sulfur cycling and methane oxidation · for an overview of the processes and pathways involved in...

41
See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/226248577 Sulfur Cycling and Methane Oxidation Chapter · February 2006 DOI: 10.1007/3-540-32144-6_8 CITATIONS 97 READS 301 2 authors: Bo Barker Jørgensen Aarhus University 402 PUBLICATIONS 29,477 CITATIONS SEE PROFILE Sabine Kasten Alfred Wegener Institute Helmholtz Ce… 123 PUBLICATIONS 2,084 CITATIONS SEE PROFILE All in-text references underlined in blue are linked to publications on ResearchGate, letting you access and read them immediately. Available from: Bo Barker Jørgensen Retrieved on: 01 November 2016

Upload: others

Post on 21-Aug-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

Seediscussions,stats,andauthorprofilesforthispublicationat:https://www.researchgate.net/publication/226248577

SulfurCyclingandMethaneOxidation

Chapter·February2006

DOI:10.1007/3-540-32144-6_8

CITATIONS

97

READS

301

2authors:

BoBarkerJørgensen

AarhusUniversity

402PUBLICATIONS29,477CITATIONS

SEEPROFILE

SabineKasten

AlfredWegenerInstituteHelmholtzCe…

123PUBLICATIONS2,084CITATIONS

SEEPROFILE

Allin-textreferencesunderlinedinbluearelinkedtopublicationsonResearchGate,

lettingyouaccessandreadthemimmediately.

Availablefrom:BoBarkerJørgensen

Retrievedon:01November2016

Page 2: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

271

8 Sulfur Cycling and Methane OxidationBO BARKER JØRGENSEN AND SABINE KASTEN

This chapter deals with the biogeochemical trans-formations of sulfur and methane in marinesediments during early diagenesis. The term ‘earlydiagenesis’ refers to the whole range of post-depositional processes that take place in aquaticsediments and are coupled either directly orindirectly to the degradation of organic matter. Wefocus on the processes that drive sulfate reduc-tion together with the manifold associated bioticand abiotic reactions that make up the sedimen-tary sulfur cycle – including the various pathwaysof sulfide oxidation. Furthermore, we give anoverview of the quantitative significance ofmicrobial sulfate reduction for the mineralizationof organic matter and oxidation of methane indifferent depositional environments and discussthe different approaches to quantify sulfate reduc-tion through radiotracer measurements or mode-ling of pore-water concentration profiles. Assedimentary pyrite represents the most importantsink for seawater sulfate, the mechanisms of pyriteformation are discussed. The sulfidization ofsediment organic matter is another sink for sulfurin the modern ocean, but will not be discussedhere. For an overview of the processes andpathways involved in the incorporation of sulfurinto organic matter, we refer to Orr and White(1990), Krein and Aizenshtat (1995), Schouten etal. (1995), and Werne et al. (2004).

Due to the profound alteration of the primarysediment geochemistry across and below thesulfate/methane transition (SMT, also called thesulfate/methane interface), where the process ofanaerobic oxidation of methane (AOM) takesplace, we also dedicate a part of this chapter tothe processes of mineral dissolution and preci-pitation occurring at and below the SMT. We willdemonstrate that sulfate reduction driven byAOM significantly alters primary mineralassociations accompanied by strong pertur-bations of mineralogical, isotopic and rockmagnetic signatures. These diagenetic processes

have a profound impact on the preservation ofnumerous paleoceanographic proxy variables andare therefore relevant for the interpretation of thegeological record.

Reviews of the sulfur cycle from a biogeo-chemical or microbiological perspective haverecently been presented, e.g. by Amend et al.(2004) and Canfield et al. (2005). The latter authorsdescribe the relationship between the differentmicroorganisms metabolizing sulfur compoundsand their environment and emphasize the orga-nisms as well as the biogeochemical processes.

8.1 Introduction

With 1.3 · 109 Tg (teragram = megaton = 1012 g) ofsulfur present as sulfate, the oceans represent oneof the largest sulfur pools on earth (Vairavamurthyet al. 1995). The main influx of sulfur to the oceansoccurs via river water carrying the products ofmechanical and chemical weathering of conti-nental rocks. Relative to this fluvial input, theatmospheric transport of sulfur is of minorimportance. It mainly consists of recycled oceanicsulfate from seaspray, volcanic sulfur gases, H2Sreleased by sulfate-reducing bacteria, organic S-bearing compounds released into seawater andsubsequently into the atmosphere by phyto-plankton, and anthropogenic emissions of sulfurdioxide. Due to the oxic conditions that prevail inthe world’s oceans, the dominant sulfur species inseawater is by far the sulfate ion (SO4

2-). Sulfate isthe second most abundant anion next to chlorideand has a concentration of 29 mM (2.71 g/kg) inocean water.

Marine sediments are the main sink for sea-water sulfate which demonstrates that the sedi-mentary sulfur cycle is a major component of theglobal sulfur cycle. The most important mecha-nisms for removing sulfate from the oceans to the

Page 3: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

272

sediments are (1) the bacterial reduction of sulfateto hydrogen sulfide, which subsequently reactswith iron to form sulfide minerals, particularlypyrite (FeS2), (2) the formation of organic sulfur,i.e. the incorporation of sulfur into sedimentaryorganic matter during early diagenesis, and (3) theprecipitation of calcium sulfate minerals inevaporites (Vairavamurthy et al. 1995). Withrespect to the relative importance of eachpathway, Vairavamurthy et al. (1995) point out thatalthough marine evaporites were important sinksfor sulfate from the Late Precambrian to the lateTertiary, their rate of formation in today’s oceans,over the last few million years, is quantitativelyinsignificant. Thus, they conclude that the burialof sulfide minerals and, to a less extent, of organicsulfur represents the major sink for oceanic sulfurin the modern ocean.

8.2 Sulfate Reduction and theDegradation of Organic Matter

Throughout the water column of the ocean, tenbillion tons of organic particles derived fromplankton production in the photic surface layersteadily sink towards the sea floor. The annualdeposition of organic material on the sea floor is

also about ten billion tons, i.e. of the same magni-tude as the total organic particle pool in theocean. As the particles sink through the watercolumn, the organic material is gradually degradedand respired back to CO2 and nutrients byzooplankton and by microorganisms. Thus, withincreasing distance from the coast and withincreasing water depth, a gradually decreasingfraction of the sinking particles remains toultimately land on the sea floor. On the shallowcontinental shelves this fraction is generally inthe range of 20-50% of the phytoplanktonproductivity in the overlying water (Jørgensen1983). In the deep sea the fraction is only 1-2%(Jahnke 1996). Deep sea sediments, therefore, playonly a minor quantitative role in the oceaniccarbon and nutrient cycles, as the following dis-cussion will show.

As the particulate organic matter is depositedon the sea floor it is immediately attacked by abroad range of organisms that all contribute to itsdegradation and gradual mineralization. At the sedi-ment surface, macrofauna plays a particular role bymechanically disintegrating the detritus and bymixing of the upper sediment layer through theirburrowing and feeding activity (bioturbation),whereby the organic material becomes repeatedlyexposed to oxygen. The benthic fauna also irrigatesthe burrows in order to transport oxygen down for

Fig. 8.1 Schematic representation of the biogeochemical zonation in marine sediments. The names of the main zones wereproposed by Froelich et al. (1979) and Berner (1981, in parenthesis). The depth scale is quasi-logarithmic; the exact depths,however, vary strongly and increase from the shelf to the deep sea. The pore water chemistry shows relevant dissolvedspecies. Peak heights and concentration scales are arbitrary. The chemical profiles reflect the depth sequence of thedominant mineralization processes through which organic matter is oxidized to CO2 (Modified from Froelich et al. 1979).

Page 4: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

273

8.2 Sulfate Reduction and the Degradation of Organic Matter

respiration, and possibly food particles, andthereby extend the oxic zone (i.e. the zone contai-ning O2) deeper down into the sediment. Due tothis macrofaunal activity, an extended surface layerof sediment remains oxidized (i.e. with positiveredox potential and with many chemical speciessuch as iron or manganese in their oxidized state).The macrofauna thus has a critical influence on theearly diagenesis in sediments and affects both thepathways and the rates of many processes.

8.2.1 Geochemical Zonation

Apart from the heterogeneous chemical structurecaused by faunal activity, marine sediments havea distinct biogeochemical zonation of the mainaerobic and anaerobic mineralization processes.Fig. 8.1 shows how the dominant oxidants formineralization change with depth, regulated partlyby their concentration in sea water and partly bythe energy yield of the process by which they areconsumed (cf. Chapter 5). Many microorganismsgain energy from the oxidation of organic matterwith an external oxidant (electron acceptor).Thermodynamically, oxygen is the most favorableelectron acceptor. The supply of oxygen from sea-water to the sediment is, however, highly transport-limited. In coastal marine sediments or in oceanareas of high productivity, e.g. upwelling regions,a high organic matter flux or low oxygen contentof the bottom water reduces the thickness of theoxic surface layer of the sediment to only a fewmm or cm. In deep sea sediments the depth ofoxygen penetration may be a dm or a m or evenmore (Chapter 6).

Where oxygen has been consumed by aerobicrespiration, the sediment is anoxic, i.e. O2-free, andmicroorganisms utilize other terminal electronacceptors for the mineralization of organic matter.Listed in an order of decreasing energy gain theseare: nitrate (NO3

-), manganese oxides (representedby Mn(IV)), iron oxides (represented by Fe(III)),and sulfate (SO4

2-) (Fig. 8.1). The sediment layerwhere NO3

-, Mn(IV) and Fe(III) reductionpredominate has been termed the suboxic zone(Froelich et al. 1979). Although these electronacceptors are energetically more favorable thansulfate, they are usually less important biogeo-chemically because of their limited supply to thesediments. The processes of O2, NO3

- and Fe(III)reduction are discussed in detail in Chapter 6 and 7.

Below the suboxic zone, sulfate reduction isthe main pathway of organic carbon oxidation and

it generally extends many meters down into thesediment. The high concentration of sulfate inseawater makes it a dominant electron acceptor(Henrichs and Reeburgh 1987). With an averageconcentration of about 29 mmol/l (Vairavamurthyet al. 1995), sulfate concentrations in seawater aremore than two orders of magnitude higher than infreshwater (about 0.1 mmol/l) (Bowen 1979). Themineralization of organic material by sulfate inmarine sediments is, therefore, much more impor-tant than in freshwater.

In the even deeper subsurface sediment, wherealso sulfate is exhausted, there is little net oxi-dation of the organic carbon but rather adegradation to CH4 and CO2. As a stable endproduct of carbon degradation in the absence ofoxidants, methane tends to accumulate in deepsub-surface sediments from where it slowlydiffuses up towards the sulfate zone. Upon entryinto the lower sulfate zone, however, also methanebecomes oxidized completely to CO2. The organicmaterial initially deposited on the sedimentsurface thus undergoes an efficient progressingdegradation as it becomes buried deeper anddeeper down through this sequence of diageneticprocesses. Only a small fraction escapes minera-lization and remains after thousands or millions ofyears to contribute to the great pool of fossil or-ganic carbon in marine sediments.

The main degradation of organic material takesplace as it gradually becomes buried downthrough the oxic and suboxic zones. The fractionthat still remains once it reaches the sulfatereduction zone therefore depends on how deepoxygen, nitrate, manganese and iron reductionpredominate. In coastal sediments with highorganic sedimentation these oxidants are rapidlydepleted and the main sulfate reduction zonestarts already a few cm below the sediment sur-face. Although it is not so apparent from thechemical zonation, sulfate reduction also occurs insuboxic and even in oxic sediment where theproduced sulfide is rapidly reoxidized (Jørgensenand Bak 1991). In the deep sea, the metal oxidesmay be exhausted only several meters below thesurface and little organic material is available onceit becomes buried down into the sulfate reductionzone. The overall sulfate reduction in sediments istherefore very sensitive to the organic depositionrate and is geographically strongly shiftedtowards the continental margins and shelf sedi-ments. In fact, a large part of the deep sea floorlies under low-productivity ocean regions where

Page 5: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

274

the organic sedimentation is so scanty that sulfatereduction is hardly detectable throughout theentire sediment column. Fig. 8.2 shows such anexample, with data taken from an expedition of theOcean Drilling Program in the eastern tropicalPacific Ocean.

Fig. 8.2 shows some of the main chemicalchanges with depth in the 320 m thick depositoverlying old ocean crust of Miocene age. Themineralization of organic carbon and organicnitrogen is apparent from the increase in porewater concentration of total CO2 and ammonium. Itis striking that these concentrations drop backtowards sea water values at the bottom of thesediment column. This is due to a slow advectionof sea water through the porous ocean crustwhich removes these products of mineralizationand thereby influences the entire chemistry of thesea bed and its exchange with sea water in thispart of the Pacific Ocean. Surprisingly, the micro-bial activity at the bottom of the sediment columnis so low that even nitrate (and possibly oxygen)remains in the crustal fluid and provides anadditional source of oxidant from below (Fig. 8.2,middle frame). The Mn2+ gradient in the upper 100m and its slight increase at 200-300 mbsf show theimportance of manganese reduction throughoutthis deep sea sediment. Consequently, sulfatereduction is inhibited by electron acceptors withhigher energy yield, and even a sediment layer of320 m thickness generates only a marginal drop insulfate concentration (Fig. 8.2, right frame).

Methane producing microorganisms are evenmore strongly inhibited and, although presentthroughout the entire sediment deposit, methanedoes not exceed a trace concentration of 0.2 µM.

8.2.2 Dissimilatory Sulfate Reduction

Two types of sulfate reduction can be distin-guished: (1) assimilatory sulfate reduction whichserves the biosynthesis of sulfur-containingorganic compounds that are part of the cellbiomass, and (2) dissimilatory sulfate reductionfrom which microorganisms conserve energy andrelease H2S to the environment. With respect tothe mineralization of organic matter and the mainsource of H2S in sediments, only dissimilatorysulfate reduction plays a role and is the processreferred to when we discuss microbial sulfatereduction. The sulfate reducing bacteria usesulfate as the terminal electron acceptor for theiranaerobic respiration (see Chapter 5). As energyand carbon source they use mostly short-chainfatty acids (acetate, formate, propionate, butyrate)and other small organic molecules that areproduced from the degradation and fermentationof sediment organic matter. H2 is also a product offermentation and serves as an important energysource for sulfate reduction in marine sediments.

Most of the known dissimilatory sulfate redu-cers are bacteria, but also some thermophilicarchaea belong to this group (Rabus et al. 2004;Stetter et al. 1993). For an overview of the most

Fig. 8.2 Pore water chemistry of a deep sea sediment from the eastern tropical Pacific Ocean (ODP Site 1225). Thecore was obtained during Leg 201 of the Ocean Drilling Program and spans the entire sediment deposit, from thesediment surface at 3760 m water depth down to the 11 million years old basaltic crust (hatched) at 320 mbsf (meterbelow sea floor). Data from D’Hondt et al. (2003).

Page 6: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

275

8.2 Sulfate Reduction and the Degradation of Organic Matter

common sulfate-reducing bacteria in marinesediments we refer to Widdel (1988). Besidessulfate there are also other oxidized sulfurcompounds, e.g. thiosulfate (S2O3

2-) and elementalsulfur (S0), that can serve as the terminal electronacceptor (Ehrlich 1996). They are, however, not ofsimilar quantitative importance as sulfate.

Dissimilatory sulfate reduction can be descri-bed by the following net equation:

2 [CH2O] + SO42- → 2 HCO3

- + H2S (8.1)

where [CH2O] simply represents organic material.When the sulfate reducing bacteria are growing, apart of the [CH2O] will be assimilated to producecell material. Hydrogen, in contrast, is used onlyas an energy source and the cell assimilates CO2or an organic subtrate:

4 H2 + SO42- + 2 H+ → 4 H2O + H2S (8.2)

8.2.3 Sulfate Reduction andOrganic Carbon Mineralization

Continental margin sediments play a major role forthe overall budget of the global carbon cycle inthe modern ocean. Most burial of organic carbontakes place on the continental shelf, in particularin deltaic and other coastal sediments (Berner1982; Hedges and Keil 1995). Thus, 82% of theburied organic carbon is stored in shelf sediments

and 16% in continental slope sediments (Wollast1998). Only 2% of the marine organic carbon burialtakes place in deep sea sediments. It is importantto note, however, that during glaciations the sealevel has in the past dropped by more than 100 mand thereby exposed vast shelf areas to erosionand transport of stored material into deeper water.The current burial of organic material in shelfsediments is, therefore, in a geological perspectivea transient state that may undergo major reallo-cation during a future glaciation. On the otherhand, in case of a long-term global warming thesea level would rise and the area and sedimentaccumulation on the continental shelves wouldincrease.

The aerobic and anaerobic mineralization oforganic material is also strongly focused towardsthe ocean margins. Table 8.1 presents examplesof the areal rates of organic carbon degradationin a number of representative sediments, rangingfrom coastal embayments to the deep sea. Thetable also shows the fraction of the overall orga-nic carbon mineralization that takes place by sul-fate reduction. Sulfate reduction predominatesespecially in sediments underlying highly pro-ductive and/or oxygen-depleted coastal waters, e.g.in the Black Sea, Cape Lookout Bight, or the up-welling areas of the Chilean shelf. In other shelfsediments sulfate reduction generally accounts for25-50% of the mineralization (Jørgensen 1982). Therelative importance of sulfate reduction drops to<1% as we move down the continental slope and it

Table 8.1 Mineralization of deposited organic carbon (Corg) in marine sediments and the role of sulfate reduction.Data range from the shelf to the deep sea and are listed according to decreasing significance of sulfate reduction.The Black Sea is included at the end as an example of an anoxic basin totally dominated by sulfate reduction. Thedata are based on radiotracer measurements of sulfate reduction rates (* represents a modeled rate).

Water depth Rate of Corg Corg mineralized degradation by sulfate

reduction[m] [mmol m-2 d-1]

Cape Lookout Bight 9 114 72% Crill and Martens (1987)Chilean shelf 34-122 10 56-79% Thamdrup and Canfield

(1996)Baltic Sea 16 9.8 44% Jørgensen (1996)

Gulf of Maine 50-300 10.7 43% Christensen (1989)St. Lawrence Estuary 335 10 26% Edenborn et al. (1987)

E. South Atlantic 3700 2.3-3.2 3-16% Ferdelman et al. (1999)

Black Sea, anoxic 130-1176 1.30-2.86 ~100% Jørgensen et al. (2004)

Location Reference

Eastern tropical Pacific

3760 0.5 0.05%* D'Hondt et al. (2004) Jahnke (1996)

Page 7: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

276

is insignificant in vast regions of the deep sea.The additional source of H2S from the anaerobicdegradation of organic sulfur plays a minor rolerelative to this dissimilatory sulfate reduction.

Radiotracer measurements of sulfate reductionrates have mainly been carried out in shallowerwaters, so that the data base for sediments below500 m water depth is very limited. A detailed studywas done by Ferdelman et al. (1999) between 855and 4766 m water depth on the continental marginof southwest Africa. This area forms part of theBenguela upwelling system. The authors demon-strated that the depth-integrated sulfate reductionrates over the upper 20 cm of the sedimentstrongly correlated with the concentrations oforganic carbon in the surface sediments. Theyfurther estimated that sulfate reduction at forexample 3700 m water depth accounts for 3 to 16 %(Table 8.1) of total oxygen consumption. Thus,even on the lower continental slope in this region,a small but significant fraction of organic matter isdegraded anaerobically through sulfate reduction.

Canfield (1993) and Canfield et al. (2005) havecompiled published data on oxygen uptake andsulfate reduction in marine sediments and groupedthese into different coastal types and into diffe-rent depth regions of the ocean. Based on theirdata, Fig. 8.3 presents the relationship betweenaerobic mineralization and sulfate-based anaerobicmineralization in sediment types comprising theentire ocean floor. The sulfate reduction rates

were determined experimentally by the radiotracermethod in shelf and slope sediments whereas theywere modeled from pore water sulfate profiles indeep sea sediments (see Section 8.6). It is strikingthat shelf and slope sediments generally fall alongor slightly below a line of equimolar carbonmineralization by oxygen and by sulfate, thusconfirming that sulfate reduction in ocean marginsediments accounts for 25-50% of the entireorganic carbon mineralization. Mangrove sedi-ments provide an exception as they have predo-minant aerobic mineralization, perhaps due to anefficient oxygen transport by the mangrove aerialroots down into the root zone where mostdegradation of organic material takes place. Below2000 m depth in the ocean sulfate reductionclearly looses importance relative to oxygen respi-ration and to suboxic processes such as nitrate,manganese or iron reduction.

This trend is illustrated in Fig. 8.4 where the 15sediment types presented in Fig. 8.3 have beengrouped into five depth zones of the global ocean.The graph shows how strongly the mineralizationof organic matter in the sea bed is shifted towardsocean margin and shelf sediments. The continen-tal shelf out to 200 m water depth comprises only8% of the global ocean area of 3.6 · 108 km2. Yet,61% of the entire benthic oxygen uptake and 68%of the sulfate reduction take place here. If weinclude also the continental slope down to 1000 mwater depth, this upper ocean margin includes

Fig. 8.3 Sulfate reduction versus oxygen uptake rates in marine sediments grouped according to ocean margin typeand water depth: (1) Salt marsh; (2) mangrove; (3) shallow, high deposition; (4) seagras beds; (5) intertidal; (6)estuaries and embayments; (7) upwelling; (8) shelf - depositional; (9) shelf - non depositional; (10) upper slope(200-1000 m); (11) lower slope (1000-2000 m); (12) rise (2000-3000 m); (13) abyss (3000-4000 m); (14) abyss(4000-5000 m); (15) abyss (>5000 m). The line indicates equimolar mineralization rates of organic carbon byoxygen consumption and by sulfate reduction. The double-logarithmic plot is based on data compiled by Canfield(1993) and Canfield et al. (2005).

Page 8: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

277

even 70% of the oxygen uptake and 96% of thesulfate reduction. In other words, according to thedata compiled by Canfield et al. (2005) only 30% ofthe global oxygen uptake of the sea floor and 4%of the sulfate reduction take place at depths below1000 m, covering 87% of the ocean. Although alarger fraction of oxygen respiration than sulfatereduction occurs in deep sea sediments below1000 m, still ca. 40% of the aerobic mineralizationbelow 1000 m ocean depth occurs along thecontinental margins (Jahnke 1996).

The accuracy of such global estimates isclearly limited by the available data from whichthey were calculated. Other estimates of the totalglobal mineralization rates in the sea bed vary byat least two-fold. Thus, the estimate of 23 · 1013

mol yr-1 by Canfield et al. (2005) is slightly higherthan the 19 · 1013 mol yr-1 estimated by Jørgensen(1983) and close to the 22 · 1013 mol yr-1 estimatedby Smith and Hollibaugh (1993). Middelburg et al.(1997) calculated a range of 15-26 · 1013 mol yr-1,depending on whether the geometric or thearithmetric mean value of each depth interval wasused. The low contribution of deep sea sedimentsto the global benthic oxygen respiration is consis-tent with the earlier budget calculations ofJørgensen (1983). Middelburg et al. (1997)estimated that 28% of the global sea floormineralization and 7% of the sulfate reduction takeplace below 2000 m water depth. It should benoted that the low contribution of sulfate reduc-tion in deep sea sediments may be affected by themethod of sulfate reduction rate determination. In

sediments of the shelf and upper slope this isprimarily by radiotracer experiments whereas inthe deep sea rates were modeled. The modelingapproach provides net rates rather than grossrates as obtained by the radiotracer method andthereby tends to underestimate the total rates, inparticular in the near-surface sediment layer (seeSection 8.6).

Since methane formation is shifted at least asstrongly as sulfate reduction towards sedimentswith high organic deposition, an equally smallfraction of the global biogenic methane produc-tion presumably takes place in the deep sea below1000 m.

The global mineralization of sediment organicmatter by oxygen was estimated in Fig. 8.4 to be23 · 1013 mol O2 yr-1 while the global sulfate reduc-tion was 7.5 · 1013 mol SO4

2- yr-1. According tosimple stoichiometries of oxygen respiration andsulfate reduction, one mol of oxygen oxidizes onemol of organic carbon to CO2 whereas one mol ofsulfate oxidizes two mol of organic carbon(Chapter 5):

[CH2O] + O2 → CO2 + H2O (8.3)

2 [CH2O] + SO42- + 2 H+ → 2 CO2 + H2S + 2 H2O

(8.4)

This means that the global estimate of sulfatereduction should be multiplied by two in order toconvert it to carbon equivalents. Thus, sulfate

Fig. 8.4 Relative distributions of total ocean area and of global ocean oxygen uptake or sulfate reduction, groupedinto five depth zones of the ocean. Based on data compiled by Canfield (1993) and Canfield et al. (2005).

8.2 Sulfate Reduction and the Degradation of Organic Matter

Page 9: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

278

oxidizes 15 · 1013 mol organic carbon per year. Thisis equivalent to 65% of the global sediment oxy-gen uptake. The major part of the sulfide produ-ced from sulfate reduction, around 90% in manyshelf sediments, does not become buried in thesediment but is directly or indirectly reoxidizedback to sulfate (see Section 8.5). The complexprocess of sulfide oxidation will thereby ultimatelyconsume oxygen corresponding to the netequation:

H2S + 2 O2 → SO42- + 2 H+ (8.5)

It is a striking consequence of these simplecalculations that about half of the entire oxygenuptake in shelf sediments, where most sulfatereduction takes place, is used for the reoxidationof sulfide or of reduced metals originally reducedby reaction with sulfide (cf. Jørgensen 1977).Thus, if 90 % of 15 · 1013 mol O2 yr-1 is consumed inthis process, it means that 13 out of 23 mol O2 yr-1

or 56% of the oxygen is somehow involved in thereoxidation of sulfide (see Section 8.5). Only theremaining half of the oxygen is therefore available

for all the animals and heterotrophic microorga-nisms that oxidize organic matter directly throughtheir aerobic respiration.

8.3 Anaerobic Oxidation ofMethane (AOM)

Below the sulfate zone, methanogenesis is themain terminal pathway of organic carbon minera-lization. Methane is produced exclusively byanaerobic archaea that utilize a narrow spectrumof substrates for the process (Whitman et al.1999). Methanogenic bacteria or eukaryotes arenot known to exist. The primary sources ofmethane formation in marine sediments are fromthe splitting of acetate (CH3COO-) and from thereduction of CO2 by hydrogen (Eq. 8.6 and 8.7):

CH3COO- + H+ + → CH4 + CO2 (8.6)

CO2 + 4 H2 → CH4 + 2 H2O (8.7)

Fig. 8.5 Profiles of pore-water sulfate and methane concentrations and of rates of sulfate reduction and methaneoxidation for a sediment core recovered from the Kattegat (Station B; 65 m water depth). The broken horizontalline denotes the depth where sulfate and methane were at equimolar concentrations - indicating the peak of thesulfate/methane transition. From Iversen and Jørgensen (1985).

Page 10: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

279

8.3 Aaerobic Oxidation of Methane (AOM)

As minor sources, also methanol, formalde-hyde, and some methylated compounds such asmethyl amines and methyl mercaptans may beused for methanogenesis. The energy yields bythese terminal steps in organic degradation arelower than by anaerobic respiration, e.g. withnitrate, metal oxides or sulfate. Sulfate reducingbacteria therefore compete effectively with themethanogenic archaea for substrates which bothgroups of organisms can utilize, primarily acetateand H2. As a result, methane is not produced asthe main end product in marine sediments untilsulfate has been largely depleted by sulfatereducing bacteria and electron acceptors for arespiratory oxidation of the organic matter aretherefore no longer available (Martens and Berner1974). Only small amounts of methane may still begenerated in the sulfate zone from non-competi-tive methylated substrates. This is a major diffe-rence from the degradation of organic matter infreshwater sediments where sulfate is highlylimited and methane is the main end product ofanaerobic mineralization.

Methane is a chemically unreactive moleculethat requires microbiological catalysis for theactivation and oxidation to CO2 at environmentaltemperatures. Aerobic bacteria that oxidize methanewith O2 are well known from diverse environments,including sea water and marine sediments, andhave been studied in pure culture in the laboratory(Bowman 2000). These earlier physiological andbiochemical studies indicated that biologicalmethane oxidation requires molecular oxygen.Anaerobic oxidation of methane (AOM) wasrecognized by geochemists already in the 1970’ies,however, as a key process in marine sediments(Reeburgh 1969, 1976; Martens and Berner 1974;Barnes and Goldberg 1976). A large number ofstudies based on reaction-transport modeling,radiotracer experiments, inhibition techniques, andstable isotope data have since then firmlyestablished that methane is oxidized biologically inthe absence of O2 at the transition between sulfateand methane. The microorganisms responsible forthe process remained elusive, however, and eventoday it has not been possible to bring anaerobicmethane oxidizers into pure culture for detailedlaboratory study. In spite of this, the identity of theorganisms has now been discovered and researchon the biogeochemistry of anaerobic oxidation ofmethane has made significant progress in recentyears (see reviews by Valentine and Reeburgh2000; Hinrichs and Boetius 2002).

8.3.1 The AOM Zone in Marine Sediments

Most of the methane produced in the sulfatedepleted sub-surface sediment diffuses upwardsalong a steep methane gradient until it reaches thelower sulfate zone and becomes oxidized. Withinthis reaction zone, referred to as the “sulfate-methane transition” (SMT), pore-water methaneand sulfate are both consumed to depletion. Thisdistinct zone is typically located one to severalmeters below the sediment surface in continentalmargin sediments and plays a key role in thebiogeochemistry of the sea bed. In this zone mostof the energy and reducing power of organiccarbon mineralized to CO2 and CH4 below thesulfate zone comes into contact with an efficientelectron acceptor, sulfate, and is finally oxidized toCO2. Since this methane flux comprises most ofthe subsurface methanogenesis, the anaerobicoxidation of methane integrates the degradation ofburied organic carbon from the lower boundary ofthe sulfate zone to very deep sediment layerswhich were deposited many thousands to millionsof years ago. The coupled sulfate-methane reac-tion has been proposed to proceed according tothe following net equation, assuming a one to onestoichiometry between methane and sulfate (e.g.Murray et al. 1978; Devol and Ahmed 1981):

CH4 + SO42- → HCO3

- + HS- + H2O (8.8)

Figure 8.5 shows an example of the sulfate-methane transition zone in a continental shelfsediment from Kattegat (Denmark) at the transi-tion between the Baltic Sea and the North Sea.The sediment was characterized by a high sedi-mentation rate (0.16 cm yr-1) and consisted of fine-grained silt and clay with an average organicmatter content of 12% (Iversen and Jørgensen1985). Sulfate diffused downwards to becomedepleted by sulfate reduction at 160 cm sedimentdepth while methane diffused upwards andpenetrated ca. 40 cm into the lower sulfate zone toca. 120 cm depth. In order to determine anaerobicmethane oxidation rates in the sulfate-methanetransition zone, Iversen and Jørgensen (1985)carried out radiotracer measurements of sulfatereduction using 35SO4

2- and of methane oxidationusing 14CH4. Fig. 8.5 (right frame) shows that therates of sulfate reduction were high, starting frombelow the 10-15 cm deep suboxic sediment, andgradually dropped with increasing depth and agein the sediment. In the SMT, the rates increased

Page 11: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

280

again with a peak just where sulfate and methanereached the same molar concentrations (smallinsert in Fig. 8.5). At that depth also the anaerobicoxidation of methane peaked with similar rates asthe sulfate reduction. The integrated rates ofmethane oxidation in the transition zone accoun-ted for 89 % of sulfate reduction at this depth.This example illustrates how bacterial sulfatereduction changes from organiclastic (based onthe degradation of organic material) to methano-trophic (based on the oxidation of methane) downthrough the sediment column.

The pore water gradients of sulfate andmethane and the depth of the AOM zone may varystrongly among sediments, depending on waterdepth, organic flux and other factors. Fig. 8.6shows an example from the continental slope ofthe Benguela upwelling area in the southernAtlantic. The sulfate profile here shows littledepletion in the top 0-2 m below which sulfatedropped almost linearly down to the sulfate-methane transition located at 5-6 m depth belowsea floor. The concentration of H2S peaked right atthe SMT from where the H2S concentrationdecreased both upwards towards the sedimentsurface and downwards (discussed below inSection 8.5). Based on pore-water concentrationprofiles, Niewöhner et al. (1998) calculated thediffusive flux of sulfate and methane into thetransition zone in the sediment. Their calculationsshowed that anaerobic methane oxidationaccounted for 100% of the deep sulfate reduction

within the sulfate-methane transition zone, i.e. itcould consume the total diffusive sulfate flux.These findings demonstrate that methane can bethe primary electron donor for sulfate reduction inthe sulfate-methane transition zone.

The methane profile in Fig. 8.6 shows thatbelow the steep gradient at the top of the methanezone the methane concentrations gradually dropagain. This is an artifact due to degassing ofmethane immediately upon core retrieval. Due tothe relatively low solubility of methane and thelarge pressure difference between in situ depthand sea surface, methane is highly supersaturatedin the pore water upon recovery of sediment coresand escapes as bubbles before sampling on boardship. For this reason, true methane concentrationgradients above atmospheric pressure are difficultto determine and only recently have pressurizedcore barrels been introduced (Dickens et al. 2003).

8.3.2 Energy Constraints and Pathway ofAOM

Microorganisms able to perform anaerobicoxidation of methane were only recently disco-vered through intensive studies, not of the“normal” subsurface SMT, but of unique marineenvironments in which AOM dominates thecarbon and sulfur cycles, such as sedimentsassociated with gas hydrates and methane seeps.(Gas hydrates are ice-like solids - generallycomposed of water and methane - which occur

Fig. 8.6 Pore-water concentration profiles from gravity core GeoB 3714-9 from the Benguela upwelling area (2060m water depth), South Atlantic. The shaded bar marks the sulfate-methane transition zone. The methane samplelabeled C.C. was taken from the core catcher immediately after core recovery. From Niewöhner et al. (1998).

Page 12: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

281

naturally in sediments under conditions of highpressure, low temperature and high methaneconcentrations; see also Chapter 14). In suchsediments from the Hydrate Ridge at 700 m waterdepth off the Oregon coast, conspicuous micro-bial aggregates, apparently responsible for theanaerobic oxidation of methane, were firstobserved (Boetius et al. 2000). The 3-5 µm largeaggregates were stained with fluorescent mole-cular probes and studied under the fluorescencemicroscope. They were all found to consist of acentral colony of archaea overgrown by sulfatereducing bacteria. In the AOM zone of thesediment, the biomass of these aggregates excee-ded the biomass of all other microorganisms by anorder of magnitude, thus indicating that they wereindeed responsible for the methane oxidation withsulfate. Since then, similar aggregates have beenfound world-wide in diverse methane enrichedsurface sediments (Orphan et al. 2001, 2002;Michaelis et al. 2002; Knittel et al. 2005).

Further evidence for the identity of theanaerobic methane oxidizers had come fromisotope analyses of specific lipid biomarkers insediments with high rates of AOM. Relative tomarine organic carbon or bicarbonate, methane ishighly enriched in the lighter carbon isotope, 12C,over the heavier carbon isotope, 13C, with δ13Cvalues ranging from -50 to -90‰ (Whiticar 1999;see also Chapter 10). The archaeal biomarkers inthese sediments were strongly enriched in 12C,which indicated that methane served as the maincarbon source for the microorganisms (Hinrichs etal. 1999; Elvert et al. 2000; Thiel et al. 2001). Alsobiomarkers derived from bacteria showed extreme

12C enrichment and indicated that the associatedsulfate reducers were also involved in the AOMprocess (Hinrichs et al. 2000; Elvert et al. 2000;Pancost et al. 2000). Final evidence that theaggregates carried out anaerobic oxidation ofmethane was provided by a recently developedSIMS technique (secondary ion mass spectro-metry) by which carbon isotopic analysis could bedone on individual microscopic AOM aggregatesand thereby confirm that their cell materialcontained the 12C depleted isotope signal ofmethane (Orphan et al. 2001).

Equation 8.8 above proposed a net equation foranaerobic oxidation of methane by sulfate that isbased on modeled pore water profiles andlaboratory experiments with sediment enrichments.No single microorganism is yet known, however,that is able to carry out such a complete oxidationof methane to bicarbonate and it remains an openquestion whether such organisms exist. Instead,the process may proceed in two steps involving anintermediate electron carrier that is transferred fromthe archaea to the sulfate reducers within the AOMaggregates. Hoehler et al. (1994) first suggestedthat AOM is carried out by a consortium of archaeaand sulfate reducing bacteria and that hydrogenmight serve as such an intermediate that istransferred from the methane oxidizers to thesulfate reducers in a two-step reaction (Fig. 8.7):

I. CH4 + 2 H2O → CO2 + 4 H2 (8.9)

II. SO42- + 4 H2 + H+ → HS- + 4 H2O (8.10)

Sum of Eq. 8.9 and 8.10 = Eq. 8.8

Fig. 8.7 Schematic of AOM aggregate with a syntrophic metabolism among two members that only in combinationcan catalyze the complete oxidization of methane with sulfate. A transfer of H2 or acetate (CH3COO-) frommethanotrophic archaea to sulfate reducing bacteria has been proposed as an intermediate in the net process. Suchan interspecies transfer of hydrogen or organic carbon is still hypothetical and has not been directly demonstrated.

8.3 Aaerobic Oxidation of Methane (AOM)

Page 13: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

282

The first step in this reaction (Eq. 8.9) isobviously a reversal of methanogenesis from CO2and H2 shown in Eq. 8.7. How can a metabolicprocess run in two opposite directions and yetyield energy for both types of archaea? Hoehleret al. (1994, 1998) showed that the energy yield ishighly dependent on the H2 concentration which,in turn, is regulated by the predominant processof mineralization. In the deep methanogenic zonethe H2 concentration is relatively high, about 10nM, so that methanogenesis (Eq. 8.7) is an exer-gonic, energy yielding process. In the sulfatezone, the sulfate reducing bacteria maintain such alow H2 concentration, about, 1 nM, that it fallsbelow the concentration required for an exergonicmethanogenesis. On the contrary, the reversal ofmethanogenesis (Eq. 8.9) becomes exergonic andAOM may proceed. This strong dependence onH2 concentration could explain why the methaneoxidizing archaea grow in aggregates togetherwith the sulfate reducing bacteria. Only when thesulfate reducers efficiently scavenge H2 as it isproduced by the archaea are these able to conti-nue oxidizing methane.

Although this idea of a syntrophic associationbased on inter-species hydrogen transfer is veryappealing, it explains only a part of the obser-vations on AOM. Since both the methanotrophicarchaea and the sulfate reducing bacteria carry thelight carbon isotopic signal of methane, it is aquestion how also carbon is transferred in thissyntrophic association. An alternative pathwaycould therefore be the conversion of methane toacetate and the subsequent oxidation of acetate(CH3COO-) by the sulfate reducers with a con-current incorporation of part of the acetate intotheir cell biomass:

CH4 + HCO3- → CH3COO- + H2O (8.11)

SO42- + CH3COO- → HS- + 2 HCO3

- (8.12)

Further possibilities exist, such as a combinedtransfer of acetate and hydrogen or a transfer offormate, both of which may explain the δ13C obser-vations and be more compatible with the energeticconstraints on AOM (Valentine and Reeburgh2000; Sørensen et al. 2001). Further research isrequired to understand the basic mechanism andregulation of this key process in the methanecycle.

Although AOM provides an almost completebarrier to methane escape from the sea floor in

sediments with diffusive methane flux, the processis surprisingly sluggish and takes place over asediment horizon and a time scale that highlyexceed that of most other microbial processes ingradient environments. Thus, the life-time ofmethane upon diffusion up into the sulfate zone isgenerally on the order of months to years(Jørgensen et al. 2004). This is a reason why weprefer the term “sulfate-methane transition” ratherthan “sulfate-methane interface”. A reason for theslow turnover may be that the methane oxidizingcommunity is operating near the theoretical mini-mum in energy yield required for microbial growth.The estimated energy yield of AOM in a numberof sedimentary environments is -20 to -25 kJ permol of methane consumed (Hoehler et al. 1994)which is less than the energy required for theformation of ATP (see Chapter 5). If the processdoes indeed involve a two-step reaction such assuggested in Eq. 8.9-8.10 or 8.11-8.12, then thisenergy even has to be shared among two partners.This may explain why the growth of AOM aggre-gates is exceedingly slow and requires manymonths for a doubling of the biomass in labora-tory incubations (Nauhaus et al. 2002). It may alsoexplain why anaerobic methane oxidizing micro-organisms have not yet been isolated in pureculture in the laboratory. They simply grow soslow that isolation may take many years. Theseobservations demonstrate, however, that proces-ses catalyzed by microbial communities in theenvironment can be highly efficient in energyconservation and operate close to the theoreticalthermodynamic limits (Schink 1997; Hoehler et al.2001; Jackson and McInerney 2002).

By the use of molecular methods based onsequence information from 16S rRNA genes, abroad diversity of archaea that all belong to theEuryarchaeota (Orphan et al. 2002; Knittel et al.2005) and sulfate reducing bacteria mostlybelonging to the Desulfosarcina-Desulfococcusbranch of the Delta-proteobacteria (Boetius et al.2000; Orphan et al. 2001) has been consistentlyfound to be associated with anaerobic oxidationof methane. Although the genetic identity of AOMmicroorganisms is now revealed, the physiologyand biochemistry of AOM are still incompletelyunderstood. A recent clue came from the discove-ry that the terminal key enzyme of methano-genesis, methyl coenzyme M reductase, and itsencoding gene seem also to be involved in theprocess of AOM in a slightly modified form, thusindicating that AOM may partly be a reversal of

Page 14: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

283

methanogenesis in certain archaea (Krüger et al.2003; Hallam et al. 2003). In this respect,microbiological and biogeochemical research iscurrently providing mutually supporting evidencefor the nature of AOM.

8.3.3 Quantitative Role of AOM

Since methane formation is largely restricted tosediment layers below the depth of sulfate pene-tration, there is relatively little organic carbon withlow reactivity left to fuel methanogenesis. Whenintegrated through the sediment column, theamount of organic carbon mineralized with theformation of methane is generally only 5-20% ofthat mineralized by sulfate reduction (Canfield1993; Canfield et al. 2005; Table 8.2), with 10%possibly representing a mean value. Since sulfatereduction typically accounts for half of theorganic mineralization in ocean margin sediments,it should only be about 5% of the total mineralizedorganic carbon that is ultimately degraded tomethane. In deep sea sediments the fraction ispresumably lower but relevant data to confirm thisare lacking. In some coastal environments withvery high organic loading, such as Cape LookoutBight on the Atlantic coast of North America,much more of the organic carbon may be buriedbelow the sulfate zone and be degraded tomethane (Crill and Martens 1986). Also in ocean

margin sediments where the methane flux isenhanced due to subduction or is focused inassociation with methane seeps or surficial gashydrate accumulations, methane may provide themain carbon source for the entire sulfate reduction(e.g. Boetius and Suess 2004). Table 8.2 providesselected examples of the contribution of methaneas a carbon and energy source for sulfatereduction in ocean margin sediments.

As discussed in Section 8.6, the sulfate reduc-tion rates determined from modeling of sulfateprofiles may underestimate the rates in the mostactive layers of near-surface sediment, and modelcomparisons of sulfate and methane fluxestherefore tend to indicate a greater role of methanefor the entire sulfate reduction than data based onexperimental rate measurements. From interstitialflux calculations, Reeburgh (1976, 1982) thusestimated that approximately 50 % of the netdownward sulfate flux at a Cariaco Trench station- an anoxic basin - could be accounted for bymethane oxidation. For Saanich Inlet sediments,75 % of the downward sulfate flux was attributedto anaerobic methane oxidation according tosimple box model calculations (Murray et al. 1978).Devol et al. (1984) obtained lower percentages of23 to 40 % of the downward sulfate flux consumedby methane oxidation for these same sedimentsusing a coupled reaction diffusion model. Iversenand Jørgensen (1985) reported that in Kattegat

Table 8.2 Role of methane as a carbon source for sulfate reduction in marine sediments. The compiled data showcumulative sulfate reduction rates measured by radiotracer technique, either over the entire sulfate zone, or in theupper 0-15 cm combined with modeling below that depth. The contribution of methane was calculated from thediffusion flux of methane up into the lower sulfate zone. In other data sets where sulfate reduction rates aredetermined only by modeling, or where also methane oxidation was measured by radiotracer technique, the calculated% of SRR from CH4 is higher than shown here. (SRR = sulfate reduction rate).

8.3 Aaerobic Oxidation of Methane (AOM)

Water depth SRR CH4 flux,[m] [mmol m-2 d-1] % of SRR

Skan Bay, Alaska 65 19 8% Alperin and Reeburgh (1988)Thode-Andersen and Jørgensen (1989)

Knab and Fossing (unpubl.)Kattegat-Skagerrak 65-200 4.8-5.6 4-6% Iversen and Jørgensen (1985)

Saanich Inlet 225 17 9% Devol et al. (1984)Namibian slope 1372-2060 1.25-2.22 10-12% Fossing et al. (2000)Chilean slope 799-2744 0.53-0.82 8-24% Treude et al. (2005)

Crill and Martens (1983, 1986)Chanton (1985)

Black Sea, anoxic zone 130-1176 0.65-1.43 7-18% Jørgensen et al. (2001)

Cape Lookout Bight 9 24-27 40%

Location References

Århus Bay, Baltic Sea 18 2.90 3%

Page 15: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

284

and Skagerrak sediments methane oxidationaccounted for only 10 % of the electron donorrequirement for the depth integrated sulfate reduc-tion. In this case sulfate reduction rates were mea-sured by radiotracer technique throughout theentire sulfate zone.

The proportion of sulfate used for organicmatter degradation in comparison to sulfate usedfor anaerobic methane oxidation may be calculatedunder molecular diffusion conditions from theshape of sulfate profiles. Borowski et al. (1996)inferred that the linear sulfate pore water concen-tration profiles found in Carolina Rise and BlakeRidge sediments imply that anaerobic methaneoxidation is the dominant sulfate consumingprocess. They calculated the upward methane fluxfrom measured sulfate pore water profilesassuming that the downward sulfate flux isstoichiometrically balanced by upward methaneflux. In this way they used the sulfate concen-tration profiles to calculate the methane flux fromthe underlying methane gas hydrates which hadbeen documented by seismic profiling. Niewöhneret al. (1998) found similar linear sulfate concen-tration profiles in sediments of the Benguelaupwelling area (see Fig. 8.6) and concluded thatanaerobic methane oxidation could account for theentire deep sulfate flux.

These studies demonstrate that the sulfate-consuming process occurring in the deepersediments, i.e. AOM, is the dominant factor deter-mining the shape of the sulfate pore waterprofiles, although oxidation of organic matterdemonstrably occurs throughout the entire sulfatezone. Linear sulfate profiles can, therefore, beused to calculate the upward methane flux (e.g.Devol and Ahmed 1981; Borowski et al. 1996;Niewöhner et al. 1998) but they do not provideaccurate sulfate reduction rates occurring in near-surface sediments. Concave-down pore waterprofiles may develop when methane dependentsulfate reduction in deeper sediment layers is lessimportant as is schematically illustrated in Figure8.8 or if transient conditions prevail in the porewater system (Hensen et al. 2003; Kasten et al.2003).

Reeburgh et al. (1993) estimated the globalmethane flux in marine sediments to a mean valueof 70 Tg yr-1 or 0.5 · 1013 mol CH4 yr-1. This isequivalent to 2% of the entire organic carbonmineralization via oxygen respiration and about7% of the global sulfate reduction. A more recentestimate by Hinrichs and Boetius (2002) of AOM

in marine sediments yielded a four-fold highernumber, 300 Tg yr-1 or about 2 · 1013 mol CH4 yr-1.These calculations do not include hot-spots ofmethane fluxes from cold seeps, hot vents orsurficial gas hydrates, since reliable estimates ofthe total areal distribution of such sites are notyet available. Methane seeps occur both at activeand passive margins but are highly focused andoften variable over time. Judd et al. (2002) made anestimate of the global flux of methane from the seabed of 16-40 Tg yr-1, i.e. much less than theestimated subsurface flux. Data compiled byHinrichs and Boetius (2002) indicate that, even ifthe area affected by methane seepage atcontinental margins is below 1%, this might have

Fig. 8.8 Schematic profiles of sulfate and methane in marinesediments. A) In sediments with low methane flux, sulfate re-duction based on oxidation of sediment organic matter predo-minates throughout the sulfate zone. B) In sediments with highmethane flux, sulfate reduction based on anaerobic oxidation ofmethane tends to straighten out the sulfate profile.

Page 16: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

285

8.4 Effects of Sulfate Reduction on Sedimentary Solid Phases

a significant impact on the total methane budget.Therefore, a more extensive mapping of seepageareas, as well as a broader data base on diffusivemethane fluxes, is needed for more accuratecalculations of the methane cycling in the globalsea bed (see also Chapter 14).

8.4 Effects of Sulfate Reductionon Sedimentary Solid Phases

Sulfate reduction, occurring either due to theoxidation of methane or the mineralization oforganic material, can lead to a pronouncedoverprint or modification of the primary sedimentcomposition by dissolution/reduction of mineralsand precipitation of authigenic mineral phases.Since most of the minerals affected are alsocommonly used for paleoceanographic and paleo-climatologic reconstructions it is crucial to consi-der and assess the extent of diagenetic alterationof the sedimentary record driven by sulfatereduction.

8.4.1 Reactions with iron

Iron sulfides represent the most importantminerals that form in association with bothorganoclastic and methanotrophic sulfate reduc-tion, or - more precisely - as a result of the hydro-gen sulfide produced by these processes. Thedifferent pathways of pyrite formation viaintermediate iron sulfides will be described in moredetail in Section 8.4.2. The first step in all pyriteforming sequences involves a reaction ofhydrogen sulfide with either dissolved Fe2+ orsolid-state iron (oxyhydr)oxides. The reactivity ofoxidized iron minerals towards sulfide varies

significantly as shown in Table 8.3. In their recentstudy, Poulton et al. (2004) demonstrated thatminerals with a lower degree of crystal order reactwithin minutes to hours, while more orderedminerals react on time scales of days to years (c.f.Chapter 7). The largest discrepancy in reactivitybetween the studies of Canfield et al. (1992) andRaiswell et al. (1994) on the one hand and Poultonet al. (2004) on the other hand exists for themineral magnetite (105 years versus 72 days;Table 8.3). This difference was explained by thesurface area of magnetite, which was taken intoconsideration in the study of Poulton et al. (2004).

Within or around the zone of anaerobicoxidation of methane, other important mineralprecipitates comprise authigenic carbonates –such as Mg-calcite, aragonite and dolomite –which precipitate due to the high concentrationsof HCO3

- ions generated by AOM (c.f., Eq. 8.8;e.g., Greinert et al. 2002; Moore et al. 2004), as wellas diagenetic barite (e.g., Brumsack 1986; Torres etal. 1996). Due to the high rates of AOM, theformation of carbonate precipitates is particularlypronounced in sedimentary settings influenced byventing and seepage of methane-rich fluids and/orthe presence of gas hydrates. The formation ofcarbonates in such methane dominated environ-ments, which are mostly driven by advectivetransport processes, is discussed in Chapters 13and 14. The factors and conditions determiningthe dissolution and preservation of the differentcarbonate phases are the subject of Chapter 9.

8.4.2 Pyrite Formation

Iron-sulfide minerals are important sinks for ironand sulfur as well as for trace metals and play animportant role in the global cycles of these ele-ments. Over the past 30 years extensive studies –

Table 8.3 Reactivity of iron (oxyhydr)oxides towards sulfide. Half-lives (t1/2) are given for reductive dissolution inseawater at pH 7.5 and at a sulfide concentration of 1000 µM. adata from Poulton et al. (2004); bdata from Canfieldet al. (1992) and Raiswell et al. (1994). (Adopted from Nüster, 2005).

M in e ra l t1 /2a t1 / 2

b

F res h ly prec ip ita ted hy dous fe rric ox ide 5 m in .

2-line fe rrihy drite 12 .3 hours 2 .8 hours

Lep idoc roc ite 10 .9 hours < 3 day s

G oeth ite 63 day s 11.5 day s

M agnet ite 72 day s 105 y ears

H em at ite 182 day s < 31 day s

Page 17: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

286

both in the field and in the laboratory – have beenperformed to elucidate the mechanisms of pyriteformation and in particular to understand the mainfactors which control the formation of pyrite innatural environments. Although these studieshave provided significant new insights, funda-mental questions remain with respect to (a) thedifferent reaction rates obtained on the basis offield studies and laboratory experiments, (b) therole of FeS surface reactions and the electronacceptor involved in the conversion to pyrite, and(c) the role of sulfate-reducing bacteria – in par-ticular their cell walls - in directing and promotingthe pyrite precipitation process. A comprehensivereview of pyrite formation in sedimentary environ-ments by Rickard et al. (1995) was recentlyupdated by Schoonen (2004). Besides the detaileddescription of the different pathways and mecha-nisms of sedimentary pyrite formation, the reviewby Schoonen (2004) also discusses the synthesisof nanoscale pyrite, trace element incorporation,and the formation of electronic defects during theformation process.

All pyrite forming pathways identified so farinvolve several reaction steps. First, hydrogensulfide, produced during sulfate reduction (Eq.8.1, 8.2 and 8.8), reacts with dissolved iron orreactive (towards sulfide) iron minerals to formamorphous iron sulfides, such as mackinawite(FeS). The precipitated amorphous iron sulfide ishighly unstable and rapidly transforms to metastableiron sulfide phases such as pyrrhotite (Fex-1S) orgreigite (Fe3S4), both of which represent interme-diates in the reaction pathways to pyrite (FeS2).The conversion of amorphous FeS to pyriterequires an electron acceptor and a change in themolar Fe:S ratio from about 1:1 to 1:2. The electronacceptor is needed to oxidize the S(-II) componentin FeS to the mean oxidation state of –I in FeS2.Concurrently, the Fe:S ratio has to decrease eithervia the addition of sulfur or the loss of iron. Three

general pathways have been reported to convertFeS to pyrite:

(1) Addition of sulfur with the sulfur speciesacting as electron acceptor (Berner 1970; Berner1984; Luther 1991). This pathway has been termedthe “polysulfide pathway”.

FeS + S0 → FeS2 (8.13)

(2) Reaction with hydrogen sulfide, i.e.addition of sulfur with a non-sulfur electron ac-ceptor (Rickard and Luther 1997). This con-version mechanism is known as the “H2S path-way”.

FeS + H2S → FeS2 + H2 (8.14)

(3) Loss of iron, combined with an (additional)electron acceptor (Wilkin and Barnes 1996), knownas the “iron-loss pathway”.

2FeS + 2H+ → FeS2 + Fe2+ + H2 (8.15)

As reviewed in detail by Schoonen (2004)these different conversion mechanisms – and inparticular the H2S pathway – have receivedcontroversial discussion. However, field studieshave shown that hydrogen sulfide can indeedsulfidize amorphous FeS and form pyrite. Rickard(1997) found that the H2S process is by far themost rapid of the pyrite-forming reactions hithertoidentified and suggested that it represents thedominant pyrite forming pathway in strictly anoxicsystems. In addition, Morse (2002) discussed thatthe oxidation of FeS by hydrogen sulfide is thefaster process compared with the oxidation byelemental sulfur. Berner (1970) suggested that, inthe presence of zero-valent sulfur, a completetransformation of FeS to pyrite should be possibleon a time scale of years. An incomplete conver-sion of FeS to pyrite, as often observed, e.g. in

Fig. 8.9 Schematic representation of the major pathways of the transformation of iron oxides to iron sulfides inanoxic marine environments, in relation to the alteration of the magnetic record. From Riedinger (2005).

Page 18: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

287

the anoxic sediments of Kau Bay, Indonesia,(Middelburg 1990) or the Amazon Fan (Kasten etal. 1998), can be due to a shortage of zero-valentsulfur during pyrite formation. Zero-valent sulfurmay form as a result of the incomplete oxidation ofH2S or FeS by oxidants such as O2, NO3

-, MnO2 orFeOOH. A limited transformation of FeS to pyritecan therefore be due to either (a) a limited amountof sulfide available to be oxidized to zero-valentsulfur or (b) a shortage in the amount of oxidantssupplied by diffusion, bioturbation or burial.

Direct precipitation of pyrite without intermedi-ate iron sulfide precursors was reported for saltmarsh sediments, where pore waters were under-saturated with respect to amorphous FeS butoversaturated with respect to pyrite (Howarth1979; Giblin and Howarth 1984). In these sedi-ments, the oxidizing activity of the roots favoredthe formation of elemental sulfur and polysulfideswhich were thought to react directly with Fe2+.The direct reaction pathway may proceed withinhours, resulting in the formation of single small,euhedral pyrite crystals (Rickard 1975; Luther etal. 1982). Framboidal pyrite, apart from that formedby the mechanism presented by Rickard (1997), isformed slowly (over years) via intermediate ironsulfides (Sweeney and Kaplan 1973; Raiswell1982).

Irrespective of the pyrite forming pathway theconversion of FeS into pyrite via intermediate ironsulfides goes along with a pronounced change inthe rock magnetic properties of the particular ironsulfide phases (Fig. 8.9, Table 8.4). AmorphousFeS is paramagnetic and therefore has a lowmagnetic susceptibility and does not contribute tothe remanent magnetization of sediments. In

contrast, the metastable iron sulfides, pyrrhotiteand greigite, which represent precursor phases ofpyrite, are ferrimagnetic and thus have a signi-ficant magnetic potential. This becomes obviousfrom maxima in magnetic susceptibility locatedslightly above enrichments of amorphous FeS inthe form of black bands in sediments of theAmazon Fan (see Fig. 8.11 below; Kasten et al.1998) and in the Black Sea (Neretin et al. 2004).The presence of greigite and pyrrhotite – res-ponsible for the peaks in magnetic susceptibility –document the progress in the sequence ofconversion of FeS to pyrite. With a further pro-gression in the conversion pathway and theformation of the stable pyrite the magnetic signalis lost again as pyrite is paramagnetic (Fig. 8.9,Table 8.4).

It is generally assumed and found that there isa gradual decrease in the amount of amorphousiron sulfides and a corresponding increase in theamount of pyrite – thus a drop in the AVS to pyriteratio – upon burial with increasing sedimentdepth. The discussion and examples presented inthe following sections will, however, show thatthis is not always the case. In particular insedimentary sequences affected by sulfidizationfronts the sequence of conversion of FeS intopyrite can be reversed with depth and the mostimmature and unstable Fe sulfides be found atgreater depth coinciding with the current depth ofthe diffusional H2S/Fe2+ boundary. Furthermore,iron sulfide formation often occurs at or alongfixed geochemical boundaries or reaction fronts.Thus, the sequence of iron sulfide formation isoften restricted to or even bound to certain depthhorizons rather than occurring gradually with

Table 8.4 Authigenic iron sulfides formed in marine sediments. Modified from Schinzel (1993).

Mineral Composition Crystal class Magneticproperties

mackinawite FeS tetragonal paramagnetic

pyrrhotite Fe1-xS hexagonal or ferrimagnetic,orthorhombic antiferrimagnetic

pyrite FeS2 cubic paramagnetic

markasite FeS2 orthorhombic diamagnetic

greigite Fe3S4 cubic ferrimagnetic

8.4 Effects of Sulfate Reduction on Sedimentary Solid Phases

Page 19: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

288

increasing sediment depth. Mineral precipitationat sulfidization fronts can also occur cyclicallywithin the sedimentary sequence – superimposedby episodic changes in depositional or otherenvironmental conditions.

8.4.3 Magnetite and Barite

Besides the precipitation of mineral phases, disso-lution of numerous minerals initially supplied tothe seafloor can occur in sediments that expe-rience sulfate reduction and/or sulfate-depletedconditions. Iron (oxyhydr)oxides and barite(BaSO4) are two examples with particular relevancefor paleoceanographic research.

Magnetic iron (oxyhydr)oxides, especiallymagnetite (Fe3O4), are the main carriers ofremanent magnetization in sediments. Magneto-stratigraphy of deep sea sediment cores is avaluable method of dating and comparing sedi-mentary records. Dissolution of the magneticminerals under sulfate- and iron-reducing condi-

tions and/or subsequent precipitation of authi-genic iron minerals may, however, alter the initialremanent magnetization and seriously compromisethe interpretation of the sedimentary geomagneticrecord (e.g. Karlin and Levi 1983, 1985; Funk et al.2003a, 2003b; Reitz et al. 2004). The effects ofAOM-driven sulfate reduction on the transfor-mation of iron minerals and the associated modi-fication of rock magnetic properties will be demon-strated below using examples from the AmazonFan and the western Argentine Basin.

A second sedimentary component importantfor paleoceanographic reconstructions, which isprone to dissolution under conditions of sulfatereduction – or more precisely in sulfate-depletedsediments - is the barium sulfate mineral, barite(BaSO4). Since a correlation has been detectedbetween barite deposition and the flux of organicmatter through the water column, the concen-tration of barite in sediments has been proposedand applied as a geochemical tracer to reconstructpast changes in ocean productivity (e.g., Bishop

Fig. 8.10 Geochemical data for core GeoB 1023-4 recovered off north Angola (17°09.6’S, 10°59.9’E, 2047 m waterdepth). Barium and sulfate pore-water concentration profiles as well as the distribution of solid-phase bariumindicate the precipitation of authigenic barite at a front slightly above the depth of complete sulfate consumption.Below the sulfate/methane transition barite becomes undersaturated and is thus subject to dissolution due to the totaldepletion of pore-water sulfate. Dissolved barium diffuses upwards into the sulfate zone where the mineral baritebecomes supersaturated and so-called authigenic or diagenetic barite precipitates at a front at the base of the sulfatezone. Modified from Gingele et al. (1999), after Kölling (1991).

Page 20: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

289

1988; Dehairs et al. 1991; Dymond et al. 1992;Gingele and Dahmke 1994; Paytan et al. 1996a).Sedimentary barite has also been used to recon-struct the strontium isotope composition of sea-water over time (e.g., Paytan et al. 1993; Mearon etal. 2003), to determine the sulfur isotope ratio ofmarine sulfate (Cecile et al. 1983; Goodfellow andJonasson 1984; Paytan et al. 1998), and tocharacterize Holocene sedimentation rates byusing excess 226Ra decay (Paytan et al. 1996b; vanBeek and Reyss 2001; van Beek et al. 2002).

Although it is frequently stated that theprocess of sulfate reduction promotes barite dis-solution it is, in fact, the complete depletion ofinterstitial sulfate which leads to an under-saturation of the pore water with respect to bariteand to its subsequent dissolution (e.g., vonBreymann et al. 1992; Torres et al. 1996; Gingele etal. 1999). This process is obvious from pore waterBa2+ concentrations which typically increase tomicromolar concentrations with depth below theSMT. Dissolution of barite below the SMT andreprecipitation of diagenetic barite slightly abovethe sulfate penetration depth in so-calledauthigenic or diagenetic barite fronts can dras-tically obscure the primary barite record and thuslead to wrong paleoceanographic interpretations.The effect of barite dissolution in the zone of totalsulfate depletion, and the subsequent reprecipi-tation of diagenetic barite at higher sedimentlevels is illustrated in Figure 8.10 for a sedimentcore recovered from the continental margin offAngola. The use of barite as a geochemical traceror archive for paleoceanographic reconstructionsis therefore limited – not to say precluded - insediment intervals that are either currently sulfate-free or have been strongly sulfate depleted in thepast (e.g. von Breymann et al. 1992; Gingele andDahmke 1994; Torres et al. 1996; Gingele et al.1999; Dickens 2001). While the current geoche-mical zonation of a sediment column can bedetermined from pore water sulfate data, a pastmigration of the SMT – and particularly a down-ward strike over time as a result of a transientdecrease in methane flux from below – has also tobe considered a possible cause that may alter thesolid-phase Ba contents.

In a novel approach, Dickens (2001) usedsedimentary Ba records to assess temporalchanges during the Late Pleistocene in the upwardflux of methane within sediments of the BlakeRidge that are rich in sub-surface gas hydrates.Due to a lack of Ba enrichment above the present

depth of the barite front the author concluded thatthe upward methane flux from the underlying gashydrate reservoir has not varied significantlyacross major changes in sea level and hydrostaticpressure. A possible means to identify the origin/source of barite enrichments is the analysis of thestable sulfur isotopic composition, δ34S, of thebarite particles. As the sulfur isotopic compositionof sulfate in the water column, where the pro-ductivity-related biogenic barite is assumed to beformed, is significantly lighter (around +21 ‰;Paytan et al. 2002) than that of pore water sulfateat the SMT, where diagenetic barite precipitates(around + 40 ‰; Torres et al. 1996), it is generallyassumed that the δ34S of barite should give insightinto its formation mechanism or origin (Paytan etal. 2004).

8.4.4 Non-Steady State Diagenesis

A particularly pronounced alteration of the sedi-mentary solid phase by mineral dissolution andauthigenic mineral precipitation can take place inconnection with sulfate reduction during non-steady-state diagenesis, i.e. during phases ofdeposition by which sedimentary conditions arenot constant over time and sediment geochemistryadjusts to the new situation. Non-steady-statediagenesis can be initiated by any changes in thefluxes of electron donors and acceptors andenvironmental conditions, for example by changesin type of depositing sediment, oxygen content ofthe bottom water, sedimentation rate, flux oforganic matter to the seafloor, and upward flux ofmethane (e.g., Kasten et al. 2003). The non-steady-state processes that occur during thetransition from one depositional situation toanother, or at the interface between two sedimenttypes, typically comprise the development of geo-chemical reaction fronts. These fronts can eitherbe fixed at particular sediment horizons for aprolonged period of time or move downwards orupwards within the sediment column. During afixation or a slow migration of reaction fronts orgeochemical boundaries the diagenetic processesacting at these fronts can produce higher concen-trations of the authigenic minerals formed atspecific sediment levels than under steady-stateconditions. On the other hand, a fast upwardmigration of a geochemical front may support theburial and thus the preservation of substantialamounts of metastable minerals (Riedinger et al.2005, see below).

8.4 Effects of Sulfate Reduction on Sedimentary Solid Phases

Page 21: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

290

In the glacial sediments of the Amazon Fan,which were deposited during the Pleistocene sealevel low-stand, Kasten et al. (1998) reported apronounced enrichment of iron sulfides locatedseveral meters below the seafloor, near thepresent-day SMT (Fig. 8.11). They concluded thatthis iron sulfide enrichment formed during aperiod of non-steady state diagenesis, when thezone of anaerobic methane oxidation became fixedat this particular sediment level for an extendedperiod. The condition that is likely to have initi-ated the non-steady state diagenetic situation wasa pronounced decrease in sedimentation and orga-nic carbon accumulation rate during the transitionfrom the Pleistocene to the Holocene. Similardistinct enrichments of authigenic minerals due toa fixation of the SMT during non-steady state

depositional conditions induced by decreases insedimentation rate or even pauses in sedimen-tation have also been reported by Raiswell (1988),Torres et al. (1996) and Bréhéret and Brumsack(2000).

A second striking feature of the Amazon Fansediments (gravity core GeoB 1514-6) is the occur-rence of a pronounced maximum in magneticsusceptibility, located slightly above the zone ofstrong enrichment of iron sulfide minerals (Fig.8.11). As the total iron sulfides did not display alocal maximum at the depth of the susceptibilitypeak, Kasten et al. (1998) suggested the presenceof a specific iron sulfide with a high magneticpotential. X-ray diffraction analyses performed onsediment samples from this depth revealed thepresence of the magnetic iron sulfide mineral,

Fig. 8.11 Solid phase profiles of total sulfur and Al-normalized concentration of Fe from total digestion (solid lines),magnetic susceptibility (Funk, unpublished data), as well as pore-water concentration profiles of SO4

2- and Fe2+ (solidcircles) for core GeoB 1514-6 recovered from the Amazon deep-sea fan (3509 m water depth). The horizontal linedemonstrates that the iron peak, which represents mainly iron sulfides, does not directly coincide with the peak inmagnetic susceptibility. Modified from Kasten et al. (1998), after Schulz et al. (1994).

Page 22: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

291

greigite (Fe3S4). These findings demonstrate thatnon-steady-state diagenesis does not only lead toa modification of the bulk sediment composition,but can also generate distinct magnetic signals ofpost-depositional origin within the sedimentaryrecord.

The continental margin off Argentina andUruguay is a highly dynamic sedimentary settingcharacterized by gravity driven mass-flowdeposits and is therefore ideally suited to studydiagenetic processes under shifting depositionalconditions. As a typical feature of the deposits inthis area, distinct minima in magnetic susceptibi-lity are found a few meters below the sedimentsurface (c.f. Fig. 8.12 ). In order to reveal theorigin of these susceptibility “gaps”, Riedinger etal. (2005) carried out extensive geochemical androck-magnetic investigations as well as numerical

transport-reaction modeling using the programCoTReM (Chapter 16). Pore water data revealedthat the conspicuous minima in susceptibilitycoincide with the current depth of the SMT (Fig.8.12). The hydrogen sulfide generated by thisprocess reacts with the abundant iron (oxyhydr)-oxides resulting in the precipitation of iron sulfi-des accompanied by a nearly complete loss of themagnetic signal. Below the sulfidic sedimentinterval, where the magnetic susceptibility had notsignificantly suffered from diagenetic overprint,high amounts of iron oxides were present.Numerical modeling of geochemical data suggeststhat these high amounts of preserved Fe(III) aswell as the distinct and spatially restricted loss insusceptibility can only be produced by extremelyhigh glacial sedimentation rates (≥100 cm/kyr) -shielding the Fe(III) minerals from complete

Fig. 8.12 Left frame: Sulfate (red circles), methane (blue circles), and sulfide (stars) pore water profiles for coreGeoB 6229-6 (3446 m water depth) from the western Argentine Basin off the Rio de la Plata (sulfate data are fromHensen et al. 2003). The magnetic susceptibility is shown in grey. Middle and right frame: Results of numericalmodeling of diagenetic alteration of magnetite to iron monosulfide with a major change of mean sedimentation rate(SR) for a sediment porosity of 75%. (a) A mean sedimentation rate of 100 cm kyr-1 leads to reduction of only aboutone third of the magnetite. (b) If the mean sedimentation rate is decreased to 5 cm kyr-1, a time interval of ~9000years is needed to reduce the total amount of magnetite initially contained within an interval of 2 m thickness.Modified from Riedinger et al. (2005).

8.4 Effects of Sulfate Reduction on Sedimentary Solid Phases

Page 23: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

292

reduction - and a subsequent drastic drop duringthe glacial/Holocene transition (Fig. 8.12). Similarto the conditions on the Amazon Fan the strongdecrease in sedimentation rate encountered duringthe last climatic transition induced a fixation of theSMT and an enhanced overprint of rock magneticand mineralogical properties at this particularsediment layer. To obtain the observed geochemi-cal and magnetic patterns, the SMT must haveremained at a fixed position for about 9000 years –a time span which closely corresponds to the timesince the Pleistocene/Holocene transition.

Sulfate reduction can also occur within dis-crete, organic-rich layers which can lead to adistinct overprint of the primary sediment compo-sition within the organic-rich layers or in thesediment above and below. The non-steady statediagenetic processes occurring in and below theorganic-rich layers (sapropels) of the EasternMediterranean have been studied by Passier et al.

(1996). They presented a model for the formationof distinct iron sulfide enrichments below thesapropels (Fig. 8.13) by the development of adownward moving sulfidization front, similar toLiesegang phenomena (formation of distinct ironsulfide bands) described by Berner (1969). ALiesegang situation exists in depositional systemsthat are characterized by intermediate contents ofreactive (towards sulfide) iron, i.e. in systems thatare neither iron nor sulfide dominated. Passier etal. (1996) concluded that excess hydrogen sulfidegenerated by dissimilatory sulfate reductionwithin the sapropel was able to migrate down-wards (downward sulfidization). This resulted inthe formation of pyrite below the sapropel by thereaction of hydrogen sulfide with solid-phaseferric iron and Fe2+ diffusing upwards fromunderlying sediments as schematically illustratedin Figure 8.14.

Downward progressing sulfidization frontshave also been reported to be initiated bytransitions from limnic to brackish/marineconditions in the Baltic Sea (Böttcher and Lepland2000; Neumann et al. 2005) and the Black Sea(Jørgensen et al. 2004; Neretin et al. 2004). Incontrast to the example from the EasternMediterranean presented above, the sulfide dri-ving the downward sulfidization in thesesedimentary settings is derived primarily fromAOM and from the increase in sulfate concen-tration in the water column during the Holocene.At the sites on the western continental slope ofthe Black Sea investigated by Jørgensen et al.

Fig. 8.13 Concentration versus depth profiles of organiccarbon (Corg), total sulfur (Stot) and Fe/Al ratio (mol/g Fedivided by mol/g Al) through sapropel S7 in gravity coreGC17 from the eastern Mediterranean. The upper dark-grey bar marks the stratigraphical position of the sapropelvisible in the core. The lower light-grey bar marks thecoincident peaks of Stot and Fe/Al below the sapropel whichrepresent a Fe-sulfide band formed by Liesegang phenome-na. Modified from Passier et al. (1996).

Fig. 8.14 Model for formation of iron sulfide bandsbelow sapropels. The schematic depicted here representsa Liesegang situation (Berner 1969). The front at whichdownward diffusing hydrogen sulfide and upwarddiffusing iron react to form iron sulfides is fixed atparticular levels below the sapropel for a prolongedperiod of time. Modified from Passier et al. (1996).

Page 24: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

293

(2004) and Neretin et al. (2004) the sulfate-methanetransition is typically located around 2 m sedimentdepth. Sulfide liberated into the pore water at thisdepth is diffusing up into the anoxic water columnand is also drawn downward to a sulfidizationfront where it reacts with iron (oxyhydr)oxides andwith Fe2+ diffusing up from the deeper iron-richlimnic deposits (Fig. 8.15). The current depthposition of this HS-/Fe2+ diffusion front is markedby a black band which mostly consists ofamorphous iron sulfides termed AVS (acid volatilesulfide) in Fig. 8.15. These mineral phases are alsoresponsible for the distinct black coloration of thesediment. Above (i.e., behind) this downwardprogressing sulfidization front the amorphous ironsulfides have been and are still converted intopyrite as seen from a distinct horizon of greigiteand pyrite formation (c.f., CRS – chromiumreducible sulfur representing pyrite; Fig. 8.15).Due to the unusual progression in the reactionsequence towards pyrite, the degree of pyriti-zation (see Section 8.4.2) in this case decreaseswith increasing age of the deposits, i.e. the leastmature or stable iron sulfides are found at greatestsediment depth.

The examples given above illustrate that sulfideproduced by dissimilatory sulfate reduction, inparticular in organic-rich layers or within the zoneof AOM, can produce a profound diagenetic alte-ration of the sediment up to thousands or hun-

dreds of thousands of years after initial deposi-tion and thereby cause a delayed chemical,mineralogical, isotopic, and magnetic lock-in, i.e. aformation of a relatively stable sedimentary signalat a defined depth. As a consequence, the age ofthe particular authigenic mineral does notcorrespond to the age of the sediment layer, inwhich it is formed, but is much younger. Counter-intuitive as it may seem, in the case of downwardmoving sulfidization fronts the age of the mineralprecipitate becomes younger with increasingsediment depth. From these considerations itbecomes obvious that the post-depositional alte-rations of mineral phases and element associa-tions generated in this way complicate or evenprevent interpretations of the geochemical envi-ronment during the time of original sedimentdeposition.

8.5 Pathways of Sulfide Oxidation

Vast amounts of sulfide, corresponding to 7 mega-ton of H2S daily, are generated in marine sedimentsas the product of bacterial sulfate reduction. Asmall fraction of this sulfide is trapped within thesediment, mainly by reaction and precipitationwith iron to form pyrite, or by the sulfidization oforganic matter, and it thereby becomes buried in

Fig. 8.15 Sulfur geochemistry of a 4-m deep sediment core from the upper slope of the western Black Sea. Leftframe: SO4

2-, H2S, CH4 and Fe2+ (notice scales) in the pore water. The smooth curves are model fits to the data basedon the PROFILE model (Berg et al. 1998). Right frame: Chromium reducible sulfur (CRS) and acid volatile sulfide(AVS), the latter showing the black band of iron sulfide at 250-300 cm depth due to the downward progressingsulfidization front. From Jørgensen et al. (2004).

8.5 Pathways of Sulfide Oxidation

Page 25: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

294

the sea bed (Brüchert and Pratt 1996). In conti-nental slope and deep sea sediments this providesa net burial of reducing power over geologicaltime that contributes to maintain an oxidizedsurface of the earth and thus, together with theburial of organic carbon, to regulate the atmos-pheric oxygen level (Berner 1982, 1989). Bernerand Raiswell (1983) found that organic carbon andreduced sulfur were buried in the present oceanbed at a mean ratio of ca. 2.8 (±1.5) on a weightbasis. This corresponds to a Corg:Sred burial ratioof 7.4 on a molar basis. Since the completeoxidation of organic carbon to CO2 represents achange of 4 oxidation steps while the completeoxidation of reduced sulfur (pyrite) to sulfaterepresents 7 oxidation steps, the burial ratio is 4.3in terms of reducing equivalents. The latter ratiocorresponds to the amount of oxygen that canpotentially accumulate in the atmosphere due tothe burial of organic carbon versus reduced sulfur.Thus, the reduced sulfur is equivalent to ca. 20%of the buried reducing equivalents in the sea bed.

In deltaic and shelf sediments, where mostsulfide production takes place, the intensivesulfide burial during interglacial periods isinterrupted during glaciations that bind oceanwater in polar ice caps and thereby may lower thesea water level by 100 m. The glacial sea waterlow-stand causes erosion of accumulated shelfsediments and partly reoxidation of the exposedpyrite, thereby returning much of the accumulatedsulfide into the oceanic sulfate pool. Berner (1982)estimated the current burial of pyrite in marinesediments to be 39 Mt S yr-1 or 0.12 · 1013 mol S yr-1.This is equivalent to the sulfide production over5-6 days per year or to 1.6% of the total sulfatereduction in the global ocean floor (cf. Fig. 8.4).Thus, 98.4% of the produced sulfide is reoxidizedback to sea water sulfate on a geological timescale (million years).

The short-term (years to thousand years)burial of sulfide in the form of pyrite in oceanmargin sediments is more efficient and generallyaccounts for 5-20% of the entire sulfide produc-tion (Jørgensen et al. 1990; Lin and Morse 1991;Canfield and Teske 1996). This burial provides asink in the dynamic cycling of sulfur that is limitedby the availability of reactive iron to bind thelarge excess of sulfide. Raiswell and Canfield(1998) found that, of the total iron in marinesediments, on average only 25-28% is highlyreactive, 23-31% is poorly reactive, and 41-42% isunreactive. In spite of the small fraction of net

pyrite burial, a much larger fraction of the sulfurcycle, however, does go through pyrite and morelabile iron sulfides, in particular in the surficialsediment. These metal-bound sulfides are recycledwithin the sediment together with the free sulfideso that 80-95% of the entire sulfide production isgradually oxidized back to sulfate. The reoxidationtakes place at all depths of the sediment, mostrapidly in the upper oxidized zone but also in thedeeper and sulfidic part where the process is moredifficult to detect.

The oxidation of organic material by sulfate re-duction yields only a fraction of the energy that isavailable by aerobic respiration of the same orga-nic compounds. Consequently, a large part of thepotential chemical energy is still conserved in theproduct, H2S, from sulfate reduction and thisenergy may become available to other micro-organisms, provided a useful oxidant such as O2or NO3

- is present. In coastal sediments where theorganic deposition, and therefore the sulfatereduction, is particularly high, the reactive metaloxides may become completely reduced by sulfide.In this extreme case, H2S may diffuse freely up tothe sediment surface and reach the thin oxic skinof the surface sediment. A H2S-O2 interfacethereby develops within the uppermost few mm ofthe sediment where the gradient-type of colorlesssulfur bacteria may flourish on the chemicalenergy from H2S. Such hotspots of sulfide oxida-tion may be recognizable from the dark colorationof the sediment surface due to black iron sulfide(“black spots”; Rusch et al. 1998). The sedimentsmay also develop a distinct coating of filamentoussulfur bacteria, such as Beggiatoa, that store lightrefracting sulfur globules in their cells and thusprovide the sediment with a distinct whitishappearence. Such white Beggiatoa mats are typi-cal of the sediments around hydrothermal ventsand cold seeps that bring H2S from the subsurfacein direct contact with oxygenated sea water(Jannasch et al. 1989). In extreme cases where thewater column overlying the sediment is anoxic,e.g. in the permanently stratified Black Sea orduring summer in some eutrophic coastalenvironments, sulfide is not retained at the sedi-ment surface but penetrates directly up into thesea water (e.g., Roden and Tuttle 1992).

In oxic marine sediments, a brown layer rich iniron and manganese oxides generally separates O2and H2S and thereby prevents a direct sulfideoxidation by oxygen (e.g. Thamdrup et al. 1994a).In this suboxic zone, neither O2 nor H2S is present

Page 26: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

295

in detectable concentrations. The metal oxidesconstitute an efficient sulfide barrier by oxidizingand binding H2S that diffuses up from the sulfidicsediment below. In this case, H2S oxidation byoxygen is the exception and requires that themetal oxide layer is penetrated by advectivetransport, for example by bioirrigation by bur-rowing animals that pump oxic water for respi-ration directly down into the sulfidic sediment.Another advective mechanism may be current-induced advective pore water transport in poroussandy sediments (Huettel et al. 1998), or oxygentransport down into the root zone of sea grassbeds (Ballbjerg et al. 1998). Through such oxygenpenetration also pyrite may be oxidized withoxygen, a process that has been extensivelystudied (e.g. Lowson 1982; Luther 1987; Mosesand Herman 1991; Morse 1991). Pyrite oxidationwith O2 is a rather fast process that may be purelyabiotic, catalyzed by an electron shuttle betweenadsorbed Fe(II) and Fe(III) ions transferringelectrons from pyrite to O2. Sulfate is the endproduct of the sulfur oxidation and iron oxidesoften coat the surface of the oxidized pyritegrains.

Most sulfide oxidation in marine sediments isanoxic (i.e. takes place in the absence of oxygen)

and generally involves the precipitation of ironsulfide and the subsequent oxidation of iron-sulfur minerals back to sulfate. Evidence foranoxic sulfide oxidation comes from studies ofpore water chemistry, solid phase distributions ofmetal oxides and sulfides, mass balance calcula-tions, and direct experiments. By the use of radio-labeled H2S added to anoxic sediment cores orslurries, a rapid transfer of the label could beobserved into sulfur fractions defined as acidvolatile sulfide (mostly FeS), chromium reduciblesulfide (mostly FeS2), elemental sulfur (S0), orsulfate (Fossing and Jørgensen 1990). Theradiolabeled FeS and S0 were readily oxidized tosulfate in the anoxic sediment. Pyrite, in contrast,is more stable and is oxidized only over longerincubations. Yet, pyrite comprises the main sulfurpool in marine sediments and undergoes slowtransport and oxidation of critical importance forthe sulfur cycle.

These processes are illustrated in Fig. 8.16.Sulfate that penetrates down into the sedimentfrom the overlying sea water is reduced to H2S bysulfate reducing bacteria that use the depositedorganic material as their energy source. Alsomethane diffusing up from below feeds sulfatereduction in the lower sulfate zone. At depth in

Fig. 8.16 The sulfur cycle in marine sediments. The cycle is energetically driven by deposited organic material andmethane, both of which are used by sulfate reducing bacteria to produce H2S. Much of the H2S reacts chemically withiron (oxyhydr)oxides to form FeS and a range of intermediate oxidation states including S0 and FeS2. The furtheroxidation of these species back to sulfate is mediated by the vertical conveyer belt of bioturbation caused byburrowing macrofauna. Reoxidation of the solid phase sulfur species to sulfate at the sediment surface may be byoxygen, nitrate or manganese oxide. The same conveyer belt brings the oxidized iron back down towards the sulfideproduction zone where it reacts with further H2S. A highly efficient recycling of sulfur is thereby achieved. (FromJørgensen and Nelson 2004).

8.5 Pathways of Sulfide Oxidation

Page 27: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

296

marine sediments, the production of H2S mayexceed the availability of reactive metal oxides andH2S accumulates in the pore water (see Fig. 8.6,8.12, 8.13, and 8.17). The H2S diffuses upwardsalong a concentration gradient that generallyreaches zero at the bottom of the suboxic zone.Concurrently, the H2S reacts with buried ironoxides to form FeS, FeS2, S

0 and a number of othersolid or dissolved intermediate products. Once thereduced sulfur is bound in the solid phase, e.g. aspyrite, its further oxidation depends on a slowreaction with further Fe(III) species or its trans-port up to near-surface layers with oxidants ofhigher redox potential.

Pyrite transport in near-surface sedimentsgenerally takes place through the conveyer belt ofbioturbation whereby burrowing macrofauna mixthe sediment or directly move sediment particlesas part of their deposit feeding behavior (Fig.8.16). As the pyrite reaches up into the suboxic

zone it may react with oxidants such as oxygen ormanganese oxide and be converted into sulfateand iron oxides (or iron (oxyhydr)oxides). The ironoxides are in turn transported downwards throughthe same conveyer belt and thereby becomeavailable for further binding of sulfide and pyriteformation. The pyrite oxidation by manganeseoxide has been implied from chemical profiles(Canfield et al. 1993) and demonstrated directlythrough experiments (Schippers and Jørgensen2001, 2002). The process is interesting in that itinvolves the reaction between two mineral phasesin the sediment that must be in close proximity forthe oxidation to proceed. The initial reaction ispurely chemical and was proposed to occur by aFe(II)/Fe(III)-shuttle in the pore fluid between themineral surfaces of FeS2 and MnO2. The immediateproducts of the oxidation are thiosulfate andpolythionates. These can be further oxidized tosulfate by manganese reducing bacteria, thus

Fig. 8.17 Biogeochemical profiles of sulfur, manganese and iron species in a coastal marine sediment (Aarhus Bay,Denmark, 16 m water depth). A) Oxygen and nitrate profiles measured with O2 and NO3

- microsensors. B) Pore waterprofiles of dissolved manganese, iron and H2S. C) Profiles of solid phase oxidized manganese and iron and of pyrite. D)Distribution of sulfate reduction rates (SRR) measured by 35S-technique. The broken line at 4 cm depth indicates thetransition between the suboxic zone and the sulfidic zone. Data in A) were measured at the same site but a different yearthan data in B)-D). (Data from Kjær 2000 and Thamdrup et al. 1994a; reproduced from Jørgensen and Nelson 2004).

Page 28: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

297

making the complete pyrite oxidation to sulfatedependent on microbial catalysis (Schippers andJørgensen 2001):

FeS2 + 7.5 MnO2 + 11 H+ → Fe(OH)3 + 2 SO4

2- + 7.5 Mn2+ + 4 H2O (8.16)

Pyrite oxidation by nitrate could not be demon-strated through short-term sediment experimentsbut FeS is readily oxidized, both by nitrate andMnO2 (Aller and Rude 1988; Schippers 2004). Theimmediate product by FeS oxidation is not thiosul-fate but polysulfide which subsequently degradesinto elemental sulfur:

FeS + 1.5 MnO2 + 3 H+ → Fe(OH)3 + S0 + 1.5 Mn2+ (8.17)

The processes of sulfide oxidation describedhere are reflected in the chemical zonations ofpore water chemistry, solid phase chemistry andbacterial activity. As an example, Fig. 8.17shows data from organic-rich coastal sedimentat the Baltic Sea - North Sea transition. Oxygenpenetrated less than 0.4 cm into the sediment andshowed no contact with the zone of detectableH2S which started only from ca. 4 cm depth.Nitrate penetrated only slightly deeper thanoxygen, with a peak at 0.2 cm depth due to aerobicoxidation of ammonium (nitrification) diffusing upfrom the sediment below. The upper 0.4-4 cm ofthe sediment comprised the suboxic zone wheremaxima in dissolved reduced manganese and ironshowed the zones where the reduction of thesemetals was most intensive (peaks at 1- 2 cm depthfor manganese and 3-4 cm depth for iron). Theconcentration of solid phase manganese oxidedropped steeply with depth from the sedimentsurface to 1 cm depth where its reduction wasmost intensive. Oxidized iron decreased moregradually as it was most intensively reduced inthe lower part of the suboxic zone. Measurementsof sulfate reduction showed that H2S wasproduced throughout the sediment with maximumrates at the top of the sulfidic zone. Sulfatereduction also took place in the suboxic zone butthe produced H2S was here rapidly reoxidized anddid not reach detectable concentrations. Only bythe short-term experimental measurements using35SO4

2- could the SRR activity therefore bedemonstrated. Due to this sulfate reduction, a partof the manganese and iron reduction was drivenby sulfide oxidation while another part was due to

the direct oxidation of organic matter by hetero-trophic, metal-reducing bacteria (Thamdrup et al.1994a).

The intermediate sulfur species in sulfideoxidation, such as thiosulfate and elemental sulfur,are not stable in the sediment but are furthertransformed by microorganisms. Thiosulfate isturned over within hours or days and generallyoccurs only in sub-micromolar concentration inthe sediment pore water (Thamdrup et al. 1994b).Elemental sulfur accumulates to much higherconcentration in the solid phase but may also turnover on a seasonal or longer time scale in near-surface sediments (Troelsen and Jørgensen 1982).The preferred pathway of bacterial thiosulfate orsulfur transformation depends strongly on thechemical environment in the sediment. In the near-surface sediment with suitable oxidants such asoxygen, nitrate or metal oxides, these sulfurspecies may be used as energy sources bychemoautotrophic bacteria and be oxidizedcompletely to sulfate (Fig. 8.17, see also Chapter 5).When formed below the suboxic zone, they maybe used as oxidants (electron acceptors) in bacte-rial respiration to oxidize organic material.Through such a bacterial thiosulfate or sulfurrespiration these intermediates are reduced backto sulfide. When the availability of oxidants andorganic material are both limited, the two sulfurspecies may be disproportionated.

The ability of certain anaerobic bacteria todisproportionate intermediate sulfur species suchas thiosulfate was discovered only in the late1980’ies (Bak and Pfennig 1987; Krämer andCypionka 1989; Finster et al. 1998) but has sincebeen shown to play an important role in the sulfurcycle of aquatic sediments (Jørgensen 1990;Jørgensen and Bak 1991; Thamdrup et al. 1993). Itis characteristic for the disproportionation thatthe sulfur species are concurrently reduced tosulfide and oxidized to sulfate. This is an energyyielding reaction under appropriate sedimentconditions and is independent of externalreductants or oxidants. The process can beconsidered a type of inorganic fermentation and itprovides sufficient energy for bacteria to live on.By thiosulfate disproportionation, the inner(sulfonate) sulfur atom changes oxidation stepfrom +5 in S2O3

2- to +6 in SO42-, while the outer

(sulfane) atom changes from -1 in S2O32- to -2 in

H2S (Vairavamurthy at al. 1993; Eq. 8.18). Byelemental sulfur with an oxidation state of 0, someof the atoms are reduced to H2S and some oxidized

8.5 Pathways of Sulfide Oxidation

Page 29: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

298

to SO42- in a proportion of 3:1 that maintains

electron balance (Eq. 8.19):

S2O32- + H2O → H2S + SO4

2- (8.18)

4 S0 + 4 H2O → 3 H2S + SO42- + 2 H+ (8.19)

Disproportionation reactions do not cause anet oxidation of the sulfur species, yet they have akey function in sulfide oxidation. Dispropor-tionation provides a shunt in the sulfur cyclewhereby the H2S formed by this reaction may beoxidized again to the same sulfur intermediate bymetal oxides. Manganese oxide, for example,rapidly oxidizes H2S to S0 without participation ofbacteria, but does not oxidize the S0 further tosulfate (Burdige 1993). The elemental sulfur may,however, be disproportionated (Eq. 8.19) wherebya fourth of it is oxidized completely to sulfatewhile the remaining three fourths return to thesulfide pool. Through repeated partial oxidation ofsulfide to elemental sulfur with manganese oxideand subsequent disproportionation of the elemen-tal sulfur to sulfate and sulfide a complete oxida-tion of sulfide to sulfate by manganese oxide maybe achieved (Fig. 8.16; Thamdrup et al. 1993;Böttcher and Thamdrup 2001):

4 H2S + 4 MnO2 → 4 S0 + 4 Mn2+ + 8 OH-

(8.20)

Sum of Eq. 8.20 and 8.19:H2S + 4 MnO2 + 2 H2O →

SO42- + 4 Mn2+ + 6 OH- (8.21)

An interesting biological mechanism of sulfideoxidation in coastal upwelling regions and otherhigh-productivity coastal ecosystems was disco-vered in the mid 1990’ies (Fossing et al. 1995). Thesediment underlying some of the most intensiveupwelling systems, e.g. off the Pacific coast ofSouth and Central America or the Atlantic coast ofsouthwest Africa, is densely populated with sulfurbacteria (species of Thioploca, Thiomargarita,and Beggiatoa) that have a peculiar mode of lifeand an unusually large cell size (Schulz et al.1999). Within a liquid vacuole inside each cell theyaccumulate nitrate from the ambient sea water andlater use this nitrate down in the sediment as anelectron acceptor for sulfide oxidation from whichthey gain energy. The sulfide is oxidized first toelemental sulfur, which is stored transiently in thecells, and then to sulfate. This adaptation is

unique in that it enables the motile types of thebacteria to commute up and down between thenitrate source in the sea water and the sulfidesource in the sediment without having access toboth at the same time (Jørgensen and Gallardo1999; Schulz and Jørgensen 2001; see Chapter 6).

8.6 Determination of Process Rates

The rates of biogeochemical processes such assulfate reduction in marine sediments can bedetermined by different approaches that all havestrengths and weaknesses and may demonstratefundamentally different aspects of the process. Itis important for each application to considercarefully the properties of the sediment to beanalyzed and the limitations of the methodapplied. The study of shallow and highly dynamicsurface sediments requires a different approachthan the study of deep sediments that haveundergone stable diagenesis over a long timeperiod. By the shallow sediment an experimentalmeasurement of the process rate may be optimalwhereas deep in the sediments generally amodeling approach is preferred. Dependent on themethod, however, either the gross or the net rateof sulfate reduction is measured. The extent towhich these differ has been determined only in afew cases.

Gross sulfate reduction rates (gross SRR) canbe measured by experimental incubation ofrecovered sediment samples on board ship or inthe laboratory. Today, this is mostly done by theuse of radioactively labeled sulfate, 35SO4

2-, as atracer in order to obtain sufficiently high sensiti-vity of the method (see Chapter 5). Originallydeveloped by Sorokin (1962), Ivanov (1968),Jørgensen (1978), this method has been modifiedand refined over the years (e.g., Howarth 1979;Canfield et al. 1986; Fossing and Jørgensen 1989;Kallmeyer et al. 2004). By adaptation of theradiotracer method it has been possible to directlymeasure sulfate reduction rates that vary overmore than 7 orders of magnitude, for example onthe Peruvian shelf, from >1000 nmol SO4

2- cm-3 day-1

at the sediment surface to <0.001 nmol SO42- cm-3

day-1 at 100 m subsurface (Parkes et al. 2005; J.Kallmeyer et al. in prep.).

The gross SRR as measured by 35SO42-

technique comes closest to the overall sulfatereduction that takes place in the sediment. The

Page 30: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

299

radiotracer experiment generally lasts for only afew hours to a day and little reoxidation ofproduced 35S-sulfide takes place during this timeinterval, unless the sediment is rich in reactiveoxidized iron which may be the case in the suboxiczone near the sediment surface. The measurementshould be kept short in order to minimize changesin the chemistry or microbial activity of thesediment under the laboratory conditions used.This is an important criterion for the use ofradiotracer instead of just following the gradualdisappearence of pore water sulfate over time ashas been used earlier and is still applied whensafety considerations prohibit the use of radio-tracer. Tracking the sulfate concentration overtime is, however, a rather insensitive method thatmostly requires incubations over many days orweeks during which significant changes in mea-sured rates occur.

Another approach is based on the experimentthat nature has already done, namely by genera-ting decreasing sulfate concentrations downthrough the sediment column due to sulfatereduction acting over many years. The resultingpore water profile of sulfate reflects the rate of netSRR, which represents the gross SRR minus the(long-term) rate of reoxidation of reduced sulfurspecies back to sulfate. Such profiles, combinedwith solid-phase data on the sulfur chemistry andorganic carbon content, have been used tocalculate the distribution of sulfate reduction over

longer periods by 1-dimensional reaction-transport modeling (e.g., Berner 1980; Canfield1991; Schulz et al. 1994). Modeling of the net SRRgenerally assumes steady state conditions ofreduction rate, transport and sedimentation andrequires qualified information on the transportcoefficients of pore water species down throughthe sediment column. Since disturbing factorssuch as burrow irrigation by the benthic macro-fauna or current-induced advective pore waterflow may strongly enhance transport in additionto molecular diffusion, it is mostly difficult toprovide accurate transport coefficients in thenear-surface sediment. At depth in the sediment,however, molecular diffusion and steady state aremore likely to prevail. Here the modeling approachhas its obvious strength relative to the radiotracermeasurements that are increasingly prone todisturbance artifacts the deeper in the sedimentthe samples are taken.

In the following, we will use data of Jørgensenet al. (2004) from the Black Sea as an example toillustrate the differences between the twoapproaches. Fig. 8.18 shows results from the deepsulfidic part of the Black Sea where macrofaunaare unable to live and where bioirrigation is there-fore absent (although current-induced advectivepore water transport may take place near the sedi-ment surface). Sulfate reduction rates were mea-sured experimentally by the radiotracer methoddown to 20 cm depth in the sediment (Fig. 8.18 A).

Fig. 8.18 Determination of sulfate reduction rates in sediment cored by multicorer (A and B) and by gravity corer(C) from 1176 m water depth in the sulfidic part of the western Black Sea. A) Sulfate reduction rates (SRR) measuredexperimentally using 35SO4

2-, either directly in situ on the sea floor using a benthic lander or shipboard in thelaboratory. B) Mean experimental sulfate reduction rates (data points) and a model curve fitted to these rates. C)Sulfate concentrations measured in the pore water (data points) and modeled sulfate profile (smooth curve) thatmatches the fitted SRR in the uppermost 20 cm. After Jørgensen et al. (2001).

8.6 Dtermination of Process Rates

Page 31: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

300

A multi-G model of Westrich and Berner (1984)was applied to fit a smooth curve to these ratedata. The model assumes that the sedimentorganic matter that feeds sulfate reduction con-sists of three pools, each of which is degradedexponentially over depth and time, but eachhaving its own pool size and exponential decayconstant. Fig. 8.18 B and C show that this modelfits well with both the measured sulfate reductionrates in the upper sediment layer and the sulfatecurve in the deeper sediment down to 330 cmdepth.

A point to observe in Fig. 8.18 C is that themeasured and modeled profiles of pore watersulfate in the uppermost 0-20 cm hardly show anycurvature and, thus, indicate zero sulfate reduc-tion in this layer. The radiotracer measurements inFig. 8.18 B, however, show that 66% of the sulfatereduction in the entire sulfate zone takes place inthe top 15 cm where, in the deep sulfidic part ofthe Black Sea, sulfate reduction is the predo-minant pathway of organic carbon mineralization(Jørgensen et al. 2001). One possible reason forthis discrepancy is that modeling has lowsensitivity over narrow depth intervals becausediffusion is fast over short distances and therebyreduces changes in concentration in spite of highreaction rates. In contrast, over the entire sulfatezone even a deep zone of relatively low sulfatereduction rates may affect the entire sulfateprofile. This is why enhanced reduction rates inthe sulfate-methane transition tend to shape theentire sulfate distribution in the sediment as seenin Fig. 8.18 C. Thus, experimental measurementsare required to determine the sulfate reduction inthe near-surface sediment. Such a discrepancybetween measured and modeled rates is charac-teristic also of other sediments and is important toconsider in relation to the global SRR estimatesmade in Fig. 8.3 and 8.4. Data from ocean marginsediments were there mostly determined by theradiotracer method (gross SRR) whereas thosefrom the deep sea were modeled (net SRR).

8.7 The Sulfur Cycle

The sea bed functions as a giant anaerobic reactorwhere element cycles are coupled in a way thatdiffers fundamentally from the cycles in the oxicocean. The microbiological and geochemical pro-cesses of sulfur transformation thereby play key

functions that control the mineralization ofdeposited organic matter and thereby the chemis-try of the ocean. In the following we briefly sum-marize some main aspects of the sulfur cycle inmarine sediments and discuss its role for thedynamics of sulfate in the ocean.

Rather than being a simple cycle, composed ofanaerobic bacterial reduction of sulfate to hydro-gen sulfide and aerobic reoxidation of H2S to SO4

2-,the transformations of sulfur in aquatic sedimentsinclude a combination of intermediate cycles orshunts and interactions with other element cycles(e.g., Jørgensen 1990; Luther and Church 1991;Thamdrup et al. 1993; van Cappellen and Wang1996) schematically illustrated in Figure 8.16.Within this complex cycle, sulfur compoundsoccur in oxidation states ranging from -2 (H2S) to+6 (SO4

2-)By bacterial sulfate reduction H2S is produced

as the extracellular end-product (Widdel andHansen 1991). During the oxidation of H2S, oxic oranoxic, chemical or biological, compounds such aszero-valent sulfur (in elemental sulfur, poly-sulfides, or polythionates), thiosulfate (S2O3

2-), andsulfite (SO3

2-) are produced (Cline and Richards1969; Pyzik and Sommer 1981; Kelly 1988; DosSantos Afonso and Stumm 1992). These inter-mediates may then be further transformed by oneor several of the following processes:

• Respiratory bacterial reduction to H2S,• bacterial or chemical oxidation,• chemical precipitation (e.g. FeS formation), or• bacterial disproportionation to H2S and SO4

2-.

By bacterial disproportionation H2S and SO42-

are produced concurrently without participationof an external electron acceptor or donor (Bak andPfennig 1987; Thamdrup et al. 1993). The biogeo-chemical transformations of sulfur in marine sedi-ments are closely coupled to the cycles of ironand manganese. Sulfate, iron oxides, and manga-nese oxides all serve as electron acceptors in therespiratory degradation of organic matter. As thereare also non-enzymatic reactions between iron,manganese and H2S within the sediment, thequantification of dissimilatory, heterotrophic Feand Mn reduction is particularly difficult.

The precipitation of (authigenic) iron sulfidesresulting from the reaction between H2S and Fephases exerts an important control on the distri-bution of H2S in marine pore waters (Goldhaberand Kaplan 1974; Canfield 1989; Canfield et al.

Page 32: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

301

1992). Depending on the abundance and reactivityof the Fe(III) minerals present as well as on therate of sulfate reduction, H2S may be undetectabledespite its production at high rates. As the H2Saccumulates in the pore water, it diffuses uptowards the sediment surface where most of it isultimately oxidized back to sulfate. Only a smallfraction, 5-20%, of the produced H2S is trapped assolid phase iron-sulfur minerals or organic sulfurwithin the sediment. Over longer periods of thou-sands to millions of years, most of this trappedsulfur becomes exposed to oxidants and againreturns to the great sulfate reservoir of the ocean.

Over geological time, ocean sulfate is thuscontinuously recycled through sulfate reductionin the sea bed. During this recycling, the sulfuratom in sulfate undergoes a major redox transfor-mation of 8 oxidation steps between sulfate (+6)and sulfide (-2) and is thereby an effective oxidantfor the mineralization of marine organic matter.Importantly, the sulfur atom in sulfate is separatedfrom the oxygen atoms during the reductionprocess (SO4

2- → H2S), and is subsequentlycombined with new oxygen atoms from sedimentpore water upon reoxidation of the sulfide. Howfast is this recombination of the elements and howlong is the residence time of sulfate in sea water?We have calculated above that the magnitude ofsulfate reduction in the global sea bed is 7.5 · 1013

mol SO42- yr-1. The mean sulfate concentration of

sea water is 29 mM and the total volume of oceanwater is 1.37 billion km3 (Garrison 1997). Thus, theocean contains 1.37 · 1021 liter of sea water with atotal pool of 4.0 · 1019 mol SO4

2-. The turnover timeof this global ocean sulfate pool through bacterialsulfate reduction is (4.0 · 1019)/(7.5 · 1013) = ca. 0.5million years.

Sulfate is reduced in the sea bed not onlythrough bacterial catalysis but also through ther-mochemical catalysis at high temperature. At mid-oceanic ridges sea water is slowly convectedthrough the hot magmatic crust where a part of itis heated to >350°C (see Chapter 13). At such hightemperatures sulfate reacts as an oxidant forferrous iron and other reduced minerals and thesulfate is reduced to H2S without the participationof microorganisms which are excluded bytemperatures above ca. 100°C (Weber andJørgensen 2002). The global sea water flux thatundergoes heating to >350°C has been estimatedto be 3-6 · 1013 liter yr-1 (Elderfield and Schultz1996). By a complete reduction of the 29 mM ofsea water sulfate, this corresponds to the

reduction of 0.08-0.16 · 1013 mol SO42- yr-1 or 1-2%

of the global microbiological sulfate reduction.The turnover time of the entire volume of oceanwater at >350°C, and thus of the oceanic sulfatepool through thermochemical reduction, in hotmid-oceanic crust is thus 20-40 million years (cf.Elderfield and Schultz 1996).

The hydrothermal systems associated with themid-oceanic ridges harbor some of the richestbiological communities on the ocean floor andinclude giant clams, mussels, tube worms andshrimps. These animals are nourished by symbi-otic bacteria that oxidize sulfide and methane fromthe vent water and use the energy for chemo-autotrophic biomass production that in turn feedsthe hosts. It is an interesting perspective thattheir energy source is independent of light andphotosynthesis but is instead of geothermalorigin and based on chemical reactions at hightemperature. (It should be remembered, however,that the oxygen used for the oxidation of sulfideand methane is derived from photosynthesis inthe surface ocean, so the chemosynthetic pro-cesses would still not run without sunlight.)Although the hydrothermal vent communities arerich oases of life on the sea floor, their totalcontribution to the oceanic carbon cycle ismarginal. Bach and Edwards (2003) estimated thatthe global chemosynthetic biomass productionfrom oxidation of new ocean crust may be up to1012 g C yr-1. The total potential for chemosyn-thetic primary production at the deep sea hydro-thermal vents is globally estimated to be about1013 g biomass per year (McCollom and Shock1997). This represents only about 0.02% of theglobal primary production by photosynthesis inthe oceans. As the chemosynthetic productiontakes place in the otherwise nutrient-poor deepsea, however, it makes a much more importantcontribution to the local carbon supply to thedeep sea floor. Based on the oxygen uptake datapresented in Fig. 8.4, the global organic carbonmineralization by oxygen in the deep sea bed at>3000 m water depth is only 5 · 1013 mol yr-1,relative to which the chemoautotrophic produc-tion at mid-oceanic ridges corresponds to 20%.

8.7 The Sulfur Cycle

Page 33: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

302

Acknowledgements

This is contribution No 0333 of the Research CenterOcean Margins (RCOM) which is financed by theDeutsche Forschungsgemeinschaft (DFG) at BremenUniversity, Germany. BBJ was also supported by theMax Planck Society and the Fonds der ChemischenIndustrie and SK by the Helmholtz Society.

8.8 Problems

Problem 1

Organic matter is oxidized in the suboxic zonethrough iron or manganese reduction. This may bea direct oxidation by metal reducing bacteria or anindirect oxidation via sulfate reduction and sulfideoxidation. Does it matter for the end productswhich pathway dominates?

Problem 2

H2S can be oxidized completely to sulfate bymanganese oxide in marine sediments althoughthe chemical reaction oxidizes H2S only toelemental sulfur, S0. How is the oxidation tosulfate then possible?

Problem 3

Which factors determine how large a fraction ofthe deposited organic material in the sea bed ismineralized through sulfate reduction relative toother mineralization pathways?

Problem 4

Only about 5% of the total mineralized organiccarbon in marine sediments is ultimately degradedto methane. Why, then, does the anaerobicoxidation of methane with sulfate often generatequasi-linear sulfate profiles that indicate methaneto be the dominant energy source for sulfatereduction?

Problem 5

Which factors may cause a lock-in of a diageneticfront in the sea bed and thereby to the diageneticoverprinting of paleoceanographic signals in aspecific sediment horizon?

References

Aller, R.C., and Rude, P.D., 1988. Complete oxidationof solid phase sulfides by manganese and bacteria inanoxic marine sediments. Geochimica etCosmochimica Acta, 52: 751-765.

Amend, J.P., Edwards, K.J., and Lyons, T.W. (eds),2004. Sulfur Biogeochemistry - Past and Present.Geological Society of America Special Paper 379,Boulder, Colorado, 205 p.

Alperin, M.J., and Reeburgh, W.S., 1988. Carbon andhydrogen isotope fractionation resulting from anae-robic methane oxidation. Global Biogeoche-micalCycles, 2: 279-288.

Bach W., and Edwards, K.J., 2003. Iron and sulfideoxidation within the basaltic ocean crust: Implica-tions for chemolithoautotrophic microbial biomassproduction. Geochimica et Cosmochimica Acta, 67:3871-3887.

Bak, F., and Pfennig, N., 1987. Chemolithotrophicgrowth of Desulfovibrio sulfodismutans sp. nov. bydisproportionation of inorganic sulfur compounds.Archives of Microbiology, 147: 184-189.

Ballbjerg, V., Mouritsen, K.N., and Finster, K., 1998.Diel cycles of sulfate reduction rates in sediments ofa Zostera marina bed (Denmark). Aquatic MicrobialEcology, 15: 97-102.

Barnes, R.O., and Goldberg, E.D., 1976. Methaneproduction and consumption in anoxic marinesediments. Geology, 4: 297-300.

Berg, P., Rysgaard-Petersen, N. and Rysgaard, S., 1998.Interpretation of measured concentration profiles inthe sediment porewater. Limnology Oceanography,43: 1500-1510.

Berner, R.A., 1969. Migration of iron and sulfur withinanaerobic sediments during early diagenesis.American Journal of Science, 267: 19-42.

Berner, R.A., 1970. Sedimentary pyrite formation. Ame-rican Journal of Science, 268: 1-23.

Berner, R.A., 1980. Early diagenesis: A theoreticalapproach. Princeton University Press, Princeton,New Jersey.

Berner, R.A., 1981. A new geochemical classification ofsedimentary environments. Journal of SedimentaryPetrology, 51: 359-365.

Berner, R.A., 1982. Burial of organic carbon and pyritesulfur in the modern ocean: Its geochemical andenvironmental significance. American Journal ofScience, 282: 451-473.

Berner, R.A., 1984. Sedimentary pyrite formation: Anupdate. Geochimica et Cosmochimica Acta, 48: 605-615.

Berner, R.A., 1989. Biogeochemical cycles of carbonand sulfur and their effect on atmospheric oxygenover phanerozoic time. Global and Planetary Change75: 97-122.

Bishop, J.K.B., 1988. The barite-opal-organic carbonassociation in oceanic particulate matter. Nature,332: 341-343.

Boetius, A., and Suess, E., 2004. Hydrate Ridge: anatural laboratory for the study of microbial lifefueled by methane from near-surface gas hydrates.Chemical Geology, 205: 291-310.

Boetius, A., Ravenschlag, K., Schubert, C., Rickert, D.,Widdel, F., Gieseke, A., Amann, R., Jørgensen, B.B.,

Page 34: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

303

Witte, U., and Pfannkuche, O., 2000. A marinemicrobial consortium apparently mediatinganaerobic oxidation of methane. Nature, 407: 623-626.

Böttcher, M.E. and Lepland, A., 2000. Biogeochemistryof sulfur in a sediment core from the west-centralBaltic Sea: Evidence from stable isotopes and pyritetextures. Journal of Marine Systems, 25: 299-312.

Böttcher, M.E., and Thamdrup, B., 2001. Anaerobic sul-fide oxidation and stable isotope fractionation asso-ciated with bacterial sulfur disproportionation in thepresence of MnO2. Geochimica et CosmochimicaActa 65: 1573-1581.

Borowski, W.S., Paull, C.K., and Ussler III, W., 1996.Marine pore-water sulfate profiles indicate in-situmethane flux from underlying gas hydrate. Geology,24: 655-658.

Bowen, H.J.M., 1979. Environmental chemistry of theelements. Academic Press, London.

Bowman, J., 2000. The Methanotrophs - The familiesMethylococcaceae and Methylocystaceae. In:Dworkin, M., Balows, A., Trüper, H.G., Harder, W.,and Schleifer, K.-H. (eds), The Prokaryotes, 3rd Ed.Springer, New York.

Bréhéret, J.-G. and Brumsack, H.-J., 2000. Bariteconcretions as evidence of pauses in sedimentationin the Marnes Bleues Formation of the VocontianBasin (SE France). Sedimentary Geology, 130: 205-228.

Brüchert, V., and Pratt, L.M., 1996. Contemporaneousearly diagenetic formation of organic and inorganicsulfur in estuarine sediments from St. Andrew Bay,Florida, USA. Geochimica et Cosmochimica Acta 60:2325-2332.

Brumsack, H.-J., 1986. The inorganic geochemistry ofCretaceous black shales (DSDP Leg 41) incomparison to modern upwelling sediments from theGulf of California. In: Shackelton, N.J. andSummerhayes, C.P. (eds) North Atlantic paleoceano-graphy. Geological Society Special Publication 21,London, pp. 447-462.

Burdige, D.J., 1993. The biogeochemistry of manganeseand iron reduction in marine sediments. EarthScience Reviews, 35: 249-284.

Canfield, D.E., 1989. Sulfate reduction and oxicrespiration in marine sediments: implications fororganic carbon preservation in euxinic environ-ments. Deep-Sea Research, 36: 121-138.

Canfield, D.E., 1991. Sulfate reduction in deep-seasediments. American Journal of Science, 291: 177-188.

Canfield, D.E., 1993. Organic matter oxidation inmarine sediments. In: Wollast, R., Mackenzie, F.T.,and Chou, L. (eds), Biogeochemical cycles and globalchange. Springer, Berlin, pp. 333-363.

Canfield, D.E., and Teske, A., 1996. Late Proterozoicrise in atmospheric oxygen concentration inferredfrom phylogenetic and sulphur-isotope studies.Nature, 382: 127-132.

Canfield, D.E., Raiswell, R., Westrich, J.T., Reaves,C.M., and Berner, R.A., 1986. The use of chromiumreduction in the analyses of reduced inorganic sulfurin sediments and shales. Chemical Geology, 54: 149-155.

Canfield, D.E., Raiswell, R. and Bottrell, S., 1992. Thereactivity of sedimentary iron minerals towardsulfide. Am. J. Sci., 292: 659-683.

Canfield, D.E., Thamdrup, B. and Hansen, J.W., 1993.

The anaerobic degradation of organic matter inDanish coastal sediments: Iron reduction, manganesereduction, and sulfate reduction. Geochimica etCosmochimica Acta, 57: 3867-3883.

Canfield, D.E., Kristensen, E., and Thamdrup, B., 2005.Aquatic Geomicrobiology. Elsevier, San Diego,California.

Cecile, M.P., Shakur, M.A. and Krouse, H.R., 1983. Theisotopic composition of western Canadian baritesand the possible derivation of oceanic sulfate δ34Sand δ18O age curves. Canadian Journal of EarthSciences, 20: 1528-1535.

Chanton, J.P., 1985. Sulfur mass balance and isotopicfractionation in an anoxic marine sediment. Ph.D. The-sis, University of North Carolina, Chapel Hill, USA.

Christensen, J.P., 1989. Sulphate reduction and carbonoxidation rates in continental shelf sediments, anexamination of offshelf carbon transport. Continen-tal Shelf Research, 9: 223-246.

Cline, J.D. and Richards, F.A., 1969. Oxygenation ofhydrogen sulfide in seawater at constant salinity,temperature, and pH. Environment Sci. Technology,3: 838-843.

Crill, P.M., and Martens, C.S., 1983. Spatial and tem-poral fluctuations of methane production in anoxiccoastal marine sediments. Limnology and Oceano-graphy, 28: 1117-1130.

Crill, P.M., and Martens, C.S., 1986. Methane produc-tion from bicarbonate and acetate in an anoxicmarine sediment. Geochimica et CosmochimicaActa, 50: 2089-2097.

Crill, P.M., and Martens, C.S., 1987. Biogeochemicalcycling in an organic-rich coastal marine basin. 6.Temporal and spatial variations in sulfate reductionrates. Geochimica et Cosmochimica Acta, 51: 1175-1186 .

Dehairs, F., Stoobants, N. and Goeyens, L., 1991.Suspended barite as a tracer of biological activity inthe Southern Ocean. Marine Chemistry, 35: 399-410.

Devol, A.H. and Ahmend, S.I., 1981. Are high rates ofsulphate reduction associated with anaerobicoxidation of methane? Nature, 291: 407-408.

Devol, A.H., Anderson, J.J., Kuivila, K., and Murray,J.W., 1984. A model for coupled sulfate reductionand methane oxidation in the sediments of SaanichInlet. Geochimica et Cosmochimica Acta, 48: 993-1004 .

D’Hondt, S., Jørgensen, B.B., Miller, J., et al., 2003.Proceedings of the ODP, Initial Reports, 201 [CD-ROM]. Available from: Ocean Drilling Program,Texas A&M University, College Station, Texas77845-9547, USA.

D’Hondt, S., Jørgensen, B.B., Miller, D.J., et al., 2004.Distributions of microbial activities in deep subsea-floor sediments. Science, 306: 2216-2221.

Dickens, G.R., 2001. Sulfate profiles and barium frontsin sediment on the Blake Ridge: Present and pastmethane fluxes through a large gas hydrate reservoir.Geochimica et Cosmochimica Acta, 65: 529-543.

Dickens, G.R., Schroeder, D., Hinrichs, K.-U., and theLeg 201 Scientific Party, 2003. The pressure coresampler (PCS) on ODP Leg 201: general operationsand gas release. In: D’Hondt, S.L., Jørgensen, B.B.,Miller, D.J., et al. (eds), Proceedings of the ODP,Initial Reports, 201, 1-22 [CD-ROM]. Availablefrom: Ocean Drilling Program, Texas A&MUniversity, College Station TX 77845-9547, USA.

References

Page 35: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

304

Dos Santos Afonso, M. and Stumm, W., 1992. Thereductive dissolution of iron (III) (hydr) oxides byhydrogen sulfide. Langmuir, 8: 1671-1676.

Dymond, J., Suess, E. and Lyle, M., 1992. Barium indeep-sea sediment: A geochemical proxy for paleo-productivity. Paleoceanography, 7: 163-181.

Edenborn, H.M., Silverberg, N., Mucci, A., and Sundby,B., 1987. Sulfate reduction in deep coastal marinesediments. Marine Chemistry, 21: 329-345.

Ehrlich, H.L., 1996. Geomicrobiology. Marcel Dekker,NY, 719 pp.

Elderfield, H., and Schultz, A., 1996. Mid-ocean ridgehydrothermal fluxes and the chemical compositionof the ocean. Annual Review of Earth and PlanetaryScience, 24: 191-224.

Elvert, M., Suess, E., Greinert, J., and Whiticar, M.J.,2000. Archaea mediating anaerobic methane oxida-tion in deep-sea sediments at cold seeps of theeastern Aleutian subduction zone. Organic Geoche-mistry, 31: 1175-1187.

Ferdelman, T. G., Fossing, H., Neumann, K., and Schulz,H.D., 1999. Sulfate reduction in surface sediments ofthe southeast Atlantic continental margin between 15degrees 38´S and 27 degrees 57´S (Angola andNamibia). Limnology and Oceanography, 44: 650-661.

Finster, K., Liesack, W., and Thamdrup, B., 1998.Elemental sulfur and thiosulfate disproportionationby Desulfocapsa sulfoexigens sp. nov., a newanaerobic bacterium isolated from marine surfacesediment. Applied and Environmental Microbiology,564: 119-125.

Fossing, H., and Jørgensen, B.B., 1989. Measurement ofbacterial sulfate reduction in sediments: Evaluationof a single-step chromium reduction method. Bio-geochemistry, 8: 205-222.

Fossing, H., and Jørgensen, B.B., 1990. Oxidation andreduction of radiolabeled inorganic sulfur compoundsin an estuarine sediment, Kysing Fjord, Denmark.Geochimica et Cosmochimica Acta, 54: 2731-2742.

Fossing, H., Ferdelman, T.G., and Berg, P., 2000. Sulfatereduction and methane oxidation in continentalmargin sediments influenced by irrigation (South-East Atlantic off Namibia). Geochimica et Cosmo-chimica Acta, 64: 897-910.

Fossing, H., Gallardo, V.A., Jørgensen, B.B., Hüttel, M.,Nielsen, L.P., Schulz, H., Canfield, D.E., Forster, S.,Glud, R.N., Gundersen, J.K., Küver, J., Ramsing,N.B., Teske, A., Thamdrup, B., and Ulloa, O., 1995.Concentration and transport of nitrate by the mat-forming sulphur bacterium Thioploca . Nature, 374:713-715.

Froelich, P.N., Klinkhammer, G.P., Bender, M.L.,Luedtke, N.A., Heath, G.R., Cullen, D., Dauphin, P.,Hammond, D., Hartman, B., and Maynard, V., 1979.Early oxidation of organic matter in pelagicsediments of the eastern equatorial Atlantic: Suboxicdiagenesis. Geochimica et Cosmochimica Acta, 43:1075-1090.

Funk, J.A., von Dobeneck, T. and Reitz, A., 2003a.Integrated rock magnetic and geochemicalquantification of redoxomorphic iron mineraