasconi, S. M., Scha- effer, P., and Kelley, D. S.: Carbon geochemistry of ser- pentinites in the Lost City Hydrothermal System (30◦ N, MAR), Geochim. Cosmochim. Ac., 72, 3681–3702, https://doi.org/10.1016/j.gca.2008.04.039, 2008. Delarue, F., Rouzaud, J.-N., Derenne, S., Bourbin, M., Westall, F., Kremer, B., Sugitani, K., Deldicque, D., and Robert, F.: The Raman-derived carbonization continuum: A tool to select the best preserved molecular structures in Archean kerogens, As- trob., 16, 407–417, https://doi.org/10.1089/ast.2015.1392, 2016. Djokic, T., Van Kranendonk, M. J., Campbell, K. A., Walter, M. R., and Ward, C. R.: Earliest signs of life on land preserved in ca. 3.5 Ga hot spring deposits, Nature Comm., 8, 15263, https://doi.org/10.1038/ncomms15263, 2017. Druffel, E. R. M. and Griffin, S.: Radiocarbon in dissolved or- ganic carbon of the South Pacific Ocean, Geophys. Res. Lett., 42, 4096–4101, https://doi.org/10.1002/2015gl063764, 2015. Erwin, J. (Ed.): Lipids and biomembranes of eukaryotic microor- ganisms, Acad. Press, New York, USA, 1973. Ferralis, N., Matys, E. D., Knoll, A. H., Hallmann, C., and Sum- mons, R. E.: Rapid, direct and non-destructive assessment of fos- sil organic matter via microRaman spectroscopy, Carbon, 108, 440–449, https://doi.org/10.1016/j.carbon.2016.07.039, 2016. French, K. L., Hallmann, C., Hope, J. M., Schoon, P. L., Zum- berge, J. A., Hoshino, Y., Peters, C. A., George, S. C., Love, G. D., Brocks, J. J., and Buick, R.: Reappraisal of hydrocar- bon biomarkers in Archean rocks, P. Natl. Acad. Sci. USA, 112, 5915–5920, https://doi.org/10.1073/pnas.1419563112, 2015. Furnes, H., McLoughlin, N., Muehlenbachs, K., Banerjee, N., Staudigel, H., Dilek, Y., de Wit, M., Van Kranendonk, M. J., and Schiffman, P.: Oceanic pillow lavas and hyaloclastites as habi- tats for microbial life through time – a review, in: Links Between Geological Processes, Microbial Activities & Evolution of Life, edited by: Dilek, Y., Furnes, H., and Muehlenbachs, K., Springer, the Netherlands, 1–68, 2008. Gérard, E., Moreira, D., Philippot, P., Van Kranendonk, M., and López-Garcia, P.: Modern subsurface bacteria in pris- tine 2.7 Ga-old fossil stromatolite drillcore samples from the Fortescue Group, Western Australia, PLoS ONE, 4, e5298, https://doi.org/10.1371/journal.pone.0005298, 2009. Glikson, M., Duck, L. J., Golding, S. D., Hofmann, A., Bol- har, R., Webb, R., Baiano, C. F., and Sly, L. I.: Micro- bial remains in some earliest Earth rocks: Comparison with a potential modern analogue, Precambrian Res., 164, 187–200, https://doi.org/10.1016/j.precamres.2008.05.002, 2008. Harris, A. C., White, N. C., McPhie, J., Bull, S. W., Line, M. A., Skrzeczynski, R., Mernagh, T. P., and Tos- dal, R. M.: Early Archean Hot Springs above Epithermal Veins, North Pole, Western Australia: New Insights from Fluid Inclusion Microanalysis, Econ. Geol., 104, 793–814, https://doi.org/10.2113/gsecongeo.104.6.793, 2009. Hawkes, J. A., Rossel, P. E., Stubbins, A., Butterfield, D., Connelly, D. P., Achterberg, E. P., Koschinsky, A., Chav- agnac, V. Hansen, C. T., Bach, W., and Dittmar, T.: Effi- cient removal of recalcitrant deep-ocean dissolved organic mat- ter during hydrothermal circulation, Nat. Geosci., 8, 856–860, https://doi.org/10.1038/ngeo2543, 2015. Hawkes, J. A., Hansen, C. T., Goldhammer, T., Bach, W., and Dittmar, T.: Molecular alteration of marine dis- solved organic