Precambrian Research 196–197 (2012) 113–127 Contents lists available at SciVerse ScienceDirect Precambrian Research journal homepage: www.elsevier.com/locate/precamres Biomarkers of black shales formed by microbial mats, Late Mesoproterozoic (1.1 Ga) Taoudeni Basin, Mauritania Martin Blumenberg a,∗ , Volker Thiel a , Walter Riegel a , Linda C. Kah b , Joachim Reitner a a b Geoscience Center, Geobiology Group, Georg-August-University Göttingen, Goldschmidtstr. 3, 37077 Göttingen, Germany Department of Earth & Planetary Sciences, University of Tennessee, 1412 Circle Drive, Knoxville, TN 37996, USA a r t i c l e i n f o Article history: Received 6 May 2011 Received in revised form 31 October 2011 Accepted 18 November 2011 Available online 28 November 2011 Keywords: Late Mesoproterozoic Biomarkers Hopanes High aromaticity Taoudeni Basin Acritarchs a b s t r a c t Hydrocarbon biomarkers in Late Mesoproterozoic black shales from the Taoudeni Basin (Mauritania, northwestern Africa) were analysed for palaeoenvironmental reconstruction. Within the Atar Group, Touirist Formation shales showed the highest content of organic carbon (>20%) at very low maturity (Rc less than 0.6). Microfacies and biomarker data indicate a shallow marine, high productivity, low oxygen setting with benthic mats as key players for the accumulation of organic matter. Both, high C/S-ratios and the occurrence of rearranged hopanes indicate that euxinic conditions were not prevalent, likely reflecting the shallow depositional environment. Steranes were not observed, indicating only a minor importance of modern, eukaryotic algae. Although low in abundance, a palynological survey, however, reveals the presence of simple ornamented acritarchs. Input of highly aromatic biopolymers typical of Mesoproterozoic acritarchs or microbial exopolymeric substances may be the origin of an unusually high aromaticity observed for the biomarker extracts. High amounts of hopanes suggest that the benthic mats were dominated by (cyano)bacteria. Furthermore, the presence of 2,3,6-trimethyl aryl isoprenoids points at contributions of organic matter from anoxygenic phototrophic bacteria. However, low concentrations of these compounds argue against major photic zone anoxia in the overlying water column and rather suggest a role of anoxygenic phototrophs as part of the benthic microbial mat community. The high abundances of hopanes in these samples suggest that nitrogen-limited conditions may have been common in the Taoudeni Basin at 1.1 Ga. These results are consistent with ideas of widespread nitrogen deficiency that emerged in Mesoproterozoic oceans due to high denitrification rates in anoxic deep waters. Cyano- and other phototrophic bacteria, as well as a limited number of acritarch species were able to cope with such conditions much better than modern algae with their higher nitrogen and trace metal demands. © 2011 Elsevier B.V. All rights reserved. 1. Introduction A wide variety of metabolisms were established in the Earth’s Oceans by 3.4 Ga (Allwood et al., 2006). However, the relative importance of different prokaryotic metabolisms and the timing of evolution and expansion of eukaryotic life remain controversial (Rasmussen et al., 2008; Javaux et al., 2010). Prior to 2.4 Ga Earth’s oceans harbored exceptionally low sulfate concentrations, and oxygen free deep waters with highly abundant ferrous Fe (Canfield, 2005). Although oxygen began to accumulate after about 2.4–2.2 Ga (the “Great Oxidation Event”) it took at least until the end of the Precambrian for deep oceanic water masses to show substantial signs of increasing sulfate concentrations and oxygenation (Kah et al., 2004; see Kah and Bartley (2011) for a review). Rather, it ∗ Corresponding author. Tel.: +49 0 551 3913756; fax: +49 0 551 397918. E-mail address: [email protected] (M. Blumenberg). 0301-9268/$ – see front matter © 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.precamres.2011.11.010 has been speculated that stable anoxic and potentially sulfide-rich deeper ocean waters were overlain by a shallow oxygenated upper water column for most of the Proterozoic (Canfield, 1998). Support for euxinic conditions in Proterozoic oceans comes from bulk C/S-ratios (Imbus et al., 1992; Shen et al., 2002) and the occurrence of biomarkers in shales from the McArthur basin (1.64 Ga), where euxinic conditions obviously extended into the photic zone (Brocks et al., 2005). If persistent, photic zone euxinia may have fueled anoxygenic phototrophy using H2 S as electron donor, yet it is unclear whether this mode of primary production occurred only in specific basins or, as it has been proposed, characterised the entire Proterozoic (Johnston et al., 2009). A more complete understanding of the distribution of low-oxygen environments and their associated microbial populations is also critical for understanding the rise and importance of eukaryotic algae through Precambrian times (Anbar and Knoll, 2002). Biomarkers in sedimentary rocks and oils may persist over billions of years and can be valuable sources for information on 114 M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 Fig. 1. Basemap showing the El Mreiti Gp. outcrops in Mauritania and the core site. ancient environments. These molecular biosignatures and their stable isotope signatures have the potential to shed light on the prevailing organisms as well as on the metabolic pathways and carbon sources. Although biomarker data is still limited in its extent, it provides substantial insights into life in Earth’s early oceans (Mycke et al., 1988; Summons et al., 1988a,b, 1999; Pratt et al., 1991; Logan et al., 1995, 2001; Brocks et al., 1999; Greenwood et al., 2004; Eigenbrode et al., 2008; Waldbauer et al., 2009). Unfortunately, recent studies (Brocks et al., 2008; Rasmussen et al., 2008; Brocks, 2011) have indicated that at least some of the biomarkers inventories presented in the above mentioned studies were not syngenetic to their host rock but rather the result of anthropogenic contaminations with hydrocarbons. Consequently, some interpretations regarding the distribution of prokaryotic and eukaryotic communities in the Precambrian are now being questioned (e.g., the early rise of modern eukaryotic algae based on findings of abundant steranes; Brocks (2009)). This and the low number of localities providing valuable biomarker information stress the need of additional studies from a variety of Precambrian settings. The Taoudeni Basin, northwestern Africa, records a discontinuous sedimentation history from the Proterozoic to the Cenozoic, and contains a thick succession of Proterozoic-aged sedimentary rocks. These rocks of the Taoudeni Basin include two supergroups, the lower of which (Supergroup 1) rests unconformably on Archean to Paleoproterozoic basement and consists of the Char, Atar/El Mreiti, and Assabet el Hassiane/Bir Amrane Groups. Carbon-isotope chemostratigraphy has determined the ∼800 m thick Atar/El Mreiti Group to be late Mesoproterozoic in age (Teal and Kah, 2005; Kah et al., 2009), which is consistent with Re-Os ages of 1107 ± 12 Ma, 1109 ± 22 Ma and 1105 ± 37 Ma for black shale within the Atar/El Mreiti Group (Rooney et al., 2010). With the exception, however, of ␦13 C analyses on stromatolitic fabrics of the Atar Group (Teal and Kah, 2005; Kah and Bartley, 2011), recent facies analysis of stromatolite-bearing strata (Kah et al., 2009), and a single report on general geochemical characteristics of the Touirist Formation organic matter (Doering et al., 2009), there has been little study concerning the biological make-up of Atar/ElMreiti Group strata and its environmental interpretation. Here we report the first biomarker studies of Taoudeni Basin strata (Touirist Formation) and use an approach to the interpretation of depositional environments based on biomarkers, stable isotope signatures, microfacies, and palynology. Results of this study provide compelling evidence on the authenticity of the biosignatures and allow for a robust reconstruction of a Late Mesoproterozoic highly productive, shallow marine setting. 2. Materials and methods 2.1. Rock samples Core samples were drilled in 2007 (for site see Fig. 1). Bisects of the core (4 samples) were used for biomarker, microfacies, and palynological analyses (for positions in the lithological section see Fig. 2). For biomarker studies, bisects were separated into exterior and interior parts, and both samples were analysed individually. A clean precision saw (Buehler IsoMet 1000) was used, which was carefully cleaned after each saw step with clean water and M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 115 2.3. Gas chromatography–mass spectrometry (GC–MS) and gas chromatography combustion isotope ratio-mass spectrometry (GC–C–IRMS) Fig. 2. Simplified lithologic section of the investigated core with TOC and Tmax trends (unpublished results) and the position of the analysed rock samples. acetone. Samples were then crushed and powdered using a pebble mill (Retsch MM 301) that was pre-cleaned with solvents and sand. Bulk C/N/S analysis was performed using a Hekatech Euro EA CNS analyser. To determine the contents of TOC and carbonate carbon (Ccarb ), each sample was analysed both before and after acidification with HCl. Bulk ␦13 C and ␦15 N isotope analysis was made via elemental analysis-isotope ratio mass spectrometry (EA-IRMS, Delta plus, Thermo Finnigan). For the detection of potential contaminations, the exterior and interior parts of the drill core bisects were compared. Generally, both yielded similar hydrocarbon compositions indicating only minor external contamination. Moreover, branched alkanes with quarternary carbon (BAQCs) were not found in any sample, which excludes the possibility of appreciable contamination from polyethylene plastic bags (Brocks et al., 2008). 