Biogeosciences, 11, 7009–7023, 2014 www.biogeosciences.net/11/7009/2014/ doi:10.5194/bg-11-7009-2014 © Author(s) 2014. CC Attribution 3.0 License. Biomarkers in the stratified water column of the Landsort Deep (Baltic Sea) C. Berndmeyer1, V. Thiel1, O. Schmale2, N. Wasmund2, and M. Blumenberg1,* 1Geobiology Group, Geoscience Center, Georg August University Göttingen, Goldschmidtstr. 3, 37077 Göttingen, Germany 2Leibniz Institute for Baltic Sea Research Warnemünde (IOW), Seestr. 15, 18199 Rostock-Warnemünde, Germany *now at: Federal Institute for Geosciences and Natural Resources (BGR), Stilleweg 2, 30655 Hanover, Germany Correspondence to: C. Berndmeyer (christine.berndmeyer@geo.uni-goettingen.de) Received: 28 May 2014 – Published in Biogeosciences Discuss.: 25 June 2014 Revised: 28 October 2014 – Accepted: 29 October 2014 – Published: 11 December 2014 Abstract. The water column of the Landsort Deep, central Baltic Sea, is stratified into an oxic, suboxic, and anoxic zone. This stratification controls the distributions of individ- ual microbial communities and biogeochemical processes. In summer 2011, particulate organic matter was filtered from these zones using an in situ pump. Lipid biomark- ers were extracted from the filters to establish water-column profiles of individual hydrocarbons, alcohols, phospholipid fatty acids, and bacteriohopanepolyols (BHPs). As a ref- erence, a cyanobacterial bloom sampled in summer 2012 in the central Baltic Sea Gotland Deep was analyzed for BHPs. The biomarker data from the surface layer of the oxic zone showed major inputs from cyanobacteria, dinoflagel- lates, and ciliates, while the underlying cold winter water layer was characterized by a low diversity and abundance of organisms, with copepods as a major group. The suboxic zone supported bacterivorous ciliates, type I aerobic methan- otrophic bacteria, sulfate-reducing bacteria, and, most likely, methanogenic archaea. In the anoxic zone, sulfate reducers and archaea were the dominating microorganisms as indi- cated by the presence of distinctive branched fatty acids: archaeol and pentamethylicosane (PMI) derivatives, respec- tively. Our study of in situ biomarkers in the Landsort Deep thus provided an integrated insight into the distribution of rel- evant compounds and describes useful tracers to reconstruct stratified water columns in the geological record. 1 Introduction The Baltic Sea is a brackish marine marginal sea with a max- imum depth of 459 m in the Landsort Deep (western central Baltic Sea; Matthäus and Schinke, 1999; Reissmann et al., 2009; Fig. 1). A positive freshwater budget and saltwater in- flows from the North Sea through Skagerrak and Kattegat lead to a permanent halocline that stratifies the water column of the central Baltic Sea at about 60 m water depth (Reiss- mann et al., 2009). Major saltwater inflows, as detected in 1993 and 2003, sporadically disturb the stratification in the eastern central Baltic Sea and oxygenate the suboxic zone and deep water. These inflows, however, rarely reach the western central Baltic Sea. Even the strong inflow from 1993 had only minor effects on Landsort Deep, where stagnating conditions prevailed throughout (Bergström and Matthäus, 1996). Therefore, the Landsort Deep offers stable environ- ments for microbial life within the oxic, suboxic, and anoxic zones and provides an excellent study site for the investi- gation of biomarker inventories that specify stratified water columns. The Black Sea, although much larger in size, is compara- ble with the Landsort Deep with respect to the existence of a permanently anoxic deep-water body. Two comprehensive in situ biomarker reports gave a wide-ranging overview of var- ious biomarkers and their producers in the Black Sea water column and identified a close coupling of microorganisms to biogeochemically defined water layers (Wakeham et al., 2007, 2012). Several other in situ biomarker water-column studies exist, but they were usually focused on certain as- pects, for example anaerobic and aerobic methanotrophy Published by Copernicus Publications on behalf of the European Geosciences Union. 7010 C. Berndmeyer et al.: Biomarkers in the stratified water column Figure 1. Map showing the sampling locations in the central Baltic Sea. (Schouten et al., 2001; Schubert et al., 2006; Blumenberg et al., 2007; Sáenz et al., 2011; Xie et al., 2014, and others). For the Baltic Sea water column, biomarker knowledge is limited as most studies so far were focused on pollution- related compounds (e.g., Beliaeff and Burgeot, 2001; Lehto- nen et al., 2006; Hanson et al., 2009). Recently, we reported the water-column distributions and 13C-content of individual bacteriohopanepolyols (BHPs) and phospholipid fatty acids (PLFA) from the Gotland Deep, located about 150 km SE of the Landsort Deep in the eastern central Baltic Sea. These studies were aimed at microbial methane turnover and con- firmed the importance of the Baltic Sea suboxic zone for bac- terial methane oxidation (Schmale et al., 2012; Berndmeyer et al., 2013; Jakobs et al., 2014). The theoretical possibility of sulfate-dependent methane oxidation in the anoxic zone was also stated (Jakobs et al., 2014) but still remains to be proven for the central Baltic Sea water column. Because the eastern central Baltic Sea is regularly dis- turbed by lateral intrusions in intermediate water depths (Jakobs et al., 2013), we chose the more stable Landsort Deep in the western central Baltic Sea as a sampling site for this biomarker study. Furthermore, published genetic studies reporting on prokaryotes and the related metabolisms in the water column of the Landsort Deep (Labrenz et al., 2007; Thureborn et al., 2013) provide a background to which the organic geochemical results can be advantageously related. The depth profiles of biomarkers from this setting not only reveal how actual biogeochemical processes are reflected by lipid abundances, distributions, and stable carbon isotope sig- natures, they also provide reference data for the reconstruc- tion of past water columns using biomarkers from the sedi- mentary record. 2 Materials and methods 2.1 Samples Samples were taken during cruise 06EZ/11/05 of R/V Elisa- beth Mann Borghese in summer 2011. The Landsort Deep is located north of Gotland (58◦35.0′ N, 18◦14.0′ E; Fig. 1). A Seabird sbe911+ CTD (conductivity, temperature, density) system and a turbidity sensor ECO FLNTU (WET Labs) were used for continuous water-column profiling. Oxygen and hydrogen sulfide concentrations were measured colomet- rically with Winkler’s method, respectively (Grasshoff et al., 1983). Filter samples of 65 to 195 L obtained from 10, 65, 70, 80, 90, 95, and 420 m water depth were taken with an in situ pump, and particulate material was filtered onto precom- busted glass microfiber filters (∅ 30 cm; 0.7 µm pore size; Munktell & Filtrak GmbH, Germany). Filters were freeze- dried and kept frozen at −20 ◦C until analysis. A cyanobacterial bloom was sampled in summer 2012 on cruise M87/4 of R/V Meteor at the Gotland Deep (57◦19.2′ N, 20◦03.0′ E; Fig. 1), east of Gotland. Water sam- ples of 10 L were taken at 1 m water depth and filtered with a 20 µm net. The samples were centrifuged and the residue freeze-dried. Samples were kept frozen at −20 ◦C until anal- ysis. 2.2 Bulk CNS analysis Three pieces (∅ 1.2 cm) from different zones of the filters were combusted together with V2O5 in a EuroVector Eu- roEA Elemental Analyzer. Particulate matter in the Baltic Sea was reported to be free of carbonate (Schneider et al., 2002), and, thus, the filters were not acidified prior to analy- sis. C, N, and S contents were calculated by comparison with peak areas from standards. Standard deviations were ±2 % for C and ±5 % for N and S. 2.3 Lipid analysis Three quarters of each filter were extracted (3× 20 min) with dichloromethane (DCM) /methanol (MeOH) (40 mL; 3 : 1, v : v) in a CEM Mars 5 microwave (Matthews, NC, USA) at 60 ◦C and 800 W. The freeze-dried residue of the cyanobac- terial bloom was extracted (3× 10 min) with DCM /MeOH (10 mL; 3 : 1, v : v) and ultrasonication. All extracts were combined. An aliquot of each filter extract and the bloom extract was acetylated using Ac2O and pyridine (1 : 1, v : v) for 1 h at 50 ◦C and then overnight at room temperature. The mixture was dried under vacuum and analyzed for BHPs using liquid- chromatography–mass-spectrometry (LC-MS). Another aliquot of each filter extract was separated into a hydrocarbon (F1), an alcohol and ketone (F2), and a polar fraction (F3) using column chromatography. The column (∅ ca. 1 cm) was filled with 7.5 g silica gel 60; samples were dried on ca. 500 mg silica gel 60 and placed on the column. Biogeosciences, 11, 7009–7023, 2014 www.biogeosciences.net/11/7009/2014/ C. Berndmeyer et al.: Biomarkers in the stratified water column 7011 The fractions were eluted with 30 mL n-hexane /DCM 8 : 2 (v : v, F1), 30 mL DCM /EtOAC 9 : 1 (v : v, F2), and 100 mL DCM /MeOH 1 : 1, (v : v) followed by an additional 100 mL MeOH (F3). F2 was dried and derivatized using a BSTFA / pyridine 3 : 2 (v : v) mixture for 1 h at 40 ◦C. A to- tal of 50 % of the polar fraction F3 was further fractionated to obtain PLFA (F3.3) according to Sturt et al. (2004). Briefly, this entails filling the column with 2 g silica gel 60 and stor- ing it at 200 ◦C until use. The F3 aliquot was dried on ca. 500 mg silica gel 60 and placed on the column. After suc- cessive elution of the column with 15 mL DCM and 15 mL acetone, the PLFA fraction was eluted with 15 mL MeOH (F3.3). F3.3 was transesterified using trimethylchlorosilane (TMCS) in MeOH (1 : 9; v : v) for 1 h at 80 ◦C. In the re- sulting fatty acid methyl ester (FAME) fractions, double- bond positions in monounsaturated compounds were deter- mined using dimethyldisulfide (DMDS; Carlson et al., 1989; Gatellier et al., 1993). The samples were dissolved in 200 µL DMDS, 100 µL n-hexane, and 30 µL I2 solution (60 mg I2 in 1 mL Et2O) and derivatized at 50 ◦C for 48 h. Subsequently, 1 mL of n-hexane and 200 µL of NaHSO4 (5 % in water) were added and the n-hexane extract was pipetted off. The procedure was repeated 3 times, and the n-hexane extracts were combined, dried on ca. 500 mg silica gel 60 and put onto a small column (ca. 1 g silica gel 60). For cleaning, the n-hexane extract was eluted with 10 dead volumes of DCM. F1, F2, F3.3, and the samples treated with DMDS were an- alyzed using gas-chromatography–mass-spectrometry (GC- MS). 