matter under experimental hydrother- mal conditions, Geochim. Cosmochim. Ac., 175, 68–85, https://doi.org/10.1016/j.gca.2015.11.025, 2016. Hayes, J. M.: Geochemical evidence bearing on the origin of aer- obiosis, a speculative hypothesis, in: The Earth’s Earliest Bio- sphere: Its Origin and Evolution, edited by: Schopf, J. W., Prince- ton University Press, New York, 291–301, 1983. Hickman, A. H.: The North Pole barite deposits, Pilbara Goldfield, Geol. Surv. West. Aust. Ann. Rep., 1972, 57–60, 1973. Hickman, A. H.: Precambrian structural geology of part of the Pil- bara region, Geol. Surv. West. Aust. Ann. Rep., 68–73, 1975. Hickman, A. H.: Geology of the Pilbara block and its environs, Geol. Surv. West. Aust. Bull., 127, 1983. Hickman, A. H.: Review of the Pilbara Craton and Fortescue Basin, Western Australia: crustal evolution providing environments for early life, Island Arc, 21, 1–31, 2012. Hickman, A. H. and Van Kranendonk, M. J.: A Billion Years of Earth History: A Geological Transect Through the Pilbara Cra- ton and the Mount Bruce Supergroup – a field guide to accom- pany 34th IGC Excursion WA-2, Geol. Surv. West. Aust., Record 2012/10, 2012. Hofmann, A.: Archaean Hydrothermal Systems in the Barberton Greenstone Belt and Their Significance as a Habitat for Early Life, in: Earliest Life on Earth: Habitats, Environments and Methods of Detection, edited by: Golding, S. and Glikson, M., Springer, Dordrecht, 51–78, https://doi.org/10.1007/978-90-481- 8794-2_3, 2011. Kaneda, T.: Iso- and Anteiso-Fatty Acids in Bacteria: Biosynthesis, Function, and Taxonomic significance, Microbiol. Rev., 55, 288– 302, 1991. Killops, S. D. and Killops, V. J.: Introduction to organic geochem- istry, Blackwell Publ., Malden, USA, 2005. Kissin, Y. V.: Catagenesis and composition of petroleum: ori- gin of n-alkanes and isoalkanes in petroleum crudes, Geochim. Cosmochim. Ac., 51, 2445–2457, https://doi.org/10.1016/0016- 7037(87)90296-1, 1987. Knoll, A. H.: Paleobiological perspectives on early eu- karyotic evolution, CSH Perspect. Biol., 6, a016121, https://doi.org/10.1101/cshperspect.a016121, 2014. Koga, Y. and Morii, H.: Biosynthesis of Ether-Type Polar Lipids in Archaea and Evolutionary Considerations, Microbiol. Mol. Biol. R., 71, 97–120, https://doi.org/10.1128/mmbr.00033-06, 2007. Konn, C., Charlou, J. L., Donval, J. P., Holm, N. G., De- hairs, F., and Bouillon, S.: Hydrocarbons and oxidized or- ganic compounds in hydrothermal fluids from Rainbow and Lost City ultramafic-hosted vents, Chem. Geol., 258, 299–314, https://doi.org/10.1016/j.chemgeo.2008.10.034, 2009. Li, W., Czaja, A. D., Van Kranendonk, M. J., Beard, B. L., Roden, E. E., and Johnson, C. M.: An anoxic, Fe(II)-rich, U-poor ocean Biogeosciences, 15, 1535–1548, 2018 www.biogeosciences.net/15/1535/2018/ J.-P. Duda et al.: The “hydrothermal pump hypothesis” 1547 3.46 billion years ago, Geochim. Cosmochim. Ac., 120, 65–79, https://doi.org/10.1016/j.gca.2013.06.033, 2013. Lindsay, J. F., Brasier, M. D., McLoughlin, N., Green, O. R., Fo- gel, M., Steele, A., and Mertzman, S. A.: The problem of deep carbon—An Archean paradox, Precambrian Res., 143, 1–22, https://doi.org/10.1016/j.precamres.2005.09.003, 2005. Lockhart, R. S., Meredith, W., Love, G. D., and Snape, C. E.: Release of bound aliphatic biomarkers via hydropyrolysis from Type II kerogen at high maturity, Org. Geochem., 39, 