2.2. Extraction and fractionation Ground samples (∼10 g) were subjected twice with distilled dichloromethane (DCM)/n-hexane (9/1; ∼100 ml each) using ultrasonication followed by extraction of hydrocarbons with n-hexane (100 ml). Extracts were combined and concentrated in a precleaned rotary evaporator, followed by reduction of the solvent to near dryness in a stream of N2 . The resulting total extracts were separated by column chromatography into a saturated fraction (F1), an aromatic fraction (F2) and a polar residue. 7.5 g of dry silica gel 60 were filled into a column with an internal diameter of about 1.5 cm. The extract was absorbed onto 500 mg of silica gel 60 and placed onto the column. F1 was eluted with 1.5 dead volumes of n-hexane (15 ml), F2 with two dead volumes n-hexane/DCM (1/1; 20 ml) and the polar residue with 20 ml DCM/methanol (1/1). Fractions were concentrated by careful rotary evaporation, dried under a stream of N2 and dissolved in 200 l n-hexane (F1) and 1500 l (F2), respectively. To remove elemental sulfur reduced copper was added to F1 and F2. Saturated (F1) and aromatic (F2) hydrocarbons were analysed by combined gas chromatography–mass spectrometry (GC–MS) using a Varian CP-3800 gas chromatograph coupled to a Varian 1200L mass spectrometer. Biomarkers were identified by comparing mass spectra and retention times with published data and/or reference compounds. The GC was equipped with a deactivated retention gap (internal diameter (i.d.) 0.53 mm) connected to a fused silica capillary column (Phenomenex Zebron ZB-1MS, 60 m, 0.10 m film thickness, 0.25 mm i.d.), with He as the carrier gas. The temperature program was 60 ◦ C (for 5 min), then increased by 10 ◦ C/min to 150◦ (held 0 min), and finally increased by 4 ◦ C/min to 310 (held for 25 min.). The MS source was operated at 200 ◦ C in electron impact mode at 70 eV ionisation energy. Fractions were injected on column using a PTV injector. The injector was initially held at 80 ◦ C for 0.2 min and then heated at 150 ◦ C/min to 320 ◦ C (held for 15 min). Biomarkers were analysed (i) in total ion current mode (from 50 to 650 amu; TIC), (ii) by selected ion monitoring (SIM) of nine ions (total cycle time 0.9 s), and (iii) by multiple reaction monitoring (MRM) with a total cycle time of 1.6 s per scan for 16 transitions. Argon was used as the collision gas. The collision energy was −10 eV. Typically, biomarkers were identified from MRM and quantified from TIC. Relative quantification of hopanes was achieved from SIM analyses, because of better peak resolutions. The detection limit for steranes was checked in MRM mode using authentic standards (5␣(H)cholestane and 4␣,23,24-trimethylcholestane (dinosterane)). The limit of detection for the first was ∼10 pg, representing for the dilutions used, ∼0.5 ppm of the saturates fraction. For dinosterane it was slightly higher (20 pg in general, representing ∼1 ppm of the saturates fraction). Trimethylated aryl isoprenoids were identified by selective MRM transitions and comparison with the aromatic fraction of a Jurassic black shale (Bächental, Toarcian), that contains abundant isorenieratene and isorenieratene degradation products (Köster et al., 1995). The ␦13 C-values of hydrocarbons were analysed (minimum of three replicates) using a Thermo Scientific Trace GC coupled to a Delta Plus isotope-ratio MS. The combustion reactor contained CuO, Ni, and Pt and was operated at 940 ◦ C. The GC was equipped with a Zebron ZB-5MS (30 m; 0.25 m film thickness, and 0.32 mm i.d.) and runs using a temperature program identical to that described above. The stable carbon isotope compositions are reported in the delta notation (␦13 C) vs. the VPDB standard. Standard deviations for ␦13 C analyses were generally less than 1‰ and sometimes higher due to bad peak separation or influences from co-eluting unresolved complex mixtures. Selected samples were subjected to Laser-ICP-MS analyses performed according Sánchez-Beristain et al. (2010) and to micro carbonate carbon stable isotope analyses according to Delecat et al. (2010), respectively. 2.4. Palynology One sample (sample MCS 4) was prepared for palynological analysis by demineralisation with HCl and HF. In order to mechanically dissociate the organic matrix and disintegrate remaining larger particles, the residue was crushed in a mortar, briefly boiled in 15% H2 O2 and exposed to ultrasonic treatment. Fine detritus was removed by screening at 15 m mesh size. The residue was stored in glycerine and three permanent preparations were mounted in glycerine jelly. Because of their scarcity within the bulk matrix, 116 M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 Fig. 3. Images of palynomorph residuals (A–D) and microfacies (E–J) of the Touirist Fm. black shales. (A) Microbial mat microstructure showing typical oval growth form in residue of sample MCS 4 and a relatively thick-walled leiosphaerid acritarch. (B) Leiosphaeridia sp. with lanceolate fold. (C) Leiosphaeridia sp. with concentric fold and thinned central area (opening?). (D) Leiosphaeridia sp. with thin wrinkled wall and irregular compression folds. (E and F) Microfacies images from MCS 4. E (transmitted light): The facies is characterised by undisturbed laminated organic-rich, silty sediments. The laminae are several millimetre-long condensed flakes of organic matter, which formerly mostl likely represented benthic microbial mats. The laminae have a thickness of 50–100 m and trapped silt and clay material. F (reflected light): Some of the M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 117 Laser micro-Raman spectroscopy was performed on selected palynomorphs (not shown). The spectra were dominated by stretching band at 1600 cm−1 (typical for C C bonds like in aromatic rings) with only weak signals at 1345 cm−1 (specific for CH3 ) (not shown). 3.3. Microfacies Fig. 4. Statistical size distribution of palynomorphs in sample MCS 4. palynomorphs were picked individually for photography under transmitted light, SEM, and Laser micro-Raman spectroscopy. 3. Results 3.1. Bulk geochemical characteristics TOC of the investigated Touirist Formation black shales were high (10.1–22.2%; Table 1) and showed even more elevated concentrations in the stratigraphic highest section (see also Fig. 2). Ccarb values were generally low (0.1–0.5%; Table 1) and sulfur concentrations were variable (0.8–3.1%), again with most elevated concentrations high in the stratigraphic section. Bulk carbon ␦13 C values showed an increasing trend from −33.9‰ vs. VPDB low in the stratigraphic section to −29.7‰ in the stratigraphically highest sample. Bulk nitrogen stable isotopes fluctuated between +4.9 and +6.1‰ (vs. air) and showed no trend with respect to core depth. 3.2. Palynology Palynomorphs recovered are predominantly simple more or less smooth-walled sphaeromorph acritarchs without distinct opening structure which can be classified as Leiosphaeridia sp. (Fig. 3A–D;). Fig. 4 shows the statistical size range of palynomorphs in sample MCS 4 with a peak between 20 and 40 m (n = 125). However, individual larger leiospheres with a diameter up to 125 m were also observed. Rather well defined round to oval microstrucutres, which are solid, lenticular in cross section and range in size mainly between 200 m and 400 m (Fig. 3A) are a striking component of the maceration residue. They occur isolated or in clusters and show a distinct orange fluorescence with blue light irradiation. They are tentatively considered here as growth forms of bacterial colonies significantly contributing to the formation of the microbial mats. Microfacies analyses revealed layered occurrences of organic matter with occasionally abundant pyrite (Fig. 3). In the stratigraphically highest sample (MCS 1) the organic matter was, compared to MCS 4, less finely dispersed and undisturbed and revealed ripple-like structural appearances. In MCS 2 we found additional interesting features (Fig. 3G–H). In this sample autochthonous carbonates of dispersed distributed euhedral nonluminescent bright dolomites were observed which were most likely diagenetically formed within the sediment. In addition, luminescent early diagentic automicrites were also observed. The automicrites form a halo around former microbial mat remains, which contain very small sized (10–20 m) strong luminescent anhedral calcite crystals and are often arranged in rows (Fig. 3H). Generally the strong luminescence is caused by high amounts of Mn within the calcite crystals (Laser ICP-MS analyses; data not shown). Assumably, these high Mn values are primary signals, reflecting negative redox conditions during mineralisation process. The later diagenetic carbonate crystals do not show any luminescence, demonstrating a different redox environment. Comparable structures were observed in methane-related carbonates from the Black Sea which were formed under anoxic conditions (Reitner et al., 2005). Therefore, the luminescent laminae are most likely lithified former biofilms with the strong luminescent anhedral calcites to be probably small lithified microbial communities. Those are embedded in a former matrix of exopolymeric substances, which is now weakly calcified with only minor red luminescence. Comparable lithified microbial laminae are also known from the Sturtian Rasthof cap carbonates from Namibia (Pruss et al., 2010). Pyrite is occasionally associated with calcified microbial mats. The small strong luminescent calcite crystals are probably lithified microbial colonies located within the microbial mats. Selected carbon stable isotope analyses of the automicrite mineral exhibit relatively 13 C depleted values (␦13 C = −16‰ VPDB) which is in the range described from carbonates formed during microbial sulfate reduction (Raiswell and Fisher, 2000) and are thus in agreement with a similar source than that for abundant pyrite crystals within the calcified microbial mats. 3.4. Biomarkers Extractable hydrocarbons consisted mainly of saturated (nalkanes, cycloalkanes, methylated alkanes, isoprenoids) and aromatic compounds, with generally low saturate/aromatic-ratios (<0.1; Table 2). 