2.4 Gas-chromatography–mass-spectrometry (GC-MS) and GC-combustion isotope ratio mass spectrometry (GC-C-IRMS) GC-MS was performed using a Varian CP-3800 chromato- graph equipped with a Phenomenex Zebron ZB-5MS fused silica column (30 m× 0.32 mm; film thickness 0.25 µm) cou- pled to a Varian 1200L mass spectrometer. Helium was used as a carrier gas. The temperature program started at 80 ◦C (3 min) and ramped up to 310 ◦C (held 25 min) with 4 ◦C min−1. Compounds were assigned by comparing mass spectra and retention times to published data. Concentrations were determined by comparison with peak areas of squalane (F2 and F3) and n-eicosane-D42 (F1) as internal standards. Compound-specific stable carbon isotope ratios of biomarkers in F2 and F3.3 were measured (twice) using a Thermo Trace gas chromatograph coupled to a Thermo Delta Plus isotope ratio mass spectrometer. The GC was oper- ated under the same conditions and with the same column as for GC-MS. The combustion reactor contained CuO, Ni, and Pt and was operated at 940 ◦C. Isotopic compositions are reported in standard delta notation relative to the Vienna PeeDee Belemnite (V-PDB) and were calculated by compar- ison with an isotopically known CO2 reference gas. GC-C- IRMS precision and linearity was checked daily using an ex- ternal n-alkane isotopic standard (provided by A. Schimmel- mann, Indiana University). 2.5 Liquid-chromatography–mass-spectrometry (LC-MS) LC-MS was performed using a Varian Prostar Dynamax high-performance liquid chromatography (HPLC) system fit- ted with a Merck Lichrocart (Lichrosphere 100; reversed phase (RP) C18e) column (250× 4 mm) and a Merck Lichro- sphere pre-column of the same material coupled to a Varian 1200L triple quadrupole mass spectrometer (both Varian). The solvents used were MeOH /water 9 : 1 (v : v; solvent A) and MeOH / propan-2-ol 1 : 1 (v : v; solvent B), and all sol- vents were Fisher Scientific HPLC grade. The solvent gradi- ent profile was 100 % A (0–1 min) to 100 % B at 35 min and then isocratic to 60 min. The MS was equipped with an at- mospheric pressure chemical ionization (APCI) source oper- ated in positive/ion mode (capillary temperature 150 ◦C, va- porizer temperature 400 ◦C, corona discharge current 8 µA, nebulizing gas flow 70 psi, auxiliary gas 17 psi). In SIM (single-ion monitoring) mode, ions obtained from acetylated BHP peaks in the samples were compared to authentic BHP standards with known concentrations (acetylated BHP and aminotriol) to determine BHP concentrations (external cal- ibration). Amino BHPs had a 7 times higher response fac- tor than nonamino BHPs, and concentrations in the samples were corrected accordingly. Comparisons with elution times of previously identified compounds further aided in BHP as- signment. The quantification error is estimated to be ±20 %. 2.6 Principle component analysis (PCA) PCA was based on the relative abundance of individual com- ponents in different water depths and was performed using R (version 3.0.2, 25 September 2013) with the “princomp” module (The R Foundation, 2014). 3 Results 3.1 Physicochemical parameters of the water column In summer 2011, the Landsort Deep showed a strong vertical stratification (Fig. 2). The oxic zone consisted of the upper- most 80 m and was divided by a strong thermocline into a warm surface layer (∼ 0–10 m) and a cold winter water layer (∼ 10–60 m). The halocline was located between 60 and 80 m. O2 concentrations rapidly decreased from > 8 mL L−1 at ∼ 50 m to < 0.2 mL L−1 at ∼ 80 m, defining the upper boundary of the suboxic zone (Tyson and Pearson, 1991). H2S was first detected at 83 m. Because O2 concentrations could methodologically only be measured in the complete absence of H2S, oxygen could not be traced below this depth. Therefore, the lower boundary of the suboxic zone was defined to be at 90 m, where H2S concentrations sharply www.biogeosciences.net/11/7009/2014/ Biogeosciences, 11, 7009–7023, 2014 7012 C. Berndmeyer et al.: Biomarkers in the stratified water column Figure 2. Physicochemical characteristics of the Landsort Deep wa- ter column in summer 2011. The suboxic zone is shaded light grey. Temperature and methane data were partially taken from Jakobs et al. (2014). increased. The upper suboxic zone also showed a sharp peak in turbidity that is possibly caused by the precipitation of Fe and Mn oxides (Dellwig et al., 2010) or zero-valent sulfur (Kamyshny Jr. et al., 2013) and can be used as an indica- tor of the O2-H2S transition (Kamyshny Jr. et al., 2013). The anoxic zone extends from 90 m to the bottom and is charac- terized by the complete absence of O2 and high concentra- tions of H2S and CH4. CH4 was highest in the deep anoxic zone and decreased strongly towards the suboxic zone but was still present in minor concentrations in the oxic zone. A small CH4 peak was detected at the suboxic–anoxic interface (Fig. 2). Particulate organic carbon (POC) was highest at 10 m (380 µg L−1), decreased to a minimum in the cold winter wa- ter layer (48 µg L−1), and showed almost constant values of ∼ 70 µg L−1 in the suboxic and anoxic zones. Generally, we follow the zonation of the Landsort Deep water column as given in Jakobs et al. (2014). We regarded the onset of H2S as the top of the anoxic zone, however, as this is better supported by our biomarker data (see below). 3.2 Lipid analysis The PCA analysis separated six groups of biomarkers ac- cording to their distribution in the water column (Fig. 3, Sect. 3.2.1-6). Out of these groups, 18 compounds were se- lected as representative biomarkers, specifying inputs from individual prokaryotes and eukaryotes (with phototrophic, chemotrophic, and/or heterotrophic metabolisms). These biomarkers and their distributions are discussed in detail in Sect. 4. The concentrations of these compounds are shown in Fig. 4, and compound-specific δ13C values are given in Ta- ble 1. Apart from the biomarker families revealed by PCA, two compound classes – n-alkanes and n-alkenes in the sea surface layer – and individual BHPs obtained from the water Table 1. δ13C values of the compounds chosen from the PCA groups. No δ13C values were available for Group 4. N.d. stands for “not detectable”. δ13C [‰] Compound oxic zone suboxic zone anoxic zone Group 1 7-Me-17 : 0 alkane n.d. n.d. n.d. β-Sitosterol −29.9 n.d. −30.1 20 : 4ω6 PLFA −30.1 −31.7 −31.6 20 : 5ω3 PLFA −29.2 n.d. n.d. 16 : 1ω7c PLFA −30.6 −28.0 −28.3 Cholesterol −26.8 −28.9 −31.7 Group 2 Dinosterol −29.9 −30.9 −32.0 Tetrahymanol −28.7 −27.9 −25.9 Ai-15 : 0 PLFA −29.3 −32.5 −34.2 Diploptene n.d. n.d. n.d. Group 3 16 : 0–18 : 1 Wax ester −28.1 −28.2 n.d. Group 5 16 : 1ω8 PLFA n.d. −45.4 n.d. Group 6 Cholestanol −27.8 −28.9 −30.1 10-Me-16 : 0 PLFA n.d. −32.5 −35.4 PMI+PMI 1 n.d. n.d. n.d. Archaeol n.d. n.d. – column and a cyanobacterial bloom are reported separately (Fig. 5, Sect. 3.2.7; Fig. 6a, Sect. 3.2.8, respectively). 3.2.1 Group 1: surface maximum The first group is defined by a strong maximum in the sur- face layer and only minor concentrations in greater depths. A subgroup of 14 compounds exclusively occurs at 10 m water depth (Fig. 3). For the other compounds, abun- dance in greater water depths increases towards the y axis. 7-methylheptadecane (52), 24-ethylcholest-5-en-3β-ol (β- sitosterol; 48), 20 : 4ω6 PLFA (34), 20 : 5ω3 PLFA (33), 16 : 1ω7c PLFA (11), and cholest-5-en-3β-ol (cholesterol; 44) were taken as representative for Group 1. Among these compounds, 16 : 1ω7 PLFA and cholesterol showed the high- est concentrations (1154 and 594 ng L−1, respectively) and 7- methylheptadecane the lowest (6 ng L−1, Fig. 4). Apart from their maximum in the surface layer, the fate of these biomark- ers in deeper water layers differed. 