1119–1124, https://doi.org/10.1016/j.orggeochem.2008.03.016, 2008. Love, G. D., Snape, C. E., Carr, A. D., and Houghton, R. C.: Release of covalently-bound alkane biomarkers in high yields from kero- gen via catalytic hydropyrolysis, Org. Geochem., 23, 981–986, https://doi.org/10.1016/0146-6380(95)00075-5, 1995. Love, G. D., Bowden, S. A., Jahnke, L. L., Snape, C. E., Campbell, C. N., Day, J. G., and Summons, R. E.: A cat- alytic hydropyrolysis method for the rapid screening of micro- bial cultures for lipid biomarkers, Org. Geochem., 36, 63–82, https://doi.org/10.1016/j.orggeochem.2004.07.010, 2005. Marshall, C. P., Love, G. D., Snape, C. E., Hill, A. C., All- wood, A. C., Walter, M. R., Van Kranendonk, M. J., Bowden, S. A., Sylva, S. P., and Summons, R. E.: Structural charac- terization of kerogen in 3.4 Ga Archaean cherts from the Pil- bara Craton, Western Australia, Precambrian Res., 155, 1–23, https://doi.org/10.1016/j.precamres.2006.12.014, 2007, 2007. Matsumi, R., Atomi, H., Driessen, A. J. M., and van der Oost, J.: Isoprenoid biosynthesis in Archaea – Biochemical and evolutionary implications, Res., Microbiol., 162, 39–52, https://doi.org/10.1016/j.resmic.2010.10.003, 2011. McCollom, T. M. and Seewald, J. S.: Carbon isotope composi- tion of organic compounds produced by abiotic synthesis un- der hydrothermal conditions, Earth Planet. Sc. Lett., 243, 74–84, https://doi.org/10.1016/j.epsl.2006.01.027, 2006. McCollom, T. M., Ritter, G., and Simoneit, B. R. T.: Lipid Synthesis Under Hydrothermal Conditions by Fischer- Tropsch-Type Reactions, Origins Life Evol. B., 29, 153–166, https://doi.org/10.1023/a:1006592502746, 1999. Meredith, W., Russell, C. A., Cooper, M., Snape, C. E., Love, G. D., Fabbri, D., and Vane, C. H.: Trapping hydropyrolysates on silica and their subsequent thermal desorption to facilitate rapid fingerprinting by GC-MS, Organic Geochem., 35, 73–89, https://doi.org/10.1016/j.orggeochem.2003.07.002, 2004. Mißbach, H., Duda, J.-P., Lünsdorf, N. K., Schmidt, B. C., and Thiel, V.: Testing the preservation of biomarkers during exper- imental maturation of an immature kerogen, Int. J. Astrob. 15, 165–175, https://doi.org/10.1017/S1473550416000069, 2016. Morag, N., Williford, K. H., Kitajima, K., Philippot, P., Van Kranendonk, M. J., Lepot, K., Thomazo, C., and Valley, J. W.: Microstructure-specific carbon isotopic signatures of organic matter from ∼3.5 Ga cherts of the Pilbara Craton support a biologic origin, Precambrian Res., 275, 429–449, https://doi.org/10.1016/j.precamres.2016.01.014, 2016. Nijman, W., de Bruijne, K. H., and Valkering, M. E.: Growth fault control of Early Archaean cherts, barite mounds and chert-barite veins, North Pole Dome, Eastern Pilbara, Western Australia, Precambrian Res. 98, 247–274, https://doi.org/10.1016/S0301- 9268(97)00062-4, 1999. Olcott-Marshall, A., Jehlicˇka, J., Rouzaud, J. N., and Mar- shall, C. P.: Multiple generations of carbonaceous ma- terial deposited in Apex chert by basin-scale pervasive hydrothermal fluid flow, Gondwana Res., 25, 284–289, https://doi.org/10.1016/j.gr.2013.04.006, 2014. Parfrey, L. W., Lahr, D. J., Knoll, A. H., and Katz, L. A.: Estimat- ing the timing of early eukaryotic diversification with multigene molecular clocks, P. Natl. Acad. Sci. USA, 108, 13624–13629, https://doi.org/10.1073/pnas.1110633108, 2011. Peters, K. E., Walters, C. C., and Moldowan, J. M.: The Biomarker Guide: Volume 2, Biomarkers and