3.4.1. Saturates (C15 +) Patterns of saturated hydrocarbons differed slightly between the samples, but all contain n-alkanes (C15 +) ranging from C15 to C35 with maxima at C17 to C19 (Fig. 5). The carbon laminae are enriched in small (10–50 m) sized pyrite crystals often organised in flat clusters and framboids within the organic matter (white arrow). These laminated pyrite occurrences hint to sulfate reducing microbes as an integral part of the mat community. (G and H) Microfacies images from MCS 2. G (transmitted light): The facies is characterised by laminated and sometimes folded organic-rich sediments (“roll up structures”). The sediment is in some parts enriched in carbonates like euhedral single crystal, non-luminescent dolomites (left arrow). The remains of the microbial mats are again some millimetre-sized organic flakes (right arrow). H (cathodoluminescence – CL micrograph): The microbial mat remains contain very small (10–20 m) sized strong luminescent anhedral calcite crystals which are often arranged in rows. The strong luminescence is caused by high amounts of Mn within the calcite crystals. Pyrite is occasionally also present. The small calcite crystals are probably lithified microbes located within the microbial mats. (I and J) Microfacies images from MCS 1. I (transmitted light): The facies is characterised by undulated and sometimes folded organic-rich sediments with high amounts of fine siliciclastic material. Silicilclastics are concentrated in microrippels which are covered by and bound to former microbial mats. The ripples are formed by moving sediments and overprinted by later compaction. Sand and silt grains are mostly angular which suppose a nearby source area. I (insert) “roll up structures” in MCS 1. J (reflected light): The microbial mat remains contain pyrite enriched in framboids and larger lenses (arrow). 118 M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 Table 1 Geochemical parameters of black shales of the Touirist Formation (Taoudeni Basin). Sample MCS MCS MCS MCS 1 2 3 4 Formation TOC (%) Ccarb (%) S (%) N (%) C/N C/S ␦13 C bulk ␦15 N bulk Touirist (Atar Group) Touirist (Atar Group) Touirist (Atar Group) Touirist (Atar Group) 22.2 17.2 16.1 10.1 0.1 0.1 0.1 0.2 3.1 2.1 0.9 0.8 0.5 0.4 0.4 0.2 45.8 43.4 40.7 44.1 7.3 8.3 17.9 12.3 −29.7 −30.6 −31.1 −33.8 +5.9 +5.2 +4.9 +6.1 Fig. 5. Total ion chromatograms (TICs) of the saturate hydrocarbon fractions (F1) of the black shales from the Touirist Formation (1.1 Ga). Filled circles denote n-alkanes, triangles denote cyclohexylalkanes. Table 2 Amounts of saturated and aromatic hydrocarbons. Sample Saturates (ppm) Aromatics (ppm) Saturates/aromatics MCS MCS MCS MCS 360 180 250 370 6040 3760 5330 4280 0.06 0.05 0.05 0.09 1 2 3 4 preference indices (CPI) were about 1 for all samples (0.95–1.05). Cyclohexylalkanes demonstrated similar patterns as n-alkanes, but maximized at slightly higher carbon numbers (Fig. 5; maximum cyclohexyalkane C18 vs. C17 for n-alkanes). Phytane, pristane and norpristane were observed in significant amounts (Fig. 5). Pristane to phytane ratios ranged from 0.64 to 1.01 (Table 3). Pristane to C17 and phytane to C18 ratios ranged from 0.24 to 0.35 and 0.21 to 0.31, respectively (except for the topmost samples; Pr/nC17 = 1.21 and Ph/n-C18 = 0.74). Acyclic isoprenoids with carbon 0.79 0.56 0.66 0.18 3 3 5 4 1 1 1 2 Sample MPI-1 Phen/MP MPR Rc (MPI-1) MCS MCS MCS MCS 0.49 0.51 0.43 0.36 0.03* 0.24 0.34 0.33 0.52 0.63 0.61 0.53 0.56 0.57 0.52 0.47 1 2 3 4 numbers higher than C20 (phytane), e.g. C40 -isoprenoids, were not observed. 0.42 0.42 0.40 0.53 1.61 1.93 1.96 2.19 np: not present (confirmed by specific MS–MS transitions) and chei.: cheilanthanes (tricyclic terpanes). a 2␣-MeC31/(2␣MeC31+C30) (%); from SIM (due to better peak resolution than from MRM), after Summons et al. (1999). b 3-MeC31/(3MeC31+C30) (%); from SIM (due to better peak resolution than from MRM), after Eigenbrode et al. (2008). c 17␣,21-hopane/(17␣,21-hopane+phytane); from TIC. 0.69 0.53 0.42 0.45 0.57 0.56 0.57 0.61 0.12 0.20 0.10 0.27 0.20 0.34 0.27 0.49 np np np np 1.03 0.95 1.05 0.98 18 18 18 17 0.74 0.21 0.23 0.30 1.21 0.24 0.24 0.28 0.49 0.39 0.50 0.49 0.94 0.64 1.01 0.98 1 2 3 4 MCS MCS MCS MCS 119 Table 4 Maturity indices (aromatic hydrocarbons) of black shales from the Touirist Fm. from the Taoudeni Basin. *Most likely affected by biodegradation. Methylphenanthrene Index (MPI-1) = 1.5 * (2-MP + 3-MP)/(Phen + 1-MP + 9-MP) (Radke and Welte, 1983). Since Phenanthrene (Phen)/methylphenanthrenes (MP) were <1 computed vitrinite reflectance was calculated after Boreham et al. (1988): Rc (MPI-1) = 0.7 * MPI-1 + 0.22. 0.09 0.12 0.24 0.35 3-Me-Index (%)b 22S/22R (C32 ) C24tetra/C30 Tricyclics/17␣ hopanes Steranes CPI n-Cmax Ph/n-C18 Pr/n-C17 Pr/(Pr + Ph) Pr/Ph Sample Table 3 Biomarker indices (saturated hydrocarbons) of black shales from the Touirist Fm. (Taoudeni Basin).. chei24/23 chei22/21 Ts/(Ts + Tm) Dia/C30 2␣-Me-Index (%)a Hopane indexc M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 3.4.2. Aromatic steroids and steranes Resolved sterane and steroid peaks were not observed. Only small humps were found in the range were steranes usually elute. These humps consist of unresolved complex mixtures (UCM) of compounds. However, we cannot exclude that trace amounts of steranes are obscured by this UCM, but if present they were under detection limit. The lower level of detection of steranes was <1 ppm of the saturates fraction for 5␣(H)-cholestane, which is in the same range given in other studies (e.g., see SOM of Brocks et al., 2005). Fig. 6 shows for comparison the SIM of the hopane specific fragment at m/z 191 and the sterane specific fragment at m/z 217. 4-Methyl steranes were also under detection limit, even when highly selective SIM or MRM modes were applied (according to Summons et al., 1987). This includes dinosteranes, which could not be detected using the 414 → 231 and 414 → 98 transitions. The presence of aromatic steroids as well as aromatic ringmethylated (e.g. 4-methyl) steranes was also tested by specific MS–MS transitions (m/z 344 → 231 (or 245; ring-methylated); 358 → 231 (or 245); 372 → 231 (or 245); 386 → 231 (or 245); 400 → 231 (or 245)). None of these transitions showed resolved peaks. 3.4.3. Hopanes and cheilanthanes The highest relative amounts of hopanes were observed in the stratigraphically highest sample and were clearly detectable even from total ion chromatograms (Fig. 5). Hopanes in all samples consist of C27 –C35 compounds, and generally show most significant abundances of 17␣(H),21(H)-C30 (17␣(H),21(H)-hopane; Fig. 7). 22S/22R-isomerisation ratios for homohopanes were at about ∼0.6 (Table 3). Diahopanes were found in varying amounts, with decreasing ratios of C30 -diahopane vs. 17␣(H),21(H)-hopane from the lowermost to the topmost sample (from 0.35 to 0.09). Minor to moderate concentrations of ring A methylated hopanes were observed with a 2␣-methyl hopane index of 1–2% (calculated after Summons et al., 1999) and a 3-methyl hopane index of 3–5% (calculated after Brocks et al., 2005). 29,30- and 28,30Bisnorhopanes (BNH) were only slightly above detection limit in all samples (Fig. 7). A suite of C19 –C25 tricyclic terpanes (cheilanthanes) were also observed in all samples. Moreover, all samples showed an abundance of tetracyclic C24 terpane (M+ 330), most likely 17,21secohopane (C24 ) (see Aquino Neto et al., 1983). 3.5. Aromatic hydrocarbons Aromatic hydrocarbons were present in all samples and contained high relative concentrations of dimethyl- and trimethyl naphthalenes, phenanthrene, dimethyl- and trimethyl phenanthrenes, chrysene and methylated chrysene(s) (Fig. 8) and were used to calculate maturity indices (Table 4). In the stratigraphically highest sample (MCS 1) low molecular weight 120 M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 Hopanes and diahopanes Ch21 Ch19 Ch20 Ch23 17,21-secohopane (C24) Tm Ts e.g. Ch24 Ch25 SIM m/z 191 no steranes and diasteranes ? e.g. ?? SIM m/z 217 time Fig. 6. Selected ion monitoring traces (SIM) from the saturated hydrocarbon fraction (F1) of MCS 4 (black shale from the lowermost Touirist Formation). m/z 191 shows mainly pentacyclic triterpenoids (hopanes) and tricyclic terpanes (cheilanthanes), whereas m/z 217 is a typical fragment of steranes and diasteranes. aromatics were clearly depleted relative to their concentration in stratigraphically lower horizons. Moreover, except for minor amounts of methylated chrysenes no polycyclic aromatic hydrocarbons (PAHs) of higher molecular weight were observed. section, whereas the stratigraphically highest samples showed an enrichment in 13 C of about 4–5‰ (similar to bulk stable carbon isotopes; compare Table 1 and Table 5). 3.5.1. Aromatic carotenoids Long-chain carotenoids (e.g. isorenieratene) and derivatives were not detected even after application of specific SIM and MS–MS transitions (according to Brocks and Schaeffer (2008)). However, alkylated trimethylbenzenes with dominantly 2,3,6-methylation were found after performing M+ → 134 transitions (exemplified for sample MCS 4; Fig. 9). The methylation pattern was confirmed after coinjection with a black shale rich in isorenieratane and its break-down products (Bächental, Toarcian; see also Köster et al. (1995)). The alkylated trimethylbenzenes in the Bächental sample are predominantly composed of 2,3,6-trimethyl aryl isoprenoids and exhibit similar distribution to other Toarcian black shales (Schwark and Frimmel, 2004). In the Touirist Formation, an additional short chain C21 diaryl isoprenoids (LIII in Koopmans et al. (1996a)) was also observed. However, in Touirist Formation shales abundant other aryl isoprenoids and aryl alkanes, in addition to 2,3,6-methylated isomers, were also found, much higher than reported from Toarcian shales rich in isorenieratene (Schwark and Frimmel, 2004). Those samples often exclusively include 2,3,6-trimethyl aryl isoprenoids. All shales demonstrated similar distributions of trimethylbenzenes (not shown). 4.1. Syngeneity of biomarkers 3.6. Stable carbon isotopes 4.2. Maturity and biodegradation Stable carbon isotope signatures of saturated and aromatic hydrocarbons were analysed for all shales. N-alkanes and acyclic isoprenoids had ␦13 C values between approximately −30 and −35‰ (VPDB; Table 5). Generally, n-alkyl hydrocarbons were similar in ␦13 C or slightly 13 C-depleted compared to phytane. Hopanes were found to be slightly less depleted in 13 C compared to acyclic compounds. Polycyclic aromatic hydrocarbons demonstrated similar ranges to that found for saturates. However, we did observe that the strongest 13 C-depletions occurred low in the stratigraphic Effects of biodegradation, namely an increase of phytane/nC18 and hopanes, and an apparent loss of low molecular weight PAHs are indicated for the stratigraphically highest sample. Moreover, almost all samples exhibit pronounced unresolved complex mixtures (UCM), although low molecular weight saturates (C15 +) were still present in high amounts. Methyl phenanthrene ratios (MPR) ranged between 0.52 and 0.63. Based on phenanthrene to methyl phenanthrene ratios <1 (∼0.24 to 0.34) (Brocks et al., 2003a) and methyl phenanthrene indices (MPI-1) between 0.36 4. Discussion Establishing a syngenetic relationship of the bitumen with its host rock is an underlying requirement for any attempt to interpret Precambrian ecosystems by means of hydrocarbon biomarkers. A number of observations allows the critical assessment of whether the measured compounds were originally part of the host rock. (i) First, the biomarker distributions in the exterior and interior parts of the cores are virtually identical (Table 6), pointing to the absence of drilling contamination. (ii) Second, the virtual lack of steranes (including dinosterane) and of higher plant-derived triterpanes suggest the absence of substantial contamination by Phanerozoic organic matter. (iii) Thirdly, indices describing the maturity of the bitumen (MPI-1; Ts/(Ts + Tm); see below) coincide and are in good agreement with bulk geochemical maturity assessments (i.e. Tmax values; Fig. 2). (iv) Finally, bulk ␦13 C and biomarker ␦13 C values are similar, suggesting similar organic sources. In conjunction, these indications yield compelling evidence that the saturated and aromatic hydrocarbons extracted from the Late Mesoproterozoic black shales of the Touirist Formation are syngenetic and have not been introduced at a later stage M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 121 Fig. 7. Multiple reaction monitoring (MRM) transitions typical for hopanes and norhopanes; MCS 4 (black shale from the lowermost Touirist Formation). and 0.51, computed vitrinite reflectance (Rc (%)) of 0.47–0.57 were calculated after Boreham et al. (1988); see Table 4 for details). These computed vitrinite reflectance values correspond to low maturities (early oil window to peak oil window; Killops and Killops (2005)) and references therein) and are consistent with observed ratios of 18␣-22,29,30-trisnorneohopane (Ts) to 17␣-22,29,30-trisnorhopane (Tm); Ts/(Ts + Tm) of about 0.40–0.53 (Table 3). Three samples display relatively low pristane/n-C17 ratios MP TMN diP DMN P DMP chrysene TMP methyl chrysenes perylene TIC time Fig. 8. Total ion chromatogram of the aromatic fraction (F2) of MCS 4 (black shale from the lowermost Touirist Formation). Individual aromatic hydrocarbons were identified by use of selected ion traces. DMN: dimethyl napthalenes; TMN: trimethyl naphthalenes; diP: dihydro phenanthrene; P: phenanthrene; MP: methyl phenanthrenes; DMP: dimethyl phenanthrenes; TMP: trimethyl phenanthrenes. 122 M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 Table 5 Carbon stable isotope signatures of biomarkers (saturates and aromatics). Note that pristane and phytane were not base line separated from n-alkanes and should – due to chromatographic reasons – in reality be slightly more enriched in 13 C. Sample MCS 1 n-C15 n-C16 Cyclohexyl-C10 n-C17 Pristane Cyclohexyl-C11 n-C18 Phytane Cyclohexyl-C12 n-C19 n-C20 n-C21 n-C22 n-C23 n-C24 18␣-22,29,30-Trisnorhopane (Ts) 17␣-22,29,30-Trisnorhopane (Tm) 30-Norhopane 17␣(H),21(H)-Hopane 22S-17␣(H),21(H)-Homohopane Phenanthrene 3-Methylphenanthrene (3-MP) 2-Methylphenanthrene (2-MP) 9-Methylphenanthrene (9-MP) 1-Methylphenanthrene (1-MP) Chrysene MCS 2 ␦13 C (mean) ± −35.6 −35.2 −34.9 −35.4 −35.2 −34.5 −35.2 −34.3 −35.1 −34.7 −34.9 −34.2 −35.3 −33.9 −35.1 0.3 0.6 0.7 0.1 1.3 0.8 1.8 3.0 1.3 0.8 0.8 1.5 1.1 0.6 −33.1 −34.2 −34.9 −35.3 −33.6 −32.9 0.5 0.8 0.9 2.9 1.1 1.6 MCS 3 MCS 4 ␦13 C (mean) ± ␦13 C (mean) ± −32.9 −32.5 −33.5 −32.6 −32.8 −33.7 −32.1 −32.2 −33.9 −33.3 −32.1 −32.3 −31.2 −32.8 0.2 0.6 0.6 1.9 0.4 0.1 0.1 1.0 0.4 1.3 1.1 1.9 1.2 −33.2 −32.4 −34.3 1.8 1.1 1.4 −32.1 −33.4 −31.2 −29.8 −31.2 −32.3 −32.1 −31.9 −31.1 −30.6 0.6 0.6 1.7 0.1 0.6 1.8 2.1 1.7 0.0 0.8 −29.6 2.1 −28.8 1.5 −30.8 −32.3 −32.5 −32.3 −31.8 −30.9 1.0 0.4 0.8 1.8 2.4 1.0 −29.7 −30.5 −30.5 −30.8 −30.1 −28.5 1.1 0.3 0.4 0.4 0.8 0.1 (0.21–0.35), in line with other Proterozoic rock samples (Fowler and Douglas, 1987; Summons et al., 1988b; Dutkiewicz et al., 2004). The pristane/n-C17 ratio in the stratigraphically highest sample (MCS 1) was considerably higher (1.21) than those of the other samples. But, even for MCS 1 these values corresponded to only light – most likely postdepositional biodegradation (PM 2) according to the Peters–Moldowan biodegradation scale (Peters and Moldowan, 1993). However, we also must consider the potential impact from calibration of this biodegradation scale for Phanerozoic organic matter, which typically shows appreciable contributions of the acyclic isoprenoids pristane and phytane. Low inputs of these isoprenoids, as observed for the Mesoproterozoic samples, may influence the pristane/n-C17 -ratios and phytane/n-C18 ratios a priori which makes it difficult to resolve the extent of biodegradation. Due to the presence of unresolved complex mixtures and the pronounced loss of low molecular weight compounds (saturates and aromatics), we assume that moderate biodegradation affected particularly the stratigraphically highest sample (MCS 1). 4.3. Palaeoenvironmental implications 4.3.1. General aspects Black shales in the Taoudeni Basin were deposited in the Late Mesoproterozoic at about 1.1 Ga (Rooney et al., 2010). The shales studied contain high amounts of organic matter (10–22%) and increase in the lithological section to almost 30% (Fig. 2). They were formed in marine, epeiric water environments, under potentially shallow water (Kah et al., in review; Bertrand-Sarfati and Moussine-Pouchkine, 1988). Our data reveal additional ␦13 C (mean) ± −30.5 −30.5 0.1 0.5 −30.5 −30.7 −29.2 −29.8 −29.9 −27.6 −30.2 −31.2 −30.0 −32.4 −31.3 −29.1 −28.5 −26.9 −31.9 −29.0 −29.7 1.1 0.8 0.3 0.9 3.3 0.6 −30.3 0.8 −30.8 0.4 information on the biota of the Taoudeni Basin and differences between the four sampled periods of black shale formation within the Touirist Formation. Increasing TOC and S-concentrations suggest that the generally high primary production (and/or degree of preservation) increased from the lower to the upper Touirist Formation. Although C/S ratios have to be used with caution for Proterozoic samples (e.g. Shen et al., 2002), a comparison with modern marine settings (e.g. euxinic Black Sea; <3; (Berner, 1984) suggests the lack of extensive euxinic conditions in the water column during deposition of the Touirist Formation (7.6–12.6). Euxinia, however, is believed to be prevalent through the Proterozoic (Canfield, 1998; Kah and Bartley, 2011). Absence of euxinia above the microbial mats deposited in the Touirist Fm. may reflect the shallow-water setting and the potential for diffusion of oxygen into the upper water column from the atmosphere or the effects of episodic storm mixing. Both the onset of widespread sulfate evaporite deposition in the Mesoproterozoic (Kah et al., 2001; Goodman and Kah, 2004), and depletion of Mo concentrations from similarly aged black shale (Lyons et al., 2009) support the potential for suboxic conditions prevailing in shallow waters by the late Mesoproterozoic. Moreover, a stable chemocline in the photic zone and euxinia below (photic zone anoxia) was recently reconstructed for the Barney Creek Formation (1.64 Ga), which yielded high amounts of biomarkers of anoxygenic phototrophs, such as isorenieratane, okenane, and respective break-down products (Brocks et al., 2005; Brocks and Schaeffer, 2008). Intact carotenoid derivatives were not found in the Touirist Formation black shales, thus evidence for photic zone euxinia in the water column above the microbial mats is lacking. Findings of potential break-down products of isorenieratane in Table 6 Biomarker indices (saturated hydrocarbons) of the interior and exterior of a black shale sample from the Touirist Fm. (Taoudeni Basin). Sample Pr/Ph Pr/n-C17 Ph/n-C18 n-Cmax CPI Steranes Tricyclics/17␣-hopanes 22S/22R (C32 ) Ts/(Ts+Tm) Dia/17␣-hopane MCS 4 interior MCS 4 exterior 0.98 0.75 0.28 0.28 0.30 0.31 17 18 0.98 1.05 np np 0.49 0.43 0.61 0.61 0.53 0.53 0.35 0.35 np: not present. M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 123 depth (Table 3); however, maturity can be excluded as a major control of the observed distribution, since all samples exhibit similar Rc values (∼0.5–0.6; Table 4). Likewise, biodegradation cannot be positively linked to the presence of diahopanes since lowest C30 diahopane/hopane-ratios were just observed in the potentially most biodegraded sample (top; MCS 1). In the Taoudeni Basin the abundant presence of diahopanes can be better explained by a shallow, suboxic depositional environment, where diahopanes often occur in high amounts (Peters et al., 2004a). The water depth during deposition of benthic mats reflected in sample MCS 1 appear to have been particularly shallow, since the occurrence of microripples suggests deposition above wave base (Fig. 3J). Fig. 9. Trimethylated aryl isoprenoids and alkanes in the aromatic fraction (F2) of MCS 4 (black shale from the lowermost Touirist Formation) as detected from specific multiple reaction monitoring (MRM) transitions (a, b, c, d, e). The 2,3,6trimethylation pattern of the denoted benzenes was confirmed by coinjection of MCS 4 with an aromatic hydrocarbon fraction from a Jurassic black shale (a , b , c , d , e ; Bächental, Toarcian; see Section 2). Asterisks denote 2,3,6-trimethyl aryl isoprenoids identified from coinjection. the Touirist Fm. shales may however be interpreted to reflect (i) euxinia in deeper waters of the basin, (ii) intermittent euxinia above the shallow-water microbial mats, and/or (iii) contributions from green-sulfur bacteria living in the stratified mats. Pristane/phytane ratios (0.8–1.0) neither suggest extremely reductive (1) nor oxidative (1) conditions of early diagenesis (Peters et al., 2004b). With appropriate caution considering problems of the use of pristane/phytane ratios to reconstruct oxidation state (ten Haven et al., 1987), and our restricted knowledge on the depositional environment, these data favor a suboxic depositional environment of low redox conditions. Further support for the idea of suboxic, shallow water depositional settings within the Taoudeni Basin comes from relatively high amounts of 17␣(H)-diahopanes (rearranged hopanes) – that also occur as R/S-doublets (Fig. 7). These compounds behave relatively refractory and have thus often been described from strongly biodegraded and/or highly mature bitumens (Brocks and Summons, 2003). However, diahopanes have also been reported from non-degraded and less mature samples, suggesting an unknown biological source. In Proterozoic organic matter, diahopanes have been occasionally found (Summons et al., 1988a,b; Dutkiewicz et al., 2004). In black shales of the Touirist Formation diahopane/hopane ratios conspicuously increase with core 4.3.2. Sources of organic matter Archaea were most likely no important contributors of organic matter, as indicated by the lack of C40 isoprenoids (i.e. biphytanes). Likewise, the prevalence of n-alkanes and cyclohexylalkanes in the low molecular weight range (C17 –C18 ) suggests a setting dominated by bacteria and/or algae. However, steranes (regulars, diasteranes, and aromatic steranes) were not found suggesting modern eukaryotic algae to be, if any, only a minor source of organic matter in the Late Mesoproterozoic Taoudeni Basin. Our results differ from findings of eukaryote-derived steranes in several other Mesoproterozoic and older rocks, for instance in the 1.1 Ga Nonesuch Formation (Pratt et al., 1991). It is possible, however, that the presence of abundant steranes in these samples result from contaminations (Brocks et al., 2008). Body fossils of eukaryotic algae are well-known from Mesoproterozoic-aged and even much older rocks (Butterfield, 2000; Javaux et al., 2004, 2010), but our results support the suggestion that low-oxygen conditions and nitrate limitation may have permitted only limited environmental expansion of eukaryotic algae prior to the end of the Mesoproterozoic (Anbar and Knoll, 2002; Johnston et al., 2009). The presence of cheilanthanes in the otherwise observed sterane-free samples from the Touirist Formation indicates that modern algae are not the primary source of these compounds, unlike the suggestion of earlier studies (e.g. Greenwood et al., 2000). In the Taoudeni Basin the vast majority of organic matter was likely contributed by bacteria, as indicated by the generally high abundance of hopanes. The observed suite of C30 to C35 homologues with 22S/R-isomers of 17␣(H),21(H)-homohopanes includes 2␣-methylated analogues, which are suggested to be sourced by cyanobacteria (Summons et al., 1999). The contributions of cyanobacteria are therefore frequently described by the 2␣-methyl hopane index, which can often exceed 10% for other Proterozoic samples (Summons et al., 1999). In the Touirist Formation black shales 2-methyl hopane indices are much lower (1–2%, Table 3). 2␣-methylated hopanes, however, are neither an exclusive biomarker for cyanobacteria nor do all cyanobacteria produce (2-methylated) hopanoids. For instance, in a recent study C-2 methylated hopanes were also found in anoxygenic phototrophs (Rashby et al., 2007) and the enzyme system necessary for C-2 methylated hopanoid synthesis was observed in few soil bacteria (Welander et al., 2010). On the other hand, many modern cyanobacteria, particularly marine taxa appear free of C-2 methylated hopanoids (Talbot et al., 2008). Consequently, low 2-methyl hopane indices as calculated for the black shales of the Touirist Formation do not exclude appreciable contributions of organic matter by cyanobacteria. Further support for organic matter considerably sourced by cyanobacteria comes from the very low abundance of gammacerane. Gammacerane is a diagenetic product of tetrahymanol, which often co-occurs with 2-methyl hopane producing proteobacteria (Bravo et al., 2001; Rashby et al., 2007). Indications for aerobic methanotrophic bacteria, which live at chemoclines, were found. Type I methanotrophic bacteria are reported to produce C-3 methylated hopanoids (Rohmer et al., 124 M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 1984; Zundel and Rohmer, 1985). Recently, the relative importance of contributions of methanotrophic bacteria to organic matter was proposed to be calculable by the 3-methyl hopane index (Brocks et al., 2005). 3-Methyl hopane indices for Touirist Formation black shale ranged between 3 and 5%. These values are similar to the 1.64 Ga Barney Creek setting (average 5.7%), where active methane cycling most likely occurred at an oxic-anoxic interface in the water column (Brocks et al., 2005). However, due to the lack of other indications for anoxia in the overlying water column, we propose that methanotrophic bacteria likely thrived within the benthic microbial mats, specifically at the boundary between suboxic and anoxic conditions in the top layers of the mats. Whereas intact aryl isoprenoids of chlorobiaceae (e.g. isorenieratane) or chromatiaceae (okenane; Schaeffer et al., 1997) were absent from the black shales studied, potential break-down products of isorenieratane, namely 2,3,6-trimethylated aryl isoprenoids were found in all samples (Fig. 9). These compounds have often been observed in black shales and oils with abundant isorenieratene. However, respective aryl isoprenoids do not necessarily derive from isorenieratene precursors. It has been shown that severe degradation of other non-aromatic carotenoids of common phototrophs (e.g., - and ␥-carotene) might also induce the formation of trimethylated aryl isoprenoids, including 2,3,6-trimethyl homologues (Hartgers et al., 1994; Koopmans et al., 1996a,b). Isorenieratene and its degradation products can be identified by relative 13 C-enrichments (e.g. Summons and Powell, 1987). However, due to co-elution with other aromatic low molecular weight components, ␦13 Canalyses were impossible for the aryl isoprenoids from the Touirist Formation. The predominance of 2,3,6-trimethylbenzenes with different alkyl chains (C14 –C18 ) and the finding of a C21 2,3,6trimethylated isoprenoid with two benzene rings (e.g., Koopmans et al., 1996a; Grice et al., 1997; Sinninghe Damste et al., 2001), however, may be interpreted to reflect minor contributions from anoxygenic phototrophs. If sourced by anoxygenic phototrophs, we propose the respective organisms played an active role within stratified microbial mats, similarly to recent microbial mats which often include anoxygenic phototrophic bacteria in the photic, anoxic, and sulfidic layer below the immediate surface of the mats (Brocks and Pearson, 2005; Ohkouchi et al., 2005). Consequently, in conjunction with the general phototrophic character of the mats this would require a very shallow environment (0–30 m), which is also indicated by microfacies results (Fig. 3). The saturated and aromatic hydrocarbons from the black shales from the Touirist Formation revealed ␦13 C-values between about −29 and −35‰ (VPDB). Similar ␦13 C-values were also observed in the kerogen of the black shales (Doering et al., 2009) as well as for other Proterozoic hydrocarbons, which most likely derived from benthic microbial mats (Logan et al., 1999). Pristane, phytane and hopanes were in the same range or slightly enriched in 13 C, in line with reports from the relationship between acyl- and isoprenoidal lipids from cyanobacteria and other bacteria (Sakata et al., 1997; Hayes, 2001). Table 5 demonstrates an increase in ␦13 C-values from the stratigraphically lowest to the highest black shale sample, which also exhibits the highest amounts of hopanes. However, the isotopic differences between the samples are small, and possible reasons for this variation are difficult to interpret. Staal et al. (2007) reported a positive relationship between bulk ␦13 C and the thickness of microbial mats grown in freshwater, while an opposite trend was observed for marine mats (explained by different degrees of carbon limitation and importance of heterotrophic turnover; Staal et al. (2007)). Given that the benthic microbial mats of the Taoudeni Basin grew in a dominantly marine environment, the observed increase in ␦13 C-values might be better explained by relatively higher microbial primary productivity, as recorded by mat thickness and enhanced TOC within the stratigraphically highest black shale (Fig. 3). This might have reduced carbon fractionation (i.e., 13 C-depletion), because high productivity may result in carbon substrate limitation during fixation (Schidlowski et al., 1994). Given the complexity of carbon isotope fractionation processes during organic matter biosynthesis, deposition, and dia-/catagenesis (e.g., Hayes, 1993), care has to be taken when interpreting relatively small isotope variations like those observed for the biomarkers in the black shales from the Touirist Formation. A marked characteristic of the Touirist Formation black shales is the high aromaticity of the bitumens. This feature has also observed in Archean- and Proterozoic-aged kerogens, but the causes are unclear (Imbus et al., 1992; Brocks et al., 2003b; Marshall et al., 2007). One common interpretation invokes evaporative fractionation in gaseous solutions which selectively enriches light aromatics in the residual oil, thus shifting it to higher aromaticity (Thompson, 1987). This process, however, appears to primarily affect oils in reservoirs, rather than source rocks. Another possible explanation for the high aromaticity in the black shales of the Touirist Formation is a specific input of acritarch biopolymers, although the number of acritarchs is low. Acritarchs did not strongly diversify until the Neoproterozoic, they were present and occasionally abundant through the Mesoproterozoic (Javaux et al., 2004; Huntley et al., 2006) and were recently observed in rocks of an age of 3.2 Ga (Javaux et al., 2010). However, studies using combined micro-Fourier transform infrared spectroscopy (FTIR) and laser micro Raman spectroscopy indicated that particularly Mesoproterozoic acritarchs contain biopolymers composed of highly condensed aromatic sub-structures (Arouri et al., 2000; Marshall et al., 2005) which may have contributed to the high aromaticity of the Touirist Formation black shales studied. And indeed, laser micro-Raman spectroscopy of palynomorphs collected from the Touirist Formation black shales revealed also highest abundances of stretching bands (data not shown) typical for highly aromatic kerogen (at 1600 cm−1 ) (Marshall et al., 2005). Against this background, it is interesting to see that 4␣,23,24-trimethylcholestane (dinosterane) was not found in these samples. Dinosterane is commonly attributed to dinoflagellates (Summons et al., 1987), but it has been reported to be abundant in early Cambrian rocks (Moldowan and Talyzina, 1998) that predate by far their advent and rapid spread during the Mesozoic. These findings suggest that specific acritarchs may be ancestors of modern dinoflagellates. Our results suggest that the simple acritarchs preserved in the black shales of the Touirist Formation were either not related to dinoflagellates (including a non eukaryotic origin of the palynomorphs; e.g., cysts of prokaryotes), were unable to produce (dino)sterane, or were too low in abundance to be reflected in the biomarker record. Although quantitative estimates from findings of microfossils and of biomarker abundances in extractable organic matter and kerogen are delicate, the latter possibility would require additional explanation for the extraordinary high aromaticity of the Touirist Formation black shales than acritarch biopolymers. Polycyclic aromatic hydrocarbons may be also sourced by proteins and polysaccharides (Almendros et al., 1997; Marynowski et al., 2001), which are abundant in nature, and particularly abundant in exopolymeric substances of benthic microbial mats (Wieland et al., 2008). In addition, if the organic matter has been mainly produced by microbial mats, those labile biopolymers were compared to aliphatic lipids less affected by biodegradation, than organic matter which had to be transported through a water column. Thus, cyclized biopolymers may provide an additional source for aromatic hydrocarbons in the Touirist Formation and perhaps also for other Proterozoic kerogens originating from benthic microbial mats. M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 4.4. On the predominance of (hopanoid-producing) prokaryotes vs. eukaryotic algae in the Mesoproterozoic Taoudeni Basin In contrast to earlier findings (Summons et al., 1988a; Brocks et al., 1999; Waldbauer et al., 2009), several biomarker studies suggest that eukaryotic algae were less important than prokaryotic organisms in Archaean and Proterozoic seas (e.g. Dutkiewicz et al., 2004; Brocks et al., 2005, 2008). This observation is consistent with other geological records (Knoll et al., 2006), and is further supported by new data from the Taoudeni Basin, which shows the prevalence of bacterial hopanes and the virtual absence of (eukaryote-derived) steranes. Although it has to be considered that steranes and hopanes behave differently during biodegradation and maturation, with hopanes showing a greater resistance to degradation than steranes (Seifert and Michael Moldowan, 1978; Peters and Moldowan, 1993), this holds only true at Peters–Moldowan level (PM) maturation and degradation levels >7; (Peters and Moldowan, 1993). This is far beyond the most degraded sample from the Taoudeni Basin (PM level ≤ 2 for MCS 1). Thus, the ratio between hopanes and steranes (including aromatic steranes) in immature and moderately biodegraded Precambrian samples may be applied as a relevant indicator on the distribution of ancient organisms. Given that selective loss of steranes was only weak, high amounts of hopanes and the observed lack of sterane peaks hint to bacteria as the predominating component of the microbial communities in the Taoudeni Basin as well as to an increasing contribution of hopanoid-producing bacteria to the organic matter with decreasing core depth. Whereas the function of hopanoids in bacteria is still under discussion (Berry et al., 1993; Moreau et al., 1997; Doughty et al., 2009; Welander et al., 2009), a recent survey demonstrated a high overlap of bacteria producing hopanoids and those being capable of nitrogen-fixation (Pearson et al., 2007). Although the direct role of hopanoids in the nitrogen-fixation process is questionable (Doughty et al., 2009), high rates of hopanoid production appear to be related to times of nitrogen deficiency and thus increased bacterial nitrogen-fixation (Eigenbrode et al., 2008; Blumenberg et al., 2009, 2010). The nitrogen cycle in the Proterozoic is only poorly understood. Several lines of evidences, however, indicate nitrogen limitation, which has been explained by high rates of denitrification at widespread oxic-anoxic boundaries. Although recently questioned (Klotz and Stein, 2008), it is generally assumed that nitrogen fixation is an ancient metabolism. Therefore, in nitrogen-limited areas of the early oceans, N2 -fixing bacteria may have had an advantage over other bacteria (Zerkle et al., 2006), and particularly over modern eukaryotes (Anbar and Knoll, 2002 and references cited therein). This situation might have prevailed through the Proterozoic, when oxygen and nitrate concentrations began to rise. Abundant hopanoids in the black shales of the Touirist Formation may therefore be seen in support of a nitrogen-limited, bacterially dominated scenario for the Taoudeni Basin, as it has been suggested for other Mesoproterozoic settings (Fennel et al., 2005). 5. Conclusions Biomarker, microfacies, and palynological analyses of black shales from the Late Mesoproterozoic Touirist Formation demonstrate that the majority of the highly abundant organic matter (>20% TOC) was produced by benthic bacterial mats thriving within the photic zone. The organic matter is of low maturity and biomarkers (e.g. high diahopanes, low aryl isoprenoids) and bulk geochemical analyses (high C/S-ratios) argue against pervasive euxinic conditions, supporting growth of the mats in a shallow, 125 suboxic environment. Unusually high amounts of aromatic hydrocarbons likely reflect a selective preservation of labile biopolymers like exopolymeric substances rich in polysaccharides or proteins in the microbial mats. Aerobic methanotrophy is indicated by relatively high ratios of 3-methylhopanes. Varying but generally high amounts of hopanes may be interpreted in terms of a nitrogen-limited setting that favored bacteria able to fix atmospheric nitrogen (diazotrophic (cyano)bacteria). Such a scenario is consistent with models of Proterozoic oceans containing low concentrations of bioavailable nitrogen (nitrate, ammonia) due to loss by denitrification. The practical lack of steranes further demonstrates that the palaeoenvironment of the Taoudeni Basin was unfavorable for eukaryotic algae other than acritarchs, supporting recent hypotheses on generally bacterially dominated Proterozoic Oceans. Acknowledgements We thank Sascha Döring, Patrick-Oliver Reynolds, Andreas Frischbutter and Geoff Gilleaudeau (University of Knoxville, Tennessee) for helpful discussions. We are also indebted to Cornelia Conradt, Andreas Reimer and Birgit Röring for analytical support, to Nadine Schäfer for help with Laser Micro-Raman Spectroscopy, and to Jens Dyckmans from the “Centre for Stable Isotope Research and Analysis” at the University of Göttingen for help with bulk and compound specific stable carbon isotope analysis. Jochen Brocks and an anonymous reviewer are thanked for their helpful comments considerably improving the manuscript. This work was financially supported by the Deutsche Forschungsgemeinschaft (grant BL971/1-2 and 1-3) and the Courant Research Center of the University Göttingen. This is publication 83 of the Courant Research Center Geobiology. References Allwood, A.C., Walter, M.R., Kamber, B.S., Marshall, C.P., Burch, I.W., 2006. Stromatolite reef from the Early Archaean era of Australia. Nature 441, 714–718. Almendros, G., Dorado, J., González-Vila, F.J., Martin, F., 1997. Pyrolysis of carbohydrate-derived macromolecules: its potential in monitoring the carbohydrate signature of geopolymers. Journal of Analytical and Applied Pyrolysis 40–41, 599–610. Anbar, A.D., Knoll, A.H., 2002. Proterozoic ocean chemistry and evolution: a bioinorganic bridge? Science 297, 1137–1142. Aquino Neto, F.R., Trendel, J.M., Restle, A., Connan, J., Albrecht, P., 1983. Occurrence and formation of tricyclic and tetracyclic terpanes in sediments and petroleums. In: Bjorøy, M. (Ed.), Advances in Organic Geochemistry 1981. John Wiley & Sons, New York, pp. 659–676. Arouri, K.R., Greenwood, P.F., Walter, M.R., 2000. Biological affinities of Neoproterozoic acritarchs from Australia: microscopic and chemical characterisation. Organic Geochemistry 31, 75–89. Berner, R.A., 1984. Sedimentary pyrite formation: an update. Geochimica et Cosmochimica Acta 48, 605–615. Berry, A., Harriott, O., Moreau, R., Osman, S., Benson, D., Jones, A., 1993. Hopanoid lipids compose the Frankia vesicle envelope, presumptive barrier of oxygen diffusion to nitrogenase. PNAS 90, 6091–6094. Bertrand-Sarfati, J., Moussine-Pouchkine, A., 1988. Is cratonic sedimentation consistent with available models? An example from the Upper Proterozoic of the West African craton. Sedimentary Geology 58, 255–276. Blumenberg, M., Mollenhauer, G., Zabel, M., Reimer, A., Thiel, V., 2010. Decoupling of bio- and geohopanoids in sediments of the Benguela Upwelling System (BUS). Organic Geochemistry 41, 1119–1129. Blumenberg, M., Seifert, R., Kasten, S., Bahlmann, E., Michaelis, W., 2009. Euphotic zone bacterioplankton sources major sedimentary bacteriohopanepolyols in the Holocene Black Sea. Geochimica et Cosmochimica Acta 73, 750–766. Boreham, C.J., Crick, I.H., Powell, T.G., 1988. Alternative calibration of the Methylphenanthrene Index against vitrinite reflectance: application to maturity measurements on oils and sediments. Organic Geochemistry 12, 289–294. Bravo, J.-M., Perzl, M., Härtner, T., Kannenberg, E.L., Rohmer, M., 2001. Novel methylated triterpenoids of the gammacerane series from the nitrogen-fixing bacterium Bradyrhizobium japonicum USDA 110. European Journal of Biochemistry 268, 1323–1331. Brocks, J.J., 2009. The succession of primary producers in Proterozoic oceans. Geochimica et Cosmochimica Acta 73, A161. 126 M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 Brocks, J.J., 2011. Millimeter-scale concentration gradients of hydrocarbons in Archean shales: live-oil escape or fingerprint of contamination? Geochimica et Cosmochimica Acta 75, 3196–3213. Brocks, J.J., Grosjean, E., Logan, G.A., 2008. Assessing biomarker syngeneity using branched alkanes with quaternary carbon (BAQCs) and other plastic contaminants. Geochimica et Cosmochimica Acta 72, 871–888. Brocks, J.J., Logan, G.A., Buick, R., Summons, R.E., 1999. Archean molecular fossils and the early rise of eukaryotes. Science 285, 1033–1036. Brocks, J.J., Buick, R., Logan, G.A., Summons, R.E., 2003a. Composition and syngeneity of molecular fossils from the 2.78 to 2.45 billion-year-old Mount Bruce Supergroup, Pilbara Craton, Western Australia. Geochimica et Cosmochimica Acta 67, 4289–4319. Brocks, J.J., Love, G.D., Snape, C.E., Logan, G.A., Summons, R.E., Buick, R., 2003b. Release of bound aromatic hydrocarbons from late Archean and Mesoproterozoic kerogens via hydropyrolysis. Geochimica et Cosmochimica Acta 67, 1521–1530. Brocks, J.J., Love, G.D., Summons, R., Knoll, A.H., Logan, G.A., Bowden, S.A., 2005. Biomarker evidence for green and purple sulphur bacteria in a stratified Palaeoproterozoic sea. Nature 437, 866–870. Brocks, J.J., Pearson, A., 2005. Building the biomarker tree of life. Reviews in Mineralogy and Geochemistry 59, 233–258. Brocks, J.J., Schaeffer, P., 2008. Okenane, a biomarker for purple sulfur bacteria (Chromatiaceae), and other new carotenoid derivatives from the 1640 Ma Barney Creek Formation. Geochimica et Cosmochimica Acta 72, 1396–1414. Brocks, J.J., Summons, R., 2003. Sedimentary hydrocarbons, biomarkers for early life. Treatise on Geochemistry 8, 63–115. Butterfield, N.J., 2000. Bangiomorpha pubescens n. gen., n. sp.: implications for the evolution of sex, multicellularity, and the Mesoproterozoic/Neoproterozoic radiation of eukaryotes. Paleobiology 26, 386–404. Canfield, D.E., 1998. A new model for Proterozoic ocean chemistry. Nature 396, 450–453. Canfield, D.E., 2005. The early history of atmospheric oxygen: homage to Robert M. Garrels. Annual Review of Earth and Planetary Sciences 33, 1–36. Delecat, S., Arp, G., Reitner, J., 2010. Aftermath of the Triassic-Jurassic boundary crisis: spiculite formation on drowned Triassic Steinplatte reef-slope by communities of hexactinellid sponges. Lecture Notes in Earth Sciences 131, 355–391. Doering, S., di Primio, R., Horsfield, B., 2009. New insights on the Precambrian petroleum system of the Taoudeni Basin based on shallow drilling results. In: 24th International Meeting on Organic Geochemistry, Elsevier, Bremen, p. 360. Doughty, D.M., Hunter, R.C., Summons, R.E., Newman, D.K., 2009. 2Methylhopanoids are maximally produced in akinetes of Nostoc punctiforme: geobiological implications. Geobiology 7, 1–9. Dutkiewicz, A., Volk, H., Ridley, J., George, S.C., 2004. Geochemistry of oil in fluid inclusions in a middle Proterozoic igneous intrusion: implications for the source of hydrocarbons in crystalline rocks. Organic Geochemistry 35, 937–957. Eigenbrode, J.L., Freeman, K.H., Summons, R.E., 2008. Methylhopane biomarker hydrocarbons in Hamersley Province sediments provide evidence for Neoarchean aerobiosis. Earth and Planetary Science Letters 273, 323–331. Fennel, K., Follows, M., Falkowski, P.G., 2005. The co-evolution of the nitrogen, carbon and oxygen cycles in the Proterozoic ocean. American Journal of Science 305, 526–545. Fowler, M.G., Douglas, A.G., 1987. Saturated hydrocarbon biomarkers in oils of Late Precambrian age from Eastern Siberia. Organic Letters 11, 201–213. Goodman, E.E., Kah, L.C., 2004. Trace sulfate concentrations as an indicator of depositional environment: examination of possible calcitized evaporates from the Proterozoic Atar Group, Mauritania. In: GSA Meeting, p. 78, Abstracts with Programs. Greenwood, P.F., Arouri, K.R., George, S.C., 2000. Tricyclic terpenoid composition of Tasmanites kerogen as determined by pyrolysis GC–MS. Geochimica et Cosmochimica Acta 64, 1249–1263. Greenwood, P.F., Arouri, K.R., Logan, G.A., Summons, R.E., 2004. Abundance and geochemical significance of C2n dialkylalkanes and highly branched C3n alkanes in diverse meso- and neoproterozoic sediments. Organic Geochemistry 35, 331–346. Grice, K., Schaeffer, P., Schwark, L., Maxwell, J.R., 1997. Changes in palaeoenvironmental conditions during deposition of the Permian Kupferschiefer (Lower Rhine Basin, northwest Germany) inferred from molecular and isotopic compositions of biomarker components. Organic Geochemistry 26, 677–690. Hartgers, W.A., Damsté, J.S.S., de Leeuw, J.W., 1994. Geochemical significance of alkylbenzene distributions in flash pyrolysates of kerogens, coals, and asphaltenes. Geochimica et Cosmochimica Acta 58, 1759–1775. Hayes, J.M., 1993. Factors controlling 13 C contents of sedimentary organic compounds: principles and evidence. Marine Geology 113, 115–125. Hayes, J.M., 2001. Fractionation of the isotopes of carbon and hydrogen in biosynthetic processes. In: Valley, J.W., Cole, D.R. (Eds.), Stable Isotope Geochemistry, Reviews in Mineralogy and Geochemistry. Mineralogical Society of America, Washington, DC, pp. 225–278. Huntley, J.W., Xiao, S., Kowalewski, M., 2006. 1.3 billion years of acritarch history: an empirical morphospace approach. Precambrian Research 144, 52–68. Imbus, S.W., Macko, S.A., Douglas Elmore, R., Engel, M.H., 1992. Stable isotope (C, S, N) and molecular studies on the Precambrian nonesuch Shale (WisconsinMichigan, U.S.A.): evidence for differential preservation rates, depositional environment and hydrothermal influence. Chemical Geology: Isotope Geoscience Section 101, 255–281. Javaux, E.J., Knoll, A.H., Walter, M.R., 2004. TEM evidence for eukaryotic diversity in mid-Proterozoic oceans. Geobiology 2, 121–132. Javaux, E.J., Marshall, C.P., Bekker, A., 2010. Organic-walled microfossils in 3.2-billion-year-old shallow-marine siliciclastic deposits. Nature 463, 934–938. Johnston, D.T., Wolfe-Simon, F., Pearson, A., Knoll, A.H., 2009. Anoxygenic photosynthesis modulated Proterozoic oxygen and sustained Earth’s middle age. Proceedings of the National Academy of Sciences 106, 16925–21692. Kah, L.C., Bartley, J.K., 2011. Protracted oxygenation of the Proterozoic biosphere. International Geology Review, 1424–1442. Kah, L.C., Bartley, J.K., Stagner, A.F., 2009. Reinterpreting a Proterozoic enigma: Conophyton–Jacutophyton stromatolites of the Mesoproterozoic Atar Group, Mauritania. IAS Special Publication 41, 277–295. Kah, L.C., Bartley, J.K., Teal, D.A., in review. Chemostratigraphy of the late Mesoproterozoic Atar Group, Mauritania: muted isotopic variability, facies correlation, and global isotopic trends. Precambrian Research. Kah, L.C., Lyons, T.W., Chesley, J.T., 2001. Geochemistry of a 1.2 Ga carbonateevaporite succession, northern Baffin and Bylot Islands: implications for Mesoproterozoic marine evolution. Precambrian Research 111, 203–234. Kah, L.C., Lyons, T.W., Frank, T.D., 2004. Low marine sulphate and protracted oxygenation of the Proterozoic biosphere. Nature 431, 834–838. Killops, S., Killops, V., 2005. Introduction to Organic Geochemistry, 2nd ed. Blackwell publishing, Oxford. Klotz, M.G., Stein, L.Y., 2008. Nitrifier genomics and evolution of the nitrogen cycle. FEMS Microbiology Letters 278, 146–156. Knoll, A.H., Javaux, E.J., Hewitt, D., Cohen, P., 2006. Eukaryotic organisms in Proterozoic oceans. Philosophical Transactions of the Royal Society B: Biological Sciences 361, 1023–1038. Koopmans, M.P., Koster, J., Van Kaam-Peters, H.M.E., Kenig, F., Schouten, S., Hartgers, W.A., de Leeuw, J.W., Sinninghe Damste, J.S., 1996a. Diagenetic and catagenetic products of isorenieratene: molecular indicators for photic zone anoxia. Geochimica et Cosmochimica Acta 60, 4467–4496. Koopmans, M.P., Schouten, S., Kohnen, M.E.L., Sinninghe Damsté, J.S., 1996b. Restricted utility of aryl isoprenoids as indicators for photic zone anoxia. Geochimica et Cosmochimica Acta 60, 4873–4876. Köster, J., Schouten, S., Sinninghe Damsté, J.S., De Leeuw, J.W., 1995. Reconstruction of depositional environment of Toarcian marlstones (Allgäu Formation, Tyrol/Austria) using biomarkers and compound specific carbon isotope analyses. In: Grimalt, J.O., Dorronsoro, C. (Eds.), Organic Geochemistry: Developments and Applications of Energy, Climate, Environment and Human History. A.I.G.O.A., San Sebastian. Logan, G.A., Calver, C.R., Gorjan, P., Summons, R.E., Hayes, J.M., Walter, M.R., 1999. Terminal Proterozoic mid-shelf benthic microbial mats in the Centralian Superbasin and their environmental significance. Geochimica et Cosmochimica Acta 63, 1345–1358. Logan, G.A., Hayes, J.M., Hieshima, G.B., Summons, R.E., 1995. Terminal Proterozoic reorganization of biogeochemical cycles. Nature 376, 53–56. Logan, G.A., Hinman, M.C., Walter, M.R., Summons, R.E., 2001. Biogeochemistry of the 1640 Ma McArthur River (HYC) lead-zinc ore and host sediments, Northern Territory, Australia. Geochimica et Cosmochimica Acta 65, 2317– 2336. Lyons, T.W., Anbar, A.D., Severmann, S., Scott, C., Gill, B.J., 2009. Tracking euxinia in the ancient acean: a multiproxy perspective and Proterozoic case study. Annual Review of Earth and Planetary Sciences 37, 507–534. Marshall, C.P., Javaux, E.J., Knoll, A.H., Walter, M.R., 2005. Combined micro-Fourier transform infrared (FTIR) spectroscopy and micro-Raman spectroscopy of Proterozoic acritarchs: a new approach to Palaeobiology. Precambrian Research 138, 208–224. Marshall, C.P., Love, G.D., Snape, C.E., Hill, A.C., Allwood, A.C., Walter, M.R., Van Kranendonk, M.J., Bowden, S.A., Sylva, S.P., Summons, R.E., 2007. Structural characterization of kerogen in 3.4 Ga Archaean cherts from the Pilbara Craton, Western Australia. Precambrian Research 155, 1–23. Marynowski, L., Czechowski, F., Simoneit, B.R.T., 2001. Phenylnaphthalenes and polyphenyls in Palaeozoic source rocks of the Holy Cross Mountains, Poland. Organic Geochemistry 32, 69–85. Moldowan, J.M., Talyzina, N.M., 1998. Biogeochemical evidence for dinoflagellate ancestors in the Early Cambrian. Science 281, 1168–1170. Moreau, R.A., Powell, M.J., Fett, W.F., Whitaker, B.D., 1997. The effect of ethanol and oxygen on the growth of Zymomonas mobilis and the levels of hopanoids and other membrane lipids. Current Microbiology 35, 124–128. Mycke, B., Michaelis, W., Degens, E.T., 1988. Biomarkers in sedimentary sulfides of Precambrian age. Organic Geochemistry 13, 619–625. Ohkouchi, N., Nakajima, Y., Okada, H., Ogawa, N.O., Suga, H., Oguri, K., Kitazato, H., 2005. Biogeochemical processes in the saline meromictic Lake Kaiike, Japan: implications from molecular isotopic evidences of photosynthetic pigments. Environmental Microbiology 7, 1009–1016. Pearson, A., Flood Page, S.R., Jorgenson, T.L., Fischer, W.W., Higgins, M.B., 2007. Novel hopanoid cyclases from the environment. Environmental Microbiology 9, 2175–2180. Peters, K.E., Moldowan, J.M., 1993. The Biomarker Guide—Interpreting Molecular Fossils in Petroleum and Ancient Sediments. Prentice Hall, Englewood Cliffs, New Jersey, p. 363. Peters, K.E., Walters, C.C., Moldowan, J.M., 2004a. The Biomarker Guide: Biomarkers and Isotopes in Petroleum Exploration and Earth History, vol. 1. The Press Syndicate of the University of Cambridge, Cambridge. Peters, K.E., Walters, C.C., Moldowan, J.M., 2004b. The Biomarker Guide: Biomarkers and Isotopes in the Environment and Human History, vol. 2. The Press Syndicate of the University of Cambridge, Cambridge. M. Blumenberg et al. / Precambrian Research 196–197 (2012) 113–127 Pratt, L.M., Summons, R.E., Hieshima, G.B., 1991. Sterane and triterpane biomarkers in the Precambrian Nonesuch Formation, North American Midcontinent Rift. Geochimica et Cosmochimica Acta 55, 911–916. Pruss, S.B., Bosak, T., Macdonald, F.A., McLane, M., Hoffman, P.F., 2010. Microbial facies in a Sturtian cap carbonate, the Rasthof Formation, Otavi Group, northern Namibia. Precambrian Research 181, 187–198. Radke, M., Welte, D.H., 1983. The Methylphenanthrene Index (MPI): a maturity parameter based on aromatic hydrocarbons. Advances in Organic Geochemistry 1981, 504–512. Raiswell, R., Fisher, Q.J., 2000. Mudrock-hosted carbonate concretions: a review of growth mechanisms and their influence on chemical and isotopic composition. Journal of the Geological Society, London 157, 239–251. Rashby, S.E., Sessions, A.L., Summons, R.E., Newman, D.K., 2007. Biosynthesis of 2methylbacteriohopanepolyols by an anoxygenic phototroph. Proceedings of the National Academy of Sciences 104, 15099–21510. Rasmussen, B., Fletcher, I.R., Brocks, J.J., Kilburn, M.R., 2008. Reassessing the first appearance of eukaryotes and cyanobacteria. Nature 455, 1101–1104. Reitner, J., Peckmann, J., Reimer, A., Schumann, G., Thiel, V., 2005. Methane-derived carbonate build-ups and associated microbial communities at cold seeps on the lower Crimean shelf (Black Sea). Facies 51, 71–84. Rohmer, M., Bouvier-Navé, P., Ourisson, G., 1984. Distribution of hopanoid triterpenes in prokaryotes. Journal of General Microbiology 130, 1137–1150. Rooney, A.D., Selby, D., Houzay, J.-P., Renne, P.R., 2010. Re-Os geochronology of a Mesoproterozoic sedimenatary succession, Taoudeni basin, Mauritania: implications for basin-wide correlation and Re-Os organic rich sediments systematics. Earth and Planetary Science Letters 289, 486–496. Sakata, S., Hayes, J.M., McTaggart, A.R., Evans, R.A., Leckrone, K.J., Togasaki, R.K., 1997. Carbon isotopic fractionation associated with lipid biosynthesis by a cyanobacterium: relevance for interpretation of biomarker records. Geochimica et Cosmochimica Acta 61, 5379–53889. Sánchez-Beristain, F., Schäfer, N., Simon, K., Reitner, J., 2010. New geochemical method to characterise mirobialites from the St. Cassian Formation, Dolomites, Northeastern Italy. Lecture Notes in Earth Sciences 131, 435–453. Schaeffer, P., Adam, P., Wehrung, P., Albrecht, A., 1997. Novel aromatic carotenoid derivatives from sulfur photosynthetic bacteria in sediments. Tetrahedron Letters 38, 8413–8416. Schidlowski, M., Gorzawksi, H., Dor, I.A., 1994. Carbon isotope variations in a solar pond microbial mat: role of environmental gradients as steering variables. Geochimica et Cosmochimica Acta 58, 2289–2298. Schwark, L., Frimmel, A., 2004. Chemostratigraphy of the Posidonia Black Shale, SWGermany II. Assessment of extent and persistence of photic-zone anoxia using aryl isoprenoid distributions. Chemical Geology 206, 231–248. Seifert, W.K., Michael Moldowan, J., 1978. Applications of steranes, terpanes and monoaromatics to the maturation, migration and source of crude oils. Geochimica et Cosmochimica Acta 42, 77–95. Shen, Y., Canfield, D.E., Knoll, A.H., 2002. Middle Proterozoic ocean chemistry: evidence from the McArthur Basin, northern Australia. American Journal of Science 302, 81–109. Sinninghe Damste, J.S., Schouten, S., van Duin, A.C.T., 2001. Isorenieratene derivatives in sediments: possible controls on their distribution. Geochimica et Cosmochimica Acta 65, 1557–1571. 127 Staal, M., Thar, R., K.ühl, M., van Loosdrecht, M.C.M., Wolf, G., de Brouwer, J.F.C., Rijstenbil, J.W., 2007. Different carbon isotope fractionation patterns during the development of phototrophic freshwater and marine biofilms. Biogeosciences 4, 613–626. Summons, R.E., Brassell, S.C., Eglinton, G., Evans, E., Horodyski, R.J., Robinson, N., Ward, D.M., 1988a. Distinctive hydrocarbon biomarkers from fossiliferous sediment of the Late Proterozoic Walcott Member, Chuar Group, Grand Canyon, Arizona. Geochimica et Cosmochimica Acta 52, 2625–2637. Summons, R.E., Jahnke, L.L., Hope, J.M., Logan, G.A., 1999. 2-Methylhopanoids as biomarkers for cyanobacterial oxygenic photosynthesis. Nature 400, 554–557. Summons, R.E., Powell, T.G., Boreham, C.J., 1988b. Petroleum geology and geochemistry of the Middle Proterozoic McArthur Basin, Northern Australia. III. Composition of extractable hydrocarbons. Geochimica et Cosmochimica Acta 52, 1747–1763. Summons, R.E., Powell, T.G., 1987. Identification of aryl isoprenoids in source rocks and crude oils: biological markers for the green sulphur bacteria. Geochimica et Cosmochimica Acta 51, 557–566. Summons, R.E., Volkman, J.K., Boreham, C.J., 1987. Dinosterane and other steroidal hydrocarbons of dinoflagellate origin in sediments and petroleum. Geochimica et Cosmochimica Acta 51, 3075–3082. Talbot, H.M., Summons, R.E., Jahnke, L.L., Cockell, C.S., Rohmer, M., Farrimond, P., 2008. Cyanobacterial bacteriohopanepolyol signatures from cultures and natural environmental settings. Organic Geochemistry 39, 232–263. Teal D.A.J., Kah L.C., 2005. Using C-isotopes to Constrain Intrabasinal Stratigraphic Correlations: Mesoproterozoic Atar Group, Mauritania, Geological Society of America (GSA), p. 45. ten Haven, H.L., De Leeuw, J.W., Rullkötter, J., Sinninghe Damste, J.S., 1987. Restricted utility of the pristane/phytane ratio as a palaeoenvironmental indicator. Nature 330, 641–643. Thompson, K.F.M., 1987. Fractionated aromatic petroleums and the generation of gas-condensates. Organic Geochemistry 11, 573–590. Waldbauer, J.R., Sherman, L.S., Sumner, D.Y., Summons, R.E., 2009. Late Archean molecular fossils from the Transvaal Supergroup record the antiquity of microbial diversity and aerobiosis. Precambrian Research 169, 28–47. Welander, P.V., Coleman, M.L., Sessions, A.L., Summons, R.E., Newman, D.K., 2010. Identification of a methylase required for 2-methylhopanoid production and implications for the interpretation of sedimentary hopanes. Proceedings of the National Academy of Sciences 107, 8537–8854. Welander, P.V., Hunter, R.C., Zhang, L., Sessions, A.L., Summons, R.E., Newman, D.K., 2009. Hopanoids play a role in membrane integrity and pH homeostasis in Rhodopseudomonas palustris TIE-1. Journal of Bacteriology. 191, 6145–6156. Wieland, A., Pape, T., Möbius, J., Klock, J.H., Michaelis, W., 2008. Carbon pools and isotopic trends in a hypersaline cyanobacterial mat. Geobiology 6, 171–186. Zerkle, A.L., House, C.H., Cox, R.P., Canfield, D.E., 2006. Metal limitation of cyanobacterial N2 fixation and implications for the Precambrian nitrogen cycle. Geobiology 4, 285–297. Zundel, M., Rohmer, M., 1985. Prokaryotic triterpenoids. 3. The biosynthesis of 2bmethylhopanoids and 3b-methylhopanoids of Methylobacterium organophilum and Acetobacter pasteurianus ssp. pasteurianus. European Journal of Biochemistry 150, 35–39.
© Copyright 2025 Paperzz