7-methylheptadecane ex- clusively occurred in the surface layer, whereas 20 : 4ω6 was traceable throughout the water column. β-sitosterol occurred in the surface and the bottom layers. Unlike the other com- pounds, cholesterol and 20 : 5ω3 PLFA did not show a linear decrease with depth; rather, there are minor occurrences right Biogeosciences, 11, 7009–7023, 2014 www.biogeosciences.net/11/7009/2014/ C. Berndmeyer et al.: Biomarkers in the stratified water column 7013 Table 2. Compounds sorted by number as shown in Fig. 3. Compounds chosen for further discussion are marked bold. Compounds: 1 13 : 0 PLFA 22 18 : 4 PLFA 43 26 : 0 PLFA 64 n-C22:1 2 i14 : 0 PLFA 23 18 : 2 PLFA 44 cholesterol 65 n-C22 : 0 3 14 : 0 PLFA 24 18 : 3 PLFA 45 cholestanol 66 n-C23 : 1 4 i15 : 0 PLFA 25 18 : 1ω9c PLFA 46 16 : 0-18.1 wax ester 67 n-C23 : 0 5 ai15 : 0 PLFA 26 18 : 1ω7c PLFA 47 18 : 0− 18 : 1 wax ester 68 n-C24 : 1 6 15 : 0 PLFA 27 18 : 1ω6c PLFA 48 β-Sitosterol 69 n-C24 : 0 7 16 : 4 PLFA 28 18 : 1ω5c PLFA 49 dinosterol 70 n-C25 : 1 8 i 16 : 0 PLFA 29 18 : 0 PLFA 50 tetrahymanol 71 n-C25 : 0 9 16 : 1ω9c PLFA 30 10-me-18 : 0 PLFA 51 archaeol 72 n-C26 : 1 10 16 : 1ω8c PLFA 31 iC19 : 0 PLFA 52 7-methylheptadecane 73 n-C26 : 0 11 16 : 1ω7c PLFA 32 19 : 0 PLFA 53 PMI +PMI D 74 n-C27 : 0 12 16 : 1ω7t PLFA 33 20 : 5ω3 PLFA 54 diploptene 75 n-C28 : 0 13 16 : 1ω5c PLFA 34 20 : 4ω6 PLFA 55 n-C17 : 1 76 n-C29 : 0 14 16 : 1ω5t PLFA 35 20 : 3 PLFA 56 n-C17 : 0 77 n-C30 : 0 15 16 : 0 PLFA 36 20 : 3 PLFA 57 n-C18 : 0 78 n-C31 : 0 16 10-me-16 : 0 PLFA 37 20 : 1 PLFA 58 n-C19 : 1 79 n-C32 : 0 17 iC17 : 0 PLFA 38 20 : 0 PLFA 59 n-C19 : 0 80 n-C33 : 0 18 ai C17 : 0 PLFA 39 22 : 6 PLFA 60 n-C20 : 1 81 n-C34 : 0 19 17 : 1 PLFA 40 22 : 4 PLFA 61 n-C20 : 0 82 n-C35 : 0 20 17 : 0 PLFA 41 22 : 0 PLFA 62 n-C21 : 1 83 n-C36 : 0 21 18 : 4 PLFA 42 24 : 0 PLFA 63 n-C21 : 0 84 total BHPs above and at the bottom of the suboxic zone. δ13C values of all compounds were between −32 and −26 ‰ (Table 1). 3.2.2 Group 2: surface and lower suboxic zone maxima Like Group 1, Group 2 shows a surface maximum, but it exhibits a stronger emphasis on the lower suboxic zone (Fig. 4). With the exception of 16 : 7ω7t, all compounds were chosen for further consideration. 4α, 23, 24-trimethyl- 5α-cholest-22E-en-3β-ol (dinosterol; 49), and gammacer- 3β-ol (tetrahymanol; 50) had their maximum concentration in the surface water (dinosterol: 66 ng L−1; tetrahymanol: 42 ng L−1) and were not detectable in the layers below un- til a sharp second maximum occurred at the bottom of the suboxic zone. Concentrations decreased again below the sub- oxic zone and remained constantly low in the bottom water. Unlike these compounds, ai 15 : 0 PLFA (5), total bacteri- ohopanepolyols (BHPs; 84), and the hopanoid hydrocarbon hop-22(29)-ene (diploptene; 54) showed steadily increasing concentrations throughout the suboxic zone and further in- creasing concentrations in the anoxic zone. The δ13C values of all compounds were between −35 and −25 ‰ (Table 1). 3.2.3 Group 3: cold winter water layer maximum The third group showed compounds that peaked in the cold winter water layer at 65 m water depth (Fig. 3). 17 : 1ω9 PLFA (19) only occurred at 70 m water depth, and n-C21 (61) occurred from 10 to 70 m, with a strong peak at 70 m. The 16 : 0− 18 : 1 (46; Fig. 4) and 18 : 0− 18 : 1 (47) wax esters Figure 3. PCA of the relative abundances of compounds in differ- ent water depths. Group 1: surface maximum; a subgroup of com- pounds exclusively occurring at the surface are listed in the box. Group 2: surface and lower suboxic zone maxima. Group 3: cold winter water layer maximum. Group 4: oxic-zone high concentra- tions. Group 5: suboxic zone maximum. Group 6: absent in oxic zone, bottom layer maximum. Compound names are given in Ta- ble 2. www.biogeosciences.net/11/7009/2014/ Biogeosciences, 11, 7009–7023, 2014 7014 C. Berndmeyer et al.: Biomarkers in the stratified water column Figure 4. Vertical distribution of biomarkers in the Landsort Deep water column. The suboxic zone is shaded grey. only occurred from 65 to 80 m, with a maximum at 65 m (287 and 228 ng L−1, respectively). Of Group 3, the 16 : 0−18 : 1 wax ester was included in the discussion. δ13C values of the wax esters were ∼−28 ‰ (Table 1). 3.2.4 Group 4: oxic-zone maximum Group 4 consisted exclusively of saturated n-alkanes from n-C21 to n-C36 as well as 26 : 0 PLFA (43). 26 : 0 PLFA only occurred at 80 m, whereas all other compounds were abundant from the surface to the upper suboxic zone at 80 m (data not shown). The homologues n-C27 (74), n-C29 (76), and n-C31 (78) show maxima at the surface (21–30 ng L−1). For the other compounds, maxima were either located at 65 or 70 m, with highest concentrations for n-C25− n-C36 (10–23 ng L−1). Below 80 m, concentrations dropped to con- stantly low values. As an example, the depth profile of n-C25 (71) is shown in Fig. 4. δ13C values for these compounds could not be obtained. Figure 5. Concentrations of n-alkanes and n-alkenes in the Land- sort Deep surface layer (oxic zone, 10 m water depth). 3.2.5 Group 5: suboxic-zone maximum Group 5 contained only two compounds: 16 : 1ω8c PLFA (10) and the n-C26 : 1 alkene (72). n-C26 : 1 occurred in very low concentrations at 10 m and peaked at 80 and 95 m (7– 8 ng L−1). 16 : 1ω8c PLFA occurred only at 80 and 90 m wa- ter depth, with highest values at 80 m (8 ng L−1; Fig. 4), and was chosen for further discussion. δ13C values of this com- pound were ∼−45 ‰ (Table 1). 3.2.6 Group 6: absent in oxic zone, bottom-layer maximum Group 6 consisted of compounds that only occurred in the suboxic zone and below and increased in concentration in the anoxic zone. An exception is 5α(H)-cholestan-3β- ol (cholestanol; 45), which was also present in the surface layer. 10-me-16 : 0 PLFA (16), the irregular C25 isoprenoid 2, 6, 10, 15, 19-pentamethylicosane (PMI) and three unsat- urated derivatives thereof (PMI 1; 53), 2, 3-di-0-isopranyl sn-glycerol diether (archaeol; 51), and cholestanol were con- sidered for further discussion. For all compounds, maxima were detected in the anoxic zone, with highest concentra- tions observed for cholestanol (35 ng L−1) followed by 10- me-16 : 0 PLFA (10 ng L−1), PMI and PMI 1 (8 ng L−1), and archaeol (1 ng L−1). Compared to other compounds, 10- me-16 : 0 PLFA shows a slight 13C depletion in the anoxic zone (−35.4 ‰ ; Table 1). Concentrations of archaeol, PMI, and PMI 1 were too low to determine δ13C. 3.2.7 n-Alkanes and n-alkenes in the sea surface layer The concentrations of n-alkanes and n-alkenes in the surface sample (10 m water depth) are given in Fig. 5. The longest n-alkane chain was n-C36, and odd carbon numbers domi- nated over even. Highest concentrations were found for n- C27 (21 ng L−1), n-C29 (30 ng L−1), and n-C31 (26 ng L−1). The longest n-alkene chain was n-C26 : 1, and highest n- alkene concentrations were measured for n-C23 : 1 (3 ng L−1) and n-C25 : 1 (3 ng L−1). Biogeosciences, 11, 7009–7023, 2014 www.biogeosciences.net/11/7009/2014/ C. Berndmeyer et al.: Biomarkers in the stratified water column 7015 3.2.8 Water-column profiles of BHPs In the Landsort Deep, seven individual BHPs were identified (Fig. 6a). In all samples, bacteriohopane-32, 33, 34, 35-tetrol (BHT) accounted for the greatest portion of the total BHPs (88–94 %). An as yet uncharacterized BHT isomer, BHT II, was present only below 70 m and showed its highest rela- tive abundance (∼ 2 %) between 70 and 90 m. BHT cycli- tol ether, BHT glucosamine, and 35-aminobacteriohopane- 32, 33, 34-triol (aminotriol) were present throughout the wa- ter column. BHT cyclitol ether and BHT glucosamine were most abundant in the oxic zone (ca. 1–4 %) but showed only minor abundances (< 1 %) below. Aminotriol was el- evated at 65 and 420 m (∼ 7 and ∼ 5 %, respectively). 35- Aminobacteriohopane-31, 32, 33, 34-tetrol (aminotetrol) oc- curred throughout the suboxic and anoxic zones, whereas 35- aminobacteriohopane-30, 31, 32, 33, 34-pentol (aminopen- tol) was observed only at 90 m and below. Both aminotetrol and aminopentol showed minor relative abundances of ∼ 2 and < 1 % of the total BHPs, respectively (Jakobs et al., 2014). For comparison, the major phytoplankton species from a cyanobacterial bloom in the Gotland Deep (2012) were deter- mined by microscopy (HELCOM manual, 2012) and the par- ticulate organic matter (POM) was analyzed for BHPs. This reference biomass contained mainly Aphanizonemon and, to a smaller extent, Anabaena and Nodularia, which were ac- companied by dinoflagellates. Three BHPs were observed in the bloom POM (Fig. 6b). Among these compounds, the most abundant was BHT (∼ 86 %), followed by BHT cyclitol ether (∼ 10 %) and BHT glucosamine (∼ 4 %). 4 Discussion In the following, we discuss several aspects of the biomarker profiles with respect to their significance as tracers for the relevant biota and biogeochemical processes in stratified wa- ter columns. 4.1 Water-column redox zones as reflected by cholestanol / cholesterol ratios Different redox states of the Landsort Deep water column and the associated microbial processes are reflected by the profiles of cholesterol and its diagenetic product, cholestanol (Fig. 4; Groups 1 and 6, respectively). Cholesterol is syn- thesized by various eukaryotic phyto- and zooplankton and higher plants (Parrish et al., 2000) and is abundant in wa- ter columns and sediments. In sediments as well as in stratified water columns, stanols are produced from sterols by anaerobic bacterial hydrogenation (Gaskell and Eglin- ton, 1975; Wakeham, 1989) and by the abiotic reduction of double bonds by reduced inorganic species such as H2S (Hebting et al., 2006; Wakeham et al., 2007). Therefore, cholestanol / cholesterol ratios typically increase under more Figure 6. Relative abundances of individual BHPs (as percent of the total) of (a) the Landsort Deep water column and (b) the Gotland Deep cyanobacterial bloom. Note that [%] axes start at 85 %. An asterisk indicates data taken from Jakobs et al. (2014). reducing conditions. In the Black Sea, low ratios of ∼ 0.1 were associated with oxygenated surface waters, and the sub- oxic zone showed ratios between 0.1 and 1, whereas the anoxic zone revealed values > 1 (Wakeham et al., 2007). In the Landsort Deep, the cholestanol / cholesterol ratios showed a slight increase with depth from the surface towards the suboxic zone but always remained < 0.1 (Fig. 4). Below, the values increased to ∼ 0.3 in the suboxic zone and further to a maximum of 0.45 in the anoxic zone. Whereas the ra- tios in the Landsort Deep are considerably lower than in the Black Sea, the depth trend still clearly mirrors the changes from oxic to suboxic, and further to anoxic conditions. It is also interesting to note that total cholesterol and cholestanol concentrations in the Landsort Deep were ten- and fourfold higher, respectively, than in the Black Sea (Wakeham et al., 2007). 