Isotopes in Petroleum Ex- ploration and Earth History, Cambridge University Press, Cam- bridge, UK, 2005. Philippot, P., Van Zuilen, M., Lepot, K., Thomazo, C., Farquhar, J., and Van Kranendonk, M. J.: Early Archaean microorganisms preferred elemental sulfur, not sulfate, Science, 317, 1534–1537, 2007. Pinti, D. L., Hashizume, K., Sugihara, A., Massault, M., and Philip- pot, P.: Isotopic fractionation of nitrogen and carbon in Pale- oarchean cherts from Pilbara craton, Western Australia: Origin of 15N-depleted nitrogen, Geochim. Cosmochim. Ac., 73, 3819– 3848, https://doi.org/10.1016/j.gca.2009.03.014, 2009. Proskurowski, G., Lilley, M. D., Seewald, J. S., Früh-Green, G. L., Olson, E. J., Lupton, J. E., Sylva, S. P., and Kelley, D. S.: Abio- genic hydrocarbon production at Lost City hydrothermal field, Science, 319, 604–607, 2008. Radke, M. and Welte, D. H.: The Methylphenanthrene Index (MPI): A Maturity Parameter based on Aromatic Hydrocarbons, in: Ad- vances in Organic Geochemistry 1981: 10th International Meet- ing on Organic Geochemistry, Bergen, September 1981, Pro- ceedings, edited by: Bjorøy, M. and European Association of Or- ganic Geochemists, 504–512, Wiley Chichester, 1983. Radke, M., Leythaeuser, D., and Teichmüller, M.: Relation- ship between rank and composition of aromatic hydrocarbons for coals of different origins, Org. Geochem., 6, 423–430, https://doi.org/10.1016/0146-6380(84)90065-2, 1984. Rippka, R., Deruelles, J., Waterbury, J. B., Herdman, M., and Stanier, R. Y.: Generic assignments, strain histories and prop- erties of pure cultures of cyanobacteria, J. Gen. Microbiol., 111, 1–61, https://doi.org/10.1099/00221287-111-1-1, 1979. Schidlowski, M.: Carbon isotopes as biogeochemical recorders of life over 3.8 Ga of Earth history: evolution of a concept, Precambrian Res., 106, 117–134, https://doi.org/10.1016/S0301- 9268(00)00128-5, 2001. Schopf, J. W.: Microfossils of the Early Archean Apex chert: new evidence of the antiquity of life, Science, 260, 640–646, https://doi.org/10.1126/science.260.5108.640, 1993. Schopf, J. W., Kudryavtsev, A. B., Agresti, D. G., Wdowiak, T. J., and Czaja, A. D.: Laser-Raman imagery of Earth’s earliest fos- sils, Nature, 416, 73–76, https://doi.org/10.1038/416073a, 2002. Shen, Y., Buick, R., and Canfield, D. E.: Isotopic evidence for mi- crobial sulphate reduction in the early Archaean era, Nature, 410, 77–81, https://doi.org/10.1038/35065071, 2001. Shirey, S. B. and Richardson, S. H.: Start of the Wilson Cycle at 3 Ga Shown by Diamonds from Subcontinental Mantle, Science, 333, 434–436, https://doi.org/10.1126/science.1206275, 2011. Simoneit, B. R. T.: Aqueous high-temperature and high-pressure organic geochemistry of hydrothermal vent systems, Geochim. Cosmochim. Ac., 57, 3231–3243, https://doi.org/10.1016/0016- 7037(93)90536-6, 1993. www.biogeosciences.net/15/1535/2018/ Biogeosciences, 15, 1535–1548, 2018 1548 J.