Formate, acetate, and propionate as substrates for sulfate reduction in sub-arctic sediments of Southwest Greenland

Volatile fatty acids (VFAs) are key intermediates in the anaerobic mineralization of organic matter in marine sediments. We studied the role of VFAs in the carbon and energy turnover in the sulfate reduction zone of sediments from the sub-arctic Godthåbsfjord (SW Greenland) and the adjacent continental shelf in the NE Labrador Sea. VFA porewater concentrations were measured by a new two-dimensional ion chromatography-mass spectrometry method that enabled the direct analysis of VFAs without sample pretreatment. VFA concentrations were low and surprisingly constant (4–6 μmol L−1 for formate and acetate, and 0.5 μmol L−1 for propionate) throughout the sulfate reduction zone. Hence, VFAs are turned over while maintaining a stable concentration that is suggested to be under a strong microbial control. Estimated mean diffusion times of acetate between neighboring cells were <1 s, whereas VFA turnover times increased from several hours at the sediment surface to several years at the bottom of the sulfate reduction zone. Thus, diffusion was not limiting the VFA turnover. Despite constant VFA concentrations, the Gibbs energies (ΔGr) of VFA-dependent sulfate reduction decreased downcore, from −28 to −16 kJ (mol formate)−1, −68 to −31 kJ (mol acetate)−1, and −124 to −65 kJ (mol propionate)−1. Thus, ΔGr is apparently not determining the in-situ VFA concentrations directly. However, at the bottom of the sulfate zone of the shelf station, acetoclastic sulfate reduction might operate at its energetic limit at ~ −30 kJ (mol acetate)−1. It is not clear what controls VFA concentrations in the porewater but cell physiological constraints such as energetic costs of VFA activation or uptake could be important. We suggest that such constraints control the substrate turnover and result in a minimum ΔGr that depends on cell physiology and is different for individual substrates.

[1]  J. Amend,et al.  Chemolithotrophy in the continental deep subsurface: Sanford Underground Research Facility (SURF), USA , 2014, Front. Microbiol..

[2]  B. Jørgensen,et al.  Direct analysis of volatile fatty acids in marine sediment porewater by two‐dimensional ion chromatography‐mass spectrometry , 2014 .

[3]  B. Jørgensen,et al.  Determination of dissimilatory sulfate reduction rates in marine sediment via radioactive 35S tracer , 2014 .

[4]  J. Amend,et al.  13. Energetic constraints on life in marine deep sediments , 2014 .

[5]  C. Martens,et al.  Apparent Minimum Free Energy Requirements for Methanogenic Archaea and Sulfate-Reducing Bacteria in an Anoxic Marine Sediment , 2013 .

[6]  B. Jørgensen,et al.  Quantifying the degradation of organic matter in marine sediments: A review and synthesis , 2013 .

[7]  J. Kallmeyer,et al.  Sulfate reduction controlled by organic matter availability in deep sediment cores from the saline, alkaline Lake Van (Eastern Anatolia, Turkey) , 2013, Front. Microbiol..

[8]  Heath J. Mills,et al.  Microbial activity in the marine deep biosphere: progress and prospects , 2013, Front. Microbiol..

[9]  B. Thamdrup,et al.  Identification of acetate-oxidizing bacteria in a coastal marine surface sediment by RNA-stable isotope probing in anoxic slurries and intact cores. , 2013, FEMS microbiology ecology.

[10]  B. Jørgensen,et al.  Microbial life under extreme energy limitation , 2013, Nature Reviews Microbiology.

[11]  M. Krüger,et al.  Geochemistry and Microbial Populations in Sediments of the Northern Baffin Bay, Arctic , 2013 .

[12]  David C. Smith,et al.  Global distribution of microbial abundance and biomass in subseafloor sediment , 2012, Proceedings of the National Academy of Sciences.

[13]  Bo Barker Jørgensen,et al.  Aerobic Microbial Respiration in 86-Million-Year-Old Deep-Sea Red Clay , 2012, Science.

[14]  J. Morse,et al.  Examination and Refinement of the Determination of Aqueous Hydrogen Sulfide by the Methylene Blue Method , 2011 .

[15]  S. D’Hondt,et al.  Gibbs energies of reaction and microbial mutualism in anaerobic deep subseafloor sediments of ODP Site 1226 , 2010 .

[16]  R. Parkes,et al.  Role of sulfate reduction and methane production by organic carbon degradation in eutrophic fjord sediments (Limfjorden, Denmark) , 2010 .

[17]  Q. Jin,et al.  Cellular energy conservation and the rate of microbial sulfate reduction , 2009 .

[18]  David C. Smith,et al.  New cell extraction procedure applied to deep subsurface sediments , 2008 .

[19]  Michael Schlüter,et al.  Rhizon sampling of porewaters near the sediment‐water interface of aquatic systems , 2005 .

[20]  B. Jørgensen,et al.  A cold chromium distillation procedure for radiolabeled sulfide applied to sulfate reduction measurements , 2004 .

[21]  B. Jørgensen,et al.  Anaerobic methane oxidation and a deep H2S sink generate isotopically heavy sulfides in Black Sea sediments , 2004 .

[22]  R. Parkes,et al.  Recent studies on bacterial populations and processes in subseafloor sediments: A review , 2000 .

[23]  C. Martens,et al.  Thermodynamic control on hydrogen concentrations in anoxic sediments , 1998 .

[24]  B. Schink Energetics of syntrophic cooperation in methanogenic degradation , 1997, Microbiology and molecular biology reviews : MMBR.

[25]  E. Shock,et al.  Inorganic species in geologic fluids: correlations among standard molal thermodynamic properties of aqueous ions and hydroxide complexes. , 1997, Geochimica et cosmochimica acta.

[26]  C. Martens,et al.  Determination of low-molecular-weight organic acid concentrations in seawater and pore-water samples via HPLC , 1997 .

[27]  E. Shock,et al.  Organic acids in hydrothermal solutions: standard molal thermodynamic properties of carboxylic acids and estimates of dissociation constants at high temperatures and pressures. , 1995, American journal of science.

[28]  D. Burdige The biogeochemistry of manganese and iron reduction in marine sediments , 1993 .

[29]  E. Oelkers,et al.  SUPCRT92: a software package for calculating the standard molal thermodynamic properties of minerals, gases, aqueous species, and reactions from 1 to 5000 bar and 0 to 1000 ° C , 1992 .

[30]  D. Shaw,et al.  Acetate in recent anoxic sediments: Direct and indirect measurements of concentration and turnover rates , 1990 .

[31]  Everett L. Shock,et al.  Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures: Standard partial molal properties of organic species , 1990 .

[32]  J. Middelburg A simple rate model for organic matter decomposition in marine sediments , 1989 .

[33]  Everett L. Shock,et al.  Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures: Correlation algorithms for ionic species and equation of state predictions to 5 kb and 1000°C , 1988 .

[34]  D. Capone,et al.  Comparison of microbial dynamics in marine and freshwater sediments: Contrasts in anaerobic carbon catabolism1 , 1988 .

[35]  F. Widdel,et al.  Acetate oxidation to CO2 in anaerobic bacteria via a novel pathway not involving reactions of the citric acid cycle , 1986, Archives of Microbiology.

[36]  R. Parkes,et al.  Demonstration, using Desulfobacter sp., of two pools of acetate with different biological availabilities in marine pore water , 1984 .

[37]  W. Reeburgh,et al.  Biogeochemistry of acetate in anoxic sediments of Skan Bay, Alaska , 1984 .

[38]  Derek G. Leaist,et al.  Diffusion in dilute aqueous acetic acid solutions at 25°C , 1984 .

[39]  A. Zehnder,et al.  Kinetics of Sulfate and Acetate Uptake by Desulfobacter postgatei , 1984, Applied and environmental microbiology.

[40]  R. Parkes,et al.  Analysis of volatile fatty acids by ion-exclusion chromatography, with special reference to marine pore water , 1983 .

[41]  W. Reeburgh Rates of biogeochemical processes in anoxic sediments. , 1983 .

[42]  P. Maloney Relationship between phosphorylation potential and electrochemical H+ gradient during glycolysis in Streptococcus lactis , 1983, Journal of bacteriology.

[43]  T. Blackburn,et al.  Turnover of 14C-labelled acetate in marine sediments , 1982 .

[44]  B. Jørgensen Mineralization of organic matter in the sea bed—the role of sulphate reduction , 1982, Nature.

[45]  D. B. Nedwell,et al.  Microbial Metabolism of Acetate, Propionate and Butyrate in Anoxic Sediment from the Colne Point Saltmarsh, Essex, U.K. , 1982 .

[46]  H. Helgeson,et al.  Theoretical prediction of the thermodynamic behavior of aqueous electrolytes by high pressures and temperatures; IV, Calculation of activity coefficients, osmotic coefficients, and apparent molal and standard and relative partial molal properties to 600 degrees C and 5kb , 1981 .

[47]  B. B. J�rgensen,et al.  Volatile Fatty Acids and Hydrogen as Substrates for Sulfate-Reducing Bacteria in Anaerobic Marine Sediment , 1981, Applied and environmental microbiology.

[48]  T. Blackburn,et al.  A method for the analysis of acetate turnover in a coastal marine sediment , 1980, Microbial Ecology.

[49]  R. Thauer,et al.  Energy conservation in chemotrophic anaerobic bacteria , 1977, Bacteriological reviews.

[50]  B. B. Jørgensen,et al.  The sulfur cycle of a marine sediment model system , 1974 .

[51]  A. Larimore,et al.  Energy conservation. , 1972, Science.

[52]  H. Helgeson,et al.  Thermodynamics of hydrothermal systems at elevated temperatures and pressures , 1969 .

[53]  Joel D. Cline,et al.  SPECTROPHOTOMETRIC DETERMINATION OF HYDROGEN SULFIDE IN NATURAL WATERS1 , 1969 .

[54]  P. Mitchell CHEMIOSMOTIC COUPLING IN OXIDATIVE AND PHOTOSYNTHETIC PHOSPHORYLATION , 1966, Biological reviews of the Cambridge Philosophical Society.

[55]  P. Lyons,et al.  Diffusion in Dilute Aqueous Acetic Acid Solutions , 1965 .

[56]  J. Bendtsen,et al.  Heat sources for glacial melt in a sub‐Arctic fjord (Godthåbsfjord) in contact with the Greenland Ice Sheet , 2011 .

[57]  Chemie Volatile Fatty Acids , 2010 .

[58]  B. Jørgensen,et al.  Acetate, lactate, propionate, and isobutyrate as electron donors for iron and sulfate reduction in Arctic marine sediments, Svalbard. , 2007, FEMS microbiology ecology.

[59]  F. Widdel,et al.  Dissimilatory Sulfate- and Sulfur-Reducing Prokaryotes , 2006 .

[60]  E. Buch Present oceanographic conditions in Greenland Waters , 2002 .

[61]  E. Oelkers,et al.  Calculation of the thermodynamic properties of aqueous species at high pressures and temperatures. Effective electrostatic radii, dissociation constants and standard partial molal properties to 1000 °C and 5 kbar , 1992 .

[62]  H. Helgeson,et al.  Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures; revised equations of state for the standard partial molal properties of ions and electrolytes , 1988 .

[63]  D. Christensen Determination of substrates oxidized by sulfate reduction in intact cores of marine sediments1 , 1984 .

[64]  S. Ferguson,et al.  Proton electrochemical gradients and energy-transduction processes. , 1982, Annual review of biochemistry.

[65]  T. Dyman,et al.  A comparison of methods for the quantification of assemblage zones , 1978 .

[66]  Bo Barker J⊘rgensen A comparison of methods for the quantification of bacterial sulfate reduction in coastal marine sediments: III. Estimation from chemical and bacteriological field data , 1978 .

[67]  Jørgensen BoBarker A comparison of methods for the quantification of bacterial sulfate reduction in coastal marine sediments , 1978 .

[68]  K. U Microbial Metabolism of Acetate , Propionate and Butyrate in Anoxic Sediment from the Colne Point Saltmarsh , Essex , , 2022 .