4.2 Phototrophic primary production As expected, in situ biomarkers for phototrophic organisms were most abundant in the surface layer and are pooled in PCA Group 1. 20 : 4ω6 PLFA is a biomarker traditionally assigned to eukaryotic phytoplankton (Nanton and Castell, 1999; Lang et al., 2011) and organisms grazing thereon, such as protozoa (Findlay and Dobbs, 1993; Pinkart et al., 2002; Risse-Buhl et al., 2011). 20 : 5ω3 PLFA is known to be a ma- jor compound in diatoms (Arao and Marada, 1994; Dunstan et al., 1994), and high concentrations of these PLFAs, as ob- served in the surface layer of the oxic zone, are in good agree- ment with such an autochthonous plankton-based source. 7-Methylheptadecane is a characteristic marker for cyanobacteria (Shiea et al., 1990; Köster et al., 1999). Its www.biogeosciences.net/11/7009/2014/ Biogeosciences, 11, 7009–7023, 2014 7016 C. Berndmeyer et al.: Biomarkers in the stratified water column most likely source are members of the subclass Nosto- cophyceae that were often reported to produce isomeric midchain branched alkanes, including 7-methylheptadecane (Shiea et al., 1990; Hajdu et al., 2007; Liu et al., 2013). Nos- tocophyceae are key members of the photoautotrophic com- munity in the Baltic Sea. Particularly the filamentous genera Nodularia and Aphanizonemon (see Sect. 3.2.8) and the pic- ocyanobacterium Synechococcus play a major role in blooms during summer time (Stal et al., 2003; Labrenz et al., 2007). The importance of cyanobacteria in the surface layer of the Landsort Deep is further reflected by the presence of C21 : 1, C23 : 1, and C25 : 1 n-alkenes (Fig. 5). These compounds have been reported from Anacystis (Gelpi et al., 1970) and Oscil- latoria (Matsumoto et al., 1990). Oscillatoria vaucher is also known to occur in the Baltic Sea but is of only minor abun- dance (Kononen et al., 1996; Vahtera et al., 2007). Unlike the n-alkenes that only occurred in the surface layer, long-chain n-alkanes were present in the whole water column, with high abundances in the oxic zone. Long-chain n-alkanes with a strong predominance of the odd-numbered n-C25 to n-C36 homologues (Eglinton and Hamilton, 1967; Bi et al., 2005) and β-sitosterol (Volkman, 1986) are typical components of higher-plant lipids, thus indicating continen- tal runoff and/or aeolian input of terrigenous organic mat- ter into the Landsort Deep. n-C27, n-C29, and n-C31 showed surface maxima (not shown), indicating similar sources as for β-sitosterol and a contribution of land plant leaf waxes. Other than β-sitosterol, most n-alkanes peaked between 65 and 70 m (n-C25 for example; Fig. 4). Apart from the surface peaks, this is also true for n-C27, n-C29, and n-C31. A pos- sible explanation is the accumulation of terrigenous higher- plant particles accumulating at the pycnocline, where density differences were highest (MacIntyre et al., 1995) 4.3 Phototrophic vs. heterotrophic dinoflagellates, and ciliates The distribution of dinoflagellates and, most likely, ciliates in the water column is reflected by two specific biomarkers: di- nosterol and tetrahymanol (see Sect. 3.2.2, Fig. 4). Dinosterol is mainly produced by dinoflagellates (Boon et al., 1979), al- though it was also reported in minor abundance from a di- atom (Navicula sp., Volkman et al., 1993). The dinosterol concentrations in the Landsort Deep showed a bimodal dis- tribution. The strong peak in the surface layer of the oxic zone probably represents contributions from phototrophic di- noflagellates. Plausible candidates are Peridiniella catenata and Scrippsiella hangoei, both of which are involved in the spring phytoplankton blooms in the central Baltic Sea (Was- mund et al., 1998; Höglander et al., 2004). The latter species was previously reported to produce dinosterol (Leblond et al., 2007). However, P. catenata as well as S. hangoei are vir- tually absent below 50 m water depth (Höglander et al., 2004) and can thus not account for the second peak of dinosterol at the suboxic–anoxic transition zone. An accumulation of surface-derived dinosterol at the bottom of the suboxic zone is unlikely, as the pycnocline, and thus the strongest density discontinuity, is located at 60–70 m water depth, i.e., about 20 m further above. Dinosterol is absent in the pycnocline and only occurs from the bottom of the suboxic zone down- wards and below. Instead, a likely source of dinosterol at this water depth is represented by heterotrophic dinoflagellates that are abundant in the suboxic zones of the central Baltic Sea (Anderson et al., 2012). Due to their enhanced produc- tivity, these environments provide good conditions to sustain communities of eukaryotic grazers (Detmer et al., 1993). A possible candidate, Gymnodinium beii, was described from the suboxic zones of the central Baltic Sea (Stock et al., 2009). Indeed, several Gymnodinium species are known to be heterotrophs (Strom and Morello, 1998) and some have been reported to produce dinosterol (Mansour et al., 1999). Like cholesterol and β-sitosterol, dinosterol was also found in the anoxic zone at 400 m water depth. The production of these compounds at this depth is unlikely, as the synthesis of sterols requires oxygen (Summons et al., 2006). Hence, the observed sterol occurrences probably reflect transport through the water column. A similar concentration distribution as for dinosterol was observed for tetrahymanol. Tetrahymanol is known to be pro- duced by ferns, fungi, and bacteria, such as the purple non- sulfur bacterium Rhodopseudomonas palustris (Zander et al., 1969; Kemp et al., 1984; Kleemann et al., 1990; Sinninghe Damsté et al., 1995; Eickhoff et al., 2013). Moreover, cil- iates ubiquitously produce tetrahymanol as a substitute for cholesterol when grazing on prokaryotes instead of eukary- otes, such as algae (Conner et al., 1968; Boschker and Mid- delburg, 2002). This is also a feasible scenario for the Baltic Sea, where the ciliate genera Metopus, Strombidium, Meta- cystis, Mesodinium, and Coleps are abundant in the sub- oxic zone and at the suboxic–anoxic interface (Detmer et al., 1993; Anderson et al., 2012). Unidentified ciliates also oc- curred in the anoxic waters of the Landsort Deep (Anderson et al., 2012). Members of the genus Rhodopseudomonas, a possible alternative source of tetrahymanol, have so far not been identified in the suboxic zone (Labrenz et al., 2007; Thureborn et al., 2013). We therefore regard bacterivorous ciliates living under suboxic to anoxic conditions as the most likely source of tetrahymanol in the suboxic zone and below. Likewise, ciliates feeding on chemoautotrophic bacteria were assumed as producers of tetrahymanol in the suboxic zone of the Black Sea (Wakeham et al., 2007). The situation is some- what different in the surface waters, where tetrahymanol shows its maximum concentrations at 10 m water depth. Al- though Rhodopseudomonas and other purple nonsulfur bac- teria usually occur under oxygen-deficient conditions, they have been genetically identified in the surface water of the Landsort Deep (Farnelid et al., 2009) and thus have to be considered as potential producers of tetrahymanol. Further- more, cholesterol is abundant in the surface waters and could be incorporated by ciliates instead of tetrahymanol. However, Biogeosciences, 11, 7009–7023, 2014 www.biogeosciences.net/11/7009/2014/ C. Berndmeyer et al.: Biomarkers in the stratified water column 7017 some ciliates seem to prefer prokaryotes as a prey. Sinking agglomerates of cyanobacteria and other bacteria are known to be covered by feeding ciliates (Gast and Gocke, 1988). Hence, in addition to R. palustris, ciliates grazing selectively on cyanobacteria would plausibly explain the abundance of tetrahymanol in the shallow waters of the Landsort Deep. δ13C values of tetrahymanol revealed an opposite trend to dinosterol. While dinosterol became isotopically more neg- ative with depth (−29.9 to −32.0 ‰), tetrahymanol became more positive (−28.7 to −25.9 ‰) and showed its highest δ13C values in the anoxic zone. Although ciliates and di- noflagellates are both grazers at the suboxic–anoxic inter- face, they seem to occupy different ecological niches and feed on different bacterial sources. 4.4 Heterotrophs in the cold winter water layer The only biomarkers with enhanced concentrations in the deep cold winter water layer are wax esters (e.g., 16 : 0−18 : 1 wax ester, Fig. 4) and, to a minor extent, cholesterol and 20 : 5ω3 PLFA. As the pycnocline, and thus a strong density discontinuity, is also located at this depth, an accumulation of settling organic debris containing these compounds has to be considered (MacIntyre et al., 1995). Living organisms, how- ever, may be also be plausible sources. Copepods are known producers of wax esters and cholesterol (Lee et al., 1971; Sargent et al., 1977; Kattner and Krause, 1989; Nanton and Castell, 1999; Falk-Petersen et al., 2002), which are often abundant in density layers, where they feed on accumulated aggregates (MacIntyre et al., 1995). These organisms syn- thesize wax esters with total chain lengths of between 28 and 44 carbon atoms (Lee et al., 1971; Kattner and Krause, 1989; Falk-Petersen et al., 2002); several of wax esters were present in the Landsort Deep (data not shown in Fig. 4), with roughly the same distribution as the most prominent 16 : 0− 18 : 1. Although copepods migrate through the water column, par- ticularly those rich in wax esters prefer deep water or near- surface cold water (Sargent et al., 1977), which is in full agreement with the high amounts of these compounds in the cold winter water layer. Copepods are abundant and diverse in the Baltic Sea, with major species being Pseudocalanus elongatus, Temora longicornis, and Acartia spp. (Möllmann et al., 2000; Möllmann and Köster, 2002). Like the wax es- ters, the 20 : 5ω3 PLFA shows higher concentrations in the cold winter water layer, but it is also abundant in the surface and at the suboxic–anoxic interface (Fig. 4). Copepods are also known to feed on diatoms and incorporate their specific fatty acids, such as 20 : 5ω3 PLFA, largely unchanged into their own tissue (Kattner and Krause, 1989). Dinoflagellates are also known producers of 20 : 5ω3 PLFA (Parrish et al., 1994; Volkman et al., 1998) and may be an alternative source in the surface layer and at the suboxic–anoxic interface; this is supported by a good correlation with dinosterol at these depths. Unlike the abovementioned compounds, all other selected biomarkers show particularly low concentrations in the cold winter water layer. This is also true for widespread com- pounds such as the 16 : 1ω7c PLFA, which is produced by eukaryotes (Pugh, 1971; Shamsudin, 1992) as well as prokaryotes (Parkes and Taylor, 1983; Vestal and White, 1989). While a mixed origin of 16 : 1ω7c PLFA has to be assumed for the oxic zone, a bacterial source is more prob- able in the suboxic zone and in the anoxic zone. Regardless of the biological source, a very low amount of this ubiqui- tous fatty acid (Fig. 4) indicates that the cold winter water layer of the Landsort Deep does not support abundant plank- tonic life. Based on microscopy, similar observations have been made for the cold winter water layers of the Gotland, Bornholm, and Gdansk basins (Gast and Gocke (1988) and citations therein). 4.5 BHPs as indicators for aerobic and anaerobic metabolisms Bacteria are the only known source of BHPs (Kannenberg and Poralla, 1999). Although the biosynthesis of BHPs and their precursor, diploptene (both part of Group 2), does not require oxygen, the production of hopanoids was long as- sumed to be restricted to aerobic bacteria, as reports from facultatively or strictly anaerobic bacteria were initially lack- ing. More recently, however, planctomycetes (Sinninghe Damsté et al., 2004), metal-reducing Geobacter (Fischer et al., 2005), and sulfate-reducing Desulfovibrio (Blumenberg et al., 2006, 2009, 2012) were identified as anaerobic produc- ers of BHPs. In the Landsort Deep, cyanobacteria are abun- dant in the surface water layer and may be considered as a major source of BHPs (cf. Talbot et al., 2008; Welander et al., 2010). Evidence for such cyanobacterial BHP contribu- tions may come from our analysis of a Gotland Deep bloom from summer 2012 (see Sect. 3.2.7). BHPs identified in this bloom were BHT, BHT cyclitol ether, and BHT glucosamine (Fig. 6b), which is in line with the BHP composition of the Landsort Deep surface layer (Fig. 6a). These three cyanobac- terial BHPs were present throughout the Landsort Deep wa- ter column, although they were present in minor amounts in the suboxic zone and below. In addition, the surface layer contained aminotriol that was also present in the whole water column. Aminotriol is an abundant BHP produced by various bacteria (e.g., Talbot and Farrimond (2007) and references therein), indicating that organisms other than cyanobacteria may contribute BHP to the surface layer. A further notable feature is the occurrence of BHT II at 70 m and below. The source of BHT II is not fully re- solved yet. It was recently related to planctomycetes, es- pecially those performing anaerobic ammonium oxidation (anammox) in sediments (Rush et al., 2014). Anammox bac- teria can also be traced by 10-me16 : 0 PLFA and ladder- ane PLFAs (not studied here; Sinninghe Damsté et al., 2005; Schubert et al., 2006). 10-me16 : 0 PLFA does indeed show www.biogeosciences.net/11/7009/2014/ Biogeosciences, 11, 7009–7023, 2014 7018 C. Berndmeyer et al.: Biomarkers in the stratified water column a peak in the lower suboxic zone, where BHT II is abun- dant. However, 10-me16 : 0 PLFA may also be contributed by sulfate-reducing bacteria (see Sect. 4.6), and no evidence for anammox has been observed in the water column of the Landsort Deep from molecular biological studies so far (Hi- etanen et al., 2012; Thureborn et al., 2013). Regardless of the biological source, BHT II was described in stratified water columns of the Arabian Sea, Peru Margin and Cariaco Basin (Sáenz et al., 2011), and the Gotland Deep (Berndmeyer et al., 2013) and has therefore been proposed as a proxy for stratified water columns. This hypothesis has been adopted to reconstruct the development of water-column stratifica- tion in the Baltic Sea during the Holocene (Blumenberg et al., 2013). Like BHT II, aminotetrol and aminopentol are absent from the surface layer (Fig. 6a). Whereas both BHPs are biomark- ers for methanotrophic bacteria, the latter typically occurs in type I methanotrophs (Talbot et al., 2001). The presence of type I methanotrophic bacteria is further supported by the co-occurrence of the specific 16 : 1ω8c PLFA (Nichols et al., 1985; Bowman et al., 1991, 1993) and its considerably de- pleted δ13C value (−45.4 ‰). Whereas a major in situ production of BHPs in the sub- oxic zone is evident from our data, the sources of BHPs in the anoxic zone are more difficult to establish. BHPs in the anoxic zone may partly derive from sinking POM as well as being newly produced by anaerobic bacteria. Sink- ing POM as a source may apply to BHT cyclitol ether and BHT glucosamine, which seem to derive from cyanobacte- ria thriving in the oxic zone, as discussed above. Aminotriol, aminotetrol, and aminopentol, however, are known products of sulfate-reducing bacteria (Blumenberg et al., 2006, 2009, 2012) and may have their origin within the anoxic zone. This interpretation is supported by the close correlation of the total BHPs with the ai-15 : 0 PLFA, which is considered as indica- tive of sulfate reducers (see Sect. 4.6; both compounds were part of the same PCA Group 2). Thus, the anoxic zone of the Landsort Deep is likely an active source for BHPs rather than solely being a pool for transiting compounds. 4.6 Microbial processes in the anoxic zone Sulfate-reducing bacteria were traced using ai-15 : 0 PLFA and 10-me-16 : 0 PLFA (Parkes and Taylor, 1983; Taylor and Parkes, 1983; Vainshtein et al., 1992). The high abundance of ai-15 : 0 PLFA in the surface layer (Fig. 4) is surprising at first glance, as sulfate reducers are not supposed to thrive in oxic environments. However, these bacteria were previously reported from oxygenated surface waters of the Gotland Deep, where they were associated with sinking cyanobac- terial agglomerates (Gast and Gocke, 1988). 10-Me-16 : 0 PLFA is otherwise absent from the oxic zone (Fig. 4). This FA was reported to occur in Desulfobacter and Desulfobac- ula (Taylor and Parkes, 1983; Kuever et al., 2001), both strictly anaerobic organisms (Szewzyk and Pfennig, 1987; Widdel, 1987; Kuever et al., 2001). Indeed, Desulfobacula toluolica was genetically identified by Labrenz et al. (2007) in suboxic and anoxic waters of the central Baltic Sea. In addition to the bacterial FA, two archaeal in situ biomarkers, archaeol and PMI, were identified. Archaeol is the most common ether lipid in archaea but is especially abundant in euryarchaeotes, including methanogens (Torn- abene and Langworthy, 1979; Koga et al., 1993). Like- wise, PMI and its unsaturated derivatives are diagnostic for methanogenic euryarchaeotes (Tornabene et al., 1979; De Rosa and Gambacorta, 1988; Schouten et al., 1997). In the Landsort Deep, both compounds are virtually absent in the oxic zone and increase in abundance with depth through the suboxic zone (Fig. 3). The same trend has been described for PMI in the Black Sea (Wakeham et al., 2007), and the pres- ence of Euryarchaeota in Landsort Deep anoxic waters has recently been proven by Thureborn et al. (2013). Given the available sample resolution, it is impossible to further elucidate the exact distribution of archaea in the anoxic zone of the Landsort Deep. Likewise, δ13C values could not be obtained for archaeol and PMI due to low com- pound concentrations, which makes statements on inputs of these lipids from archaea involved in the sulfate-dependent anaerobic oxidation of methane (AOM) impossible (cf. Hin- richs et al., 1999; Thiel et al., 2001). Whereas it has been shown that AOM is theoretically possible in the anoxic zone of the Landsort Deep and anaerobic methane consumption has recently been demonstrated to occur (Jakobs et al., 2013), clear evidence for abundant AOM is as yet lacking and re- quires further investigations focused on the anoxic water bodies of the Baltic Sea. 5 Conclusions The Landsort Deep in the western central Baltic Sea is char- acterized by a stratified water column. Marine microbial or- ganisms have adapted to the vertical chemical limitations of their ecosystems, and their distributions in the water column can be reconstructed using diverse in situ biomarkers. Ac- cording to their behavior in the water column, PCA analy- sis revealed six groups of biomarkers for distinct groups of (micro)organisms and the related biogeochemical processes. Within the oxic zone, a clear preference for the surface layer became obvious for distinctive biomarkers. Among these compounds, 7-methylheptadecane, different alkenes, BHT cyclitol ether, and BHT glucosamine were indicative of the presence of bacterial primary producers, namely cyanobac- teria. Dinosterol concentrations and −δ13C values revealed a phototrophic dinoflagellate population in the surface wa- ters and a second, heterotrophic community thriving at the suboxic–anoxic interface. Similarly, abundant tetrahymanol at the surface indicated ciliates feeding on cyanobacterial ag- glomerates, but a second maximum at the suboxic–anoxic interface suggested a further ciliate population that grazed Biogeosciences, 11, 7009–7023, 2014 www.biogeosciences.net/11/7009/2014/ C. Berndmeyer et al.: Biomarkers in the stratified water column 7019 on chemoautotrophic bacteria. The cold winter water layer at the bottom of the oxic zone showed only low concentra- tions of biomarkers and seemed to be avoided by most organ- isms, except copepods. In contrast, biomarkers obtained from the suboxic zone reflected a high abundance and diversity of eukaryotes and prokaryotes. Whereas 16 : 1ω8 PLFA and aminopentol revealed the presence of type I aerobic methane- oxidizing bacteria, ai-15 : 0 PLFA, 10-me-16 : 0, and total BHPs indicated the distribution of sulfate-reducing bacte- ria in the Landsort Deep water column. The close coupling of ai-15 : 0 PLFA with total BHPs suggests that these bac- teria represent a major in situ source for hopanoids in the anoxic zone. The anoxic zone was also inhabited by most likely Euryarchaeota, as shown by the presence of archaeol and PMI and its derivatives. Our study in the water column of the Landsort Deep gives insights into the recent distribu- tions and actual sources of organic matter as reflected by lipid biomarkers. The results may also aid the interpretation of or- ganic matter preserved in the sedimentary record and thus help to better constrain changes in the geological history of the Baltic Sea. Acknowledgements. We thank the captains and crews of R/Vs Elisabeth Mann Borghese and Meteor for assistance during the cruises. We thank C. Conradt and L. Kammel for laboratory assistance, T. Licha and K. Nödler for help with LC–MS, and N. Cerveau for help with R. We thank S. Bühring and an anony- mous reviewer for helpful comments on our manuscript. The German Research Foundation (Deutsche Forschungsgemeinschaft, DFG) is acknowledged for financial support (Grants BL 971/1-3 and 971/3-1). This Open Access Publication is funded by the University of Göttingen. Edited by: J. Middelburg References Anderson, R., Winter, C., and Jürgens, K.: Protist grazing and viral lysis as prokaryotic mortality factors at Baltic Sea oxic-anoxic interfaces, Mar. Ecol. Progr., 467, 1–14, 2012. Arao, T. and Yamada, M.: Biosythesis of polyunsaturated fatty acids in the marine diatom, Phaeodactylum tricornutum, Phytochem- istry, 35, 1177–1181, 1994. Beliaeff, B. and Burgeot, T.: Integrated biomarker response: a useful tool for ecological risk assessment, Environ. Toxicol. Chem., 21, 1316–1322, 2001. Bergström, S. and Matthäus, W.: Meteorology, hydrology and hy- drography, in: Third periodic assessment of the state of the marine environment of the Baltic Sea, 1989–1993; background document, Baltic Sea environment proceedings 64b, HELCOM, Helsinki, 9–18, 1996. Berndmeyer, C., Thiel, V., Schmale, O., and Blumenberg, M.: Biomarkers for aerobic methanotrophy in the water column of the stratified Gotland Deep (Baltic Sea), Org. Geochem., 55, 103–111, 2013. Bi, X., Sheng, G., Liu, X., Li, C., and Fu, J.: Molecular and car- bon and hydrogen isotopic composition of n-alkanes in plant leaf waxes, Org. Geochem., 36, 1405–1417, 2005. Blumenberg, M., Krüger, M., Nauhaus, K., Talbot, H. M., Opper- mann, B. I., Seifert, R., Pape, T., and Michaelis, W.: Biosynthesis of hopanoids by sulfate-reducing bacteria (genus Desulfovibrio), Environ. Microbiol., 8, 1220–1227, 2006. Blumenberg, M., Seifert, R., and Michaelis, W.: Aerobic methan- otrophy in the oxic-anoxic transition zone of the Black Sea water column, Org. Geochem., 38, 84–91, 2007. Blumenberg, M., Oppermann, B. I., Guyoneaud, R., and Michaelis, W.: Hopanoid production by Desulfovibrio bastinii isolated from oilfield formation water, FEMS Microbiol. Lett., 293, 73–78, 2009. Blumenberg, M., Hoppert, M., Krüger, M., Dreier, A., and Thiel, V.: Novel findings on hopanoid occurrences among sulfate re- ducing bacteria: Is there a direct link to nitrogen fixation?, Org. Geochem., 49, 1–5, 2012. Blumenberg, M., Berndmeyer, C., Moros, M., Muschalla, M., Schmale, O., and Thiel, V.: Bacteriohopanepolyols record strat- ification, nitrogen fixation and other biogeochemical perturba- tions in Holocene sediments of the central Baltic Sea, Biogeo- sciences, 10, 2725–2735, doi:10.5194/bg-10-2725-2013, 2013. Boon, J. J., Rijpstra, W. I. C., De Lange, F., and De Leeuw, J. W.: Black Sea sterol – a molecular fossil for dinoflagellate blooms, Nature, 277, 125–127, 1979. Boschker, H. T. S. and Middelburg, J. J.: Stable isotopes and biomarkers in microbial ecology, FEMS Microbiol. Ecol., 40, 85–95, 2002. Bowman, J. P., Skeratt, J. H., Nichols, P. D., and Sly, L. I.: Phop- sholipid fatty acid and lipopolysaccharide fatty acid signature lipids in methane-utilizing bacteria, FEMS Microbiol. Ecol., 85, 15–22, 1991. Bowman, J. P., Sly, L. I., Nichols, P. D., and Hayward, A. C.: Re- vised taxonomy of the methanotrophs: Description of Methy- lobacter gen. nov., emendation of Methylococcus, validation of Methylosinus and Methylocystis species, and a proposal that the family Methylococcaceae includes only the group I methan- otrophs, Int. J. Syst. Bacteriol., 43, 735–753, 1993. Carlson, D. R., Roan, C.-S., Yost, R. A., and Hector, J.: Dimethyl disulfide derivatives of long chain alkenes, alkadiens, and alkatrienes for gas chromatography/mass spectrometry, Anal. Chem., 61, 1564–1571, 1989. Conner, R. L., Landrey, J. R., Burns, C. H., and Mallory, F. B.: Cholesterol inhibition of pentacyclic triterpenoid biosythesis in Tetrahymena pyriformis, J. Protozool., 15, 600–605, 1968. De Rosa, M. and Gambacorta, A.: The lipids of archaebacteria, Prog. Lip. Res., 27, 153–175, 1988. Dellwig, O., Leipe, T., März, C., Glockzin, M., Pollehne, F., Schnet- ger, B., Yakushev, E. V., Böttcher, M. E., and Brumsack, H.-J.: A new particulate Mn–Fe–P-shuttle at the redoxcline of anoxic basins, Geochim. Cosmochim. Ac., 74, 7100–7115, 2010. Detmer, A. E., Giesenhagen, H. C., Trenkel, V. M., Auf dem Venne, H., and Jochem, F.: Phototrophic and hetreotrophic pico- and nanoplankton in anoxic depths of the central Baltic Sea, Mar. Ecol. Progr., 99, 197–203, 1993. www.biogeosciences.net/11/7009/2014/ Biogeosciences, 11, 7009–7023, 2014 7020 C. Berndmeyer et al.: Biomarkers in the stratified water column Dunstan, G. A., Volkman, J. K., Barrett, S. M., Leroi, J.-M., and Jeffrey, S. W.: Essential polyunsaturated fatty acids from 14 species of diatom (Bacillariophyceae), Phytochemistry, 35, 155– 161, 1994. Eglinton, G. and Hamilton, R. J.: Leaf epicuticular waxes, Science, 156, 1322–1335, 1967. Eickhoff, M., Birgel, D., Talbot, H. M., Peckmann, J., and Kappler, A.: Oxidation of Fe(II) leads to increased C2 methylation of pen- tacyclic triterpenoids in the anoxygenic phototrophic bacterium Rhodopseudomonas palustris strain TIE-1, Geobiology, 11, 268– 278, 2013. Falk-Petersen, S., Dahl, T. M., Scott, C. L., Sargent, J. R., Gulliksen, B., Kwasniewski, S., Hop, H., and Millar, R.-M.: Lipid biomark- ers and trophic linkages between ctenophores and copepods in Svalbard waters, Mar. Ecol. Progr., 227, 187–194, 2002. Farnelid, H., Öberg, T., and Riemann, L.: Identity and dynamics of putative N2-fixing picoplankton in the Baltic proper suggest complex patterns of regulation, Environmental Microbiology Re- ports, 1, 145–154, 2009. Findlay, R. H. and Dobbs, F. C.: Quantitative description of micro- bial communities using lipid analysis, in: Handbook of methods in aquatic microbial ecology, edited by: Cole, J. J., CRC Press, Florida, USA, 271–284, 1993. Fischer, W. W., Summons, R. E., and Pearson, A.: Targeted ge- nomic detection of biosynthetic pathways: anaerobic produc- tion of hopanoid biomarkers by a common sedimentary microbe, Geobiology, 3, 33–40, 2005. Gaskell, S. J. and Eglinton, G.: Rapid hydrogenation of sterols in a contemporary lacustrine sediment, Nature, 254, 209–211, 1975. Gast, V. and Gocke, K.: Vertical distribution of number, biomass and size-class spectrum of bacteria in relation to oxic/anoxic con- ditions in the Central Baltic Sea, Mar. Ecol. Progr., 45, 179–186, 1988. Gatellier, J.-P. L. A., de Leeuw, J. W., Sinninghe Damsté, J. S., Derenne, S., Largeau, C., and Metzger, P.: A comparative study of macromolecular substances of a Coorongite and cell walls of the extant alga Botryococcus braunii, Geochim. Cosmochim. Ac., 57, 2053–2068, 1993. Gelpi, E., Schneider, H., Mann, J., and Oró, J.: Hydrocarbons of geochemical significance in microscopic algae, Phytochemistry, 9, 603–612, 1970. Grasshoff, K., Kremling, K., and Ehrhardt, M.: Methods of seawater analysis, Verlag Chemie, Weinheim, Germany, 75–89, 1983. Hajdu, S., Höglander, H., and Larsson, U.: Phytoplankton verti- cal distributions and compositions in Baltic Sea cyanobacterial blooms, Harmful Algae, 6, 189–205, 2007. Hanson, N., Förlin, L., and Larsson, A.: Evaluation of long-term biomarker data from perch (Perca fluviatilis) in the Baltic Sea suggests increasing exposure to environmental pollutants, Envi- ron. Toxicol. Chem., 28, 364–373, 2009. Hebting, Y., Schaeffer, P., Behrens, A., Adam, P., Schmitt, G., Sch- neckenburger, P., Bernasconi, S. M., and Albrecht, P.: Biomarker evidence for a major preservation pathway of sedimentary or- ganic carbon, Science, 312, 1627–1631, 2006. HELCOM COMBINE: Manual for marine monitoring in the COM- BINE programme of HELCOM, Part C, Programme for moni- toring of eutrophication and its effects, Annex C-6, Guidelines concerning phytoplankton species composition, abundance and biomass, HELCOM, 310–325, 2012. Hietanen, S., Jäntti, H., Buizert, C., Jürgens, K., Labrenz, M., Voss, M., and Kuparinen, J.: Hypoxia and nitrogen processing in the Baltic Sea water column, Limnol. Oceanogr., 57, 325–337, 2012. Hinrichs, K. U., Hayes, J. M., Sylva, S. P., Brewer, P. G., and De- Long, E. F.: Methane-consuming archaebacteria in marine sedi- ments, Nature, 398, 802–805, 1999. Höglander, H., Larsson, U., and Hajdu, S.: Vertical distribution and settling of spring phytoplankton in the offshore NW Baltic Sea proper, Mar. Ecol. Progr., 283, 15–27, 2004. Jakobs, G., Rehder, G., Jost, G., Kießlich, K., Labrenz, M., and Schmale, O.: Comparative studies of pelagic microbial methane oxidation within the redox zones of the Gotland Deep and Land- sort Deep (central Baltic Sea), Biogeosciences, 10, 7863–7875, doi:10.5194/bg-10-7863-2013, 2013. Jakobs, G., Holtermann, P., Berndmeyer, C., Rehder, G., Blumen- berg, M., Jost, G., Nausch, G., and schmale, O.: Seasonal and spatial methane dynamic in the water column of the central Baltic Sea (Gotland Sea), Cont. Shelf Res., 91, 12–25, 2014. Kamyshny Jr., A., Yakushev, E. V., Jost, G., and Podymov, O. I.: Role of sulfide oxidation intermediates in the redox balance of the oxic-anoxic interface of the Gotland Deep, Baltic Sea, in: The handbook of environmental chemistry, edited by: Yakushev, E. V., Springer, Berlin Heidelberg, 22, 95–119, 2013. Kannenberg, E. L. and Poralla, K.: Hopanoid biosythesis and func- tion in bacteria, Naturwissenschaften, 86, 168–176, 1999. Kattner, G. and Krause, M.: Seasonal variations of lipids (wax es- ters, fatty acids and alcohols) in calanoid copepods from the North Sea, Mar. Chem., 26, 261–275, 1989. Kemp, P., Lander, D. J., and Orpin, C. G.: The lipids of the ru- men fungus Piromonas communis, J. Gen. Microbiol., 130, 27– 37, 1984. Kleemann, G., Poralla, K., Englert, G., Kjøsen, H., Liaaen-Jensen, S., Neunlist, S., and Rohmer, M.: Tetrahymanol from the pho- totrophic bacterium Rhodopseudomonas palustris: first report of a gammacerane triterpene from a prokaryote, J. Gen. Microbiol., 136, 2551–2553, 1990. Koga, Y., Nishihara, M., Morii, H., and Akagawa-Matsushita, M.: Ether polar lipids of methanogenic bacteria: structures, compar- ative aspects, and biosynthesis, Microbiol. Revs., 57, 164–182, 1993. Kononen, K., Kuparinen, J., Mäkelä, K., Laanemets, J., Pavelson, J., and Nõmman, S.: Initiation of cyanobacterial blooms in a frontal region at the entrance to the Gulf of Finland, Baltic Sea, Limnol. Oceanogr., 41, 98–112, 1996. Köster, J., Volkman, J. K., Rullkötter, J., Scholz-Böttcher, B. M., Rethmeier, J., and Fischer, U.: Mono-, di- and trimethyl- branched alkanes in cultures of the filamentous cyanobacterium Calothrix scopulorum, Org. Geochem., 30, 1367–1379, 1999. Kuever, J., Könnecke, M., Galushko, A., and Drzyzga, O.: Re- classification of Desulfobacterium phenolicum as Desulfobacula phenolica comb. nov. and description of strain SaxT as Desul- fotignum balticum gen. nov., sp. nov., Int. J. Syst. Evol. Micro- biol., 51, 171–177, 2001. Labrenz, M., Jost, G., and Jürgens, K.: Distribution of abundant prokaryotic organsims in the water column of the central Baltic Sea with an oxic-anoxic interface, Aquat. Microb. Ecol., 46, 177–190, 2007. Lang, I., Hodac, L., Friedl, T., and Feussner, I.: Fatty acid profiles and their distribution patterns in microalgae: a comprehensive Biogeosciences, 11, 7009–7023, 2014 www.biogeosciences.net/11/7009/2014/ C. Berndmeyer et al.: Biomarkers in the stratified water column 7021 analysis of more than 2000 strains from the SAG culture collec- tion, BMC Plant Biol., 11, 124–140, 2011. Leblond, J. D., Anderson, B., Kofink, D., Logares, R., Rengefors, K., and Kremp, A.: Fatty acid and sterol composition of two evo- lutinary closely related dinoflagellate morphospecies from cold Scandinavian brackish and freshwaters, Eur. J. Phycol., 41, 303– 311, 2007. Lee, R. F., Nevenzel, J. C., and Paffenhöfer, G.-A.: Importance of wax esters and other lipids in the marine food chain: phytoplank- ton and copepods, Mar. Biol., 9, 99–108, 1971. Lehtonen, K. K., Schiedek, D., Köhler, A., Lang, T., Vuorinen, P. J., Förlin, L., Baršiene˙, J., Pempkowiak, J., and Gercken, J.: The BEEP project in the Baltic Sea: Overview of results and outline for a regional biological effects monitoring strategy, Mar. Pollut. Bull., 53, 523–537, 2006. Liu, A., Zhu, T., Lu, X., and Song, L.: Hydrocarbon profiles and phylogenetic analyses of diversified cyanobacterial species, Appl. Energ., 111, 383–393, 2013. MacIntyre, S., Alldredge, A. L., and Gotschalk, C. C.: Accumula- tion of marine snow at density discontinuities in the water col- umn, Limnol. Oceanogr., 40, 449–468, 1995. Mansour, M. P., Volkman, J. K., Jackson, A. E., and Blackburn, S. I.: The fatty acid and sterol composition of five marine dinoflag- ellates, J. Phycol., 35, 710–720, 1999. Matsumoto, G. I., Akiyama, M., Watanuki, K., and Torii, T.: Unusual distribution of long-chain n-alkanes and n-alkenes in Antarctic soil, Org. Geochem., 15, 403–412, 1990. Matthäus, W. and Schinke, H.: The influence of river runoff on deep water conditions of the Baltic Sea, Hydrobiologia, 393, 1–10, 1999. Möllmann, C. and Köster, F. W.: Population dynamics of calanoid copepods and the implications of their predation by clupeid fish in the Central Baltic Sea, J. Plankton. Res., 24, 959–977, 2002. Möllmann, C., Kornilovs, G., and Sidrevics, L.: Long-term dynam- ics of main mesozooplankton species in the central Baltic Sea, J. Plankton. Res., 22, 2015–2038, 2000. Nanton, D. A. and Castell, J. D.: The effects of temperature and dietary fatty acids on the fatty acid composition of harpacticoid copepods, for use as a live food for fish larvae, Aquaculture, 175, 167–181, 1999. Nichols, P. D., Smith, G. A., Antworth, C. P., Hanson, R. S., and White, D. C.: Phospholipid and lipopolysaccharide normal and hydroxy fatty acids as potential signatures for methane-oxidizing bacteria, FEMS Microbiol. Ecol., 0, 327–335, 1985. Parkes, R. J. and Taylor, J.: The relationship between fatty acid dis- tributions and bacterial respiratory types in contemporary marine sediments, Estuar. Coast. Shelf Science, 16, 173–189, 1983. Parrish, C. C., Bodennec, G., and Gentien, P.: Time courses of intracellular and extracellular lipid classes in batch cultures of the toxic dinoflagellate, Gymnodinium cf. nagasakiense, Mar. Chem., 48, 71–82, 1994. Parrish, C. C., Abrajano, T. A., Budge, S. M., Helleur, R. J., Hudson, E. D., Pulchan, K., and Ramos, C.: Lipid and Phenolic Biomark- ers in Marine Ecosystems: Analysis and Applications, in: Mar. Chem., edited by: Wangersky, P. J., The Handbook of Environ- mental Chemistry, Springer Berlin Heidelberg, 193–223, 2000. Pinkart, H. C., Ringelberg, D. B., Piceno, Y. M., Macnaughton, S. J., and White, D. C.: Biochemical approaches to biomass measure- ments and community structure analysis, in: Manual of environ- mental microbiology, edited by: Hurt, C. J., ASM Press, Wash- ington, DC, 2, 101–113, 2002. Pugh, P. R.: Changes in the fatty acid composition of Coscinodiscus eccentricus with cultur-age and salinity, Mar. Biol., 11, 118–124, 1971. Reissmann, J. H., Burchard, H., Feistel, R., Hagen, E., Lass, H. U., Mohrholz, V., Nausch, G., Umlauf, L., and Wieczorek, G.: State- of-the-art review on vertical mixing in the Baltic Sea and conse- quences for eutrophication, Progr. Oceanogr., 82, 47–80, 2009. Risse-Buhl, U., Trefzeger, N., Seifert, A.-G., Schönborn, W., Gleixner, G., and Küsel, K.: Tracking the autochthonous carbon transfer in stream biofilm food webs, FEMS Microbiol. Ecol., 79, 118–131, 2011. Rush, D., Sinninghe Damsté, J. S., Poulton, S. W., Thamdrup, B., Garside, A. L., González, J. A., Schouten, S., Jetten, M. S. M., and Talbot, H. M.: Anaerobic ammonium-oxidising bacteria: a biological source of the bacteriohopanetetrol stereoisomer in ma- rine sediments, Geochim. Cosmochim. Ac., 140, 50–64, 2014. Sáenz, J. P., Wakeham, S. G., Eglinton, T. I., and Summons, R. E.: New constraints on the provenance of hopanoids in the marine geologic record: Bacteriohopanepolyols in marine suboxic and anoxic environments, Org. Geochem., 42, 1351–1362, 2011. Sargent, J. R., Gatten, R. R., and McIntosh, R.: Wax esters in the marine environment – their occurence, formation, transformation and ultimate fates, Mar. Chem., 5, 573–584, 1977. Schmale, O., Blumenberg, M., Kießlich, K., Jakobs, G., Bernd- meyer, C., Labrenz, M., Thiel, V., and Rehder, G.: Aero- bic methanotrophy within the pelagic redox-zone of the Got- land Deep (central Baltic Sea), Biogeosciences, 9, 4969–4977, doi:10.5194/bg-9-4969-2012, 2012. Schneider, B., Nausch, G., Kubsch, H., and Petersohn, I.: Accumu- lation of total CO2 during stagnation in the Baltic Sea deep water and its relationship to nutrient and oxygen concentrations, Mar. Chem., 77, 277–291, 2002. Schouten, S., Van der Maarel, M. J. E. C., Huber, R., and Sin- ninghe Damsté, J. S.: 2, 6, 10, 15, 19-Pentamethylicosenes in Methanolobus bombayensis, a marine methanogenic archaeon, and in Methanosarcina mazei, Org. Geochem., 26, 409–414, 1997. Schouten, S., Wakeham, S. G., and Sinninghe Damsté, J. S.: Ev- idence for anaerobic methane oxidation by archaea in euxinic waters of the Black Sea, Org. Geochem., 32, 1277–1281, 2001. Schubert, C. J., Coolen, M. J., Neretin, L. N., Schippers, A., Abbas, B., Durisch-Kaiser, E., Wehrli, B., Hopmans, E. C., Damste, J. S., Wakeham, S., and Kuypers, M. M.: Aerobic and anaerobic methanotrophs in the Black Sea water column, Environ. Micro- biol., 8, 1844–1856, 2006. Shamsudin, L.: Lipid and fatty acid composition of microalgae used in Malaysian aquaculture as live food for the early stage of pe- naeid larvae, J. Appl. Phycol., 4, 371–378, 1992. Shiea, J., Brassel, S. C., and Ward, D. M.: Mid-chain branched mono- and dimethyl alkanes in hot spring cyanobacterial mats: A direct biogenic source for branched alkanes in ancient sedi- ments?, Org. Geochem., 15, 223–231, 1990. Sinninghe Damsté, J. S., Kenig, F., Koopmans, M. P., Köster, J., Schouten, S., Hayes, J. M., and de Leeuw, J. W.: Evidence for gammacerane as an indicator of water column stratification, Geochim. Cosmochim. Ac., 59, 1895–1900, 1995. www.biogeosciences.net/11/7009/2014/ Biogeosciences, 11, 7009–7023, 2014 7022 C. Berndmeyer et al.: Biomarkers in the stratified water column Sinninghe Damsté, J. S., Rijpstra, W. I. C., Schouten, S., Fuerst, J. A., Jetten, M. S. M., and Strous, M.: The occurrence of hopanoids in planctomycetes: implications for the sedimentary biomarker record, Org. Geochem., 35, 561–566, 2004. Sinninghe Damsté, J. S., Rijpstra, Geenevasen, J. A. J., Strous, M., and Jetten, M. S. M.: Structural identification of ladderane and other membrane lipids of planctomycetes capable of anaero- bic ammonium oxidation (anammox), FEBS J., 272, 4270–4283, 2005. Stal, L. J., Albertano, P., Bergmann, B., von Bröckel, K., Gallon, J. R., Hayes, P. K., Sivonen, K., and Walsby, A. E.: BASIC: Baltic Sea cyanobacteria. An investigation of the structure and dynamics of water blooms of cyanobacteria in the Baltic Sea – re- sponses to a changing environment, Cont. Shelf Res., 23, 1695– 1714, 2003. Stock, A., Jürgens, K., Bunge, J., and Stoeck, T.: Protistan diversity in suboxic and anoxic waters of the Gotland Deep (Baltic Sea) as revealed by 18S rRNA clone libraries, Aquat. Microb. Ecol., 55, 267–284, 2009. Strom, S. L. and Morello, T. A.: Comparative growth rates and yields of ciliates and heterotrophic dinoflagellates, J. Plankton. Res., 20, 571–584, 1998. Summons, R., Bradley, A. S., Jahnke, L. L., and Waldbauer, J. R.: Steroids, triterpenoids and molecular oxygen, Phil. Trans. R. Soc. B, 361, 951–968, 2006. Sturt, H. F., Summons, R. E., Smith, K., Elvert, M., and Hin- richs, K. U.: Intact polar membrane lipids in prokaryotes and sediments deciphered by high-performance liquid chromatogra- phy/electrospray ionization multistage mass spectrometry – new biomarkers for biogeochemistry and microbial ecology, Rapid Commun. Mass Sp., 18, 617–628, 2004. Szewzyk, R. and Pfennig, N.: Complete oxidation of catechol by the strictly anaerobic sulfate-reducing Desulfobacterium cate- cholicum sp. nov., Arch. Microbiol., 147, 163–168, 1987. Talbot, H. M. and Farrimond, P.: Bacterial populations recorded in diverse sedimentary biohopanoid distributions, Org. Geochem., 38, 1212–1225, 2007. Talbot, H. M., Watson, D. F., Murrel, J. C., Carter, J. F., and Farri- mond, P.: Analysis of intact bacteriohopanepolyols from methan- otrophic bacteria by reversed-phase high-performance liquid chromatography-atmospheric pressure chemical ionisation mass spectrometry, J. Chrom., 921, 175–185, 2001. Talbot, H. M., Summons, R. E., Jahnke, L. L., Cockell, C. S., Rohmer, M., and Farrimond, P.: Cyanobacterial bacterio- hopanepolyol signatures from cultures and natural environmental settings, Org. Geochem., 39, 232–263, 2008. Taylor, J. and Parkes, J.: The cellular fatty acids of the sulphate- reducing bacteria, Desulfobacter sp., Desulfobulbus sp. and Desulfovibrio desulfuricans, J. Gen. Microbiol., 129, 3303– 3309, 1983. The R Foundation For Statistical Computing, R: A language and environment for statistical computing, R website, available at: http://www.r-project.org/, last access: 15 September 2014. Thiel, V., Peckmann, J., Richnow, H. H., Luth, U., Reitner, J., and Michaelis, W.: Molecular signals for anaerobic methane oxida- tion in Black Sea seep carbonates and a microbial mat, Mar. Chem. 73, 97–112, 2001 Thureborn, P., Lundin, D., Plathan, J., Poole, A. M., Sjöberg, B.-M., and Sjöling, S: a metagenomics transect into the deepest point of the Baltic Sea reveals clear stratification of microbial functional capacities, PLOS One, 8, e74983, doi:10.1371/journal.pone.0074983, 2013. Tornabene, T. G. and Langworthy, T. A.: Diphytanyl and dibiphy- tanyl glycerol ether lipids of methanogenic archaebacteria, Sci- ence, 203, 51–53, 1979. Tornabene, T. G., Langworthy, T. A., Holzer, G., and Orò, J.: Squalenes, phytanes and other isoprenoids as major neutral lipids of methanogenic and thermoacidophilic “archaebacteria”, J. Mol. Evol., 13, 73–83, 1979. Tyson, R. V. and Pearson, T. H.: Modern and ancient continental shelf anoxia: an overview, Geological Society, London, Special Publications, 58, 1–24, 1991. Vahtera, E., Conley, D. J., Gustafsson, B. G., Kuosa, H., Pitkänen, H., Savchuk, O. P., Tamminen, T., Viitasalo, M., Voss, M., Was- mund, N., and Wulff, F.: Internal ecosystem feedbacks enhance nitrogen-fixing cyanobacteria blooms and complicate manage- ment in the Baltic Sea, Ambio, 36, 186–194, 2007. Vainshtein, M., Hippe, H., and Kroppenstedt, R. M.: Cellular fatty acid composition of Desulfovibrio species and its use in classi- fication of sulfate-reducing bacteria, Syst. Appl. Microbiol., 15, 554–556, 1992. Vestal, R. J. and White, D. C.: Lipid analysis in microbial ecology, BioScience, 39, 535–541, 1989. Volkman, J. K.: A review of sterol markers for marine and terrige- nous organic matter, Org. Geochem., 9, 83–99, 1986. Volkman, J. K., Barrett, S. M., Dunstan, G. A., and Jeffrey, S. W.: Geochemical significance of the occurrence of dinosterol and other 4-methyl sterols in a marine diatom, Org. Geochem., 20, 7–15, 1993. Volkman, J. K., Barrett, S. M., Blackburn, S. I., Mansour, M. P., Sikes, E. L., and Gelin, F.: Microalgal biomarkers: A review of recent research developments, Org. Geochem., 29, 1163–1179, 1998. Wakeham, S. G.: Reduction of stenols to stanols in particulate mat- ter at oxic-anoxic boundaries in sea water, Nature, 342, 787–790, 1989. Wakeham, S. G., Amann, R., Freeman, K. H., Hopmans, E. C., Jør- gensen, B. B., Putnam, I. F., Schouten, S., Sinninghe Damsté, J. S., Talbot, H. M., and Woebken, D.: Microbial ecology of the stratified water column of the Black Sea as revealed by a compre- hensive biomarker study, Org. Geochem., 38, 2070–2097, 2007. Wakeham, S. G., Turich, C., Schubotz, F., Podlaska, A., Xiaona, N. L., Varela, R., Astor, Y., Sáenz, J. P., Rush, D., Sinninghe Damsté, J. S., Summons, R. E., Scranton, M. I., Taylor, G. T., and Hinrichs, K. U.: Biomarkers, chemistry and microbiology show chemoautotrophy in a multilayer chemocline in the Cariaco Basin, Deep Sea Res. Pt. I, 163, 133–156, 2012. Wasmund, N., Nausch, G., and Matthäus, W.: Phytoplankton spring blooms in the southern Baltic Sea – spatio temporal development and long-term trends, J. Plankton. Res., 20, 1099–1117, 1998. Welander, P. V., Coleman, M., Sessions, A. L., Summons, R. E., and Newman, D. K.: Identification of a methylase required for 2-methylhopanoid production and implications for the interpre- tation of sedimentary hopanes, PNAS, 107, 8537–8542, 2010. Widdel, F.: New types of acetate-oxidizing, sulfate-reducing Desul- fobacter species, D. hydrogenophilus sp. nov., D. latus sp. nov., and D. curvatus sp. nov., Arch. Microbiol., 148, 286–291, 1987. Biogeosciences, 11, 7009–7023, 2014 www.biogeosciences.net/11/7009/2014/ C. Berndmeyer et al.: Biomarkers in the stratified water column 7023 Xie, S., Liu, X.-L., Schubotz, F., Wakeham, S. G., and Hinrichs, K. U.: Distribution of glycerol ether lipids in the oxygen min- imum zone of the Eastern Tropical North Pacific Ocean, Org. Geochem., 71, 60–71, 2014. Zander, J. M., Caspi, E., Pandey, G. N. and Mitra, C. R.: The pres- ence of tetrahymanol in Oleandra wallichii, Phytochemistry, 8, 2265–2267, 1969. www.biogeosciences.net/11/7009/2014/ Biogeosciences, 11, 7009–7023, 2014