-P. Duda et al.: The “hydrothermal pump hypothesis” Smithies, R. H., Champion, D. C., Van Kranendonk, M. J., Howard, H. M., and Hickman, A. H.: Modern-style subduction processes in the Mesoarchaean: Geochemical evidence from the 3.12 Ga Whundo intra-oceanic arc, Earth Planet. Sc. Lett., 231, 221–237, https://doi.org/10.1016/j.epsl.2004.12.026, 2005. Snape, C. E., Bolton, C., Dosch, R. G., and Stephens, H. P.: High liquid yields from bituminous coal via hydropy- rolysis with dispersed catalysts, Ener. Fuel., 3, 421–425, https://doi.org/10.1021/ef00015a028, 1989. Summons, R. E. and Hallmann, C.: Organic geochemical signatures of early life on earth, in: Treatise on geochemistry 2nd Ed., edited by: Falkowski, P. and Freeman, K., Elsevier, the Netherlands, 33– 46, https://doi.org/10.1016/B978-0-08-095975-7.01005-6, 2014. Terabayashi, M., Masada, Y., and Ozawa, H.: Archean ocean- floor metamorphism in the North Pole area, Pilbara Cra- ton, Western Australia, Precambrian Res., 127, 167–180, https://doi.org/10.1016/S0301-9268(03)00186-4, 2003. Ueno, Y., Isozaki, Y., Yurimoto, H., and Maruyama, S.: Car- bon Isotopic Signatures of Individual Archean Microfos- sils(?) from Western Australia, Int. Geol. Rev., 43, 196–212, https://doi.org/10.1080/00206810109465008, 2001. Ueno, Y., Yoshioka, H., Maruyama, S., and Isozaki, Y.: Carbon iso- topes and petrography of kerogens in ∼3.5-Ga hydrothermal sil- ica dikes in the North Pole area, Western Australia, Geochim. Cosmochim. Ac., 68, 573–589, https://doi.org/10.1016/S0016- 7037(03)00462-9, 2004. Ueno, Y., Yamada, K., Yoshida, N., Maruyama, S., and Isozaki, Y.: Evidence from fluid inclusions for microbial methano- genesis in the early Archaean era, Nature, 440, 516–519, https://doi.org/10.1038/nature04584, 2006. Van Kranendonk, M. J.: North Shaw, W.A. Sheet 2755: West. Aust. Geol. Surv. 1 : 100 000, Geol. Series, 1999. Van Kranendonk, M. J.: Volcanic degassing, hydrothermal circu- lation and the flourishing of early life on Earth: A review of the evidence from c. 3490–3240 Ma rocks of the Pilbara Supergroup, Pilbara Craton, Western Australia, Earth-Sci. Rev., 74, 197–240, https://doi.org/10.1016/j.earscirev.2005.09.005, 2006. Van Kranendonk, M. J.: Morphology as an Indictor of Biogenic- ity for 3.5–3.2 Ga Fossil Stromatolites from the Pilbara Cra- ton, Western Australia, in: Advances in Stromatolite Geobiol- ogy, edited by: Reitner, J., Quéric, N.-V., and Arp, G., 537–554, Springer Berlin Heidelberg, 2011. Van Kranendonk, M. J. and Pirajno, F.: Geochemistry of metabasalts and hydrothermal alteration zones associated with c. 3.45 Ga chert and barite deposits: implications for the geological setting of the Warrawoona Group, Pilbara Craton, Australia, Geochem.-Explor. Env. A., 4, 253–278, https://doi.org/10.1144/1467-7873/04-205, 2004. Van Kranendonk, M. J., Webb, G. E., and Kamber, B. S.: Ge- ological and trace element evidence for a marine sedimen- tary environment of deposition and biogenicity of 3.45 Ga stromatolitic carbonates in the Pilbara Craton, and support for a reducing Archaean ocean, Geobiology, 1, 91–108, https://doi.org/10.1046/j.1472-4669.2003.00014.x, 2003. Van Kranendonk, M. J., Philippot, P., Lepot, K., Bodor- kos, S., and Pirajno, F.: Geological setting of Earth’s old- est fossils in the ca. 3.5 Ga Dresser Formation, Pilbara Craton, Western Australia. Precambrian Res., 167, 93–124, https://doi.org/10.1016/j.precamres.2008.07.003, 2008. Walter, M. R., Buick, R., and Dunlop, J. S. R.: Stromatolites 3,400– 3,500 Myr old from the North Pole area, Western Australia, Na- ture, 284, 443–445, 1980. Yui, T. F., Huang, E., and Xu, J.: Raman spectrum of carbona- ceous material: a possible metamorphic grade indicator for low- grade metamorphic rocks, J. Metamorph. Geol., 14, 115–124, https://doi.org/10.1046/j.1525-1314.1996.05792.x, 1996. Biogeosciences, 15, 1535–1548, 2018 www.biogeosciences.net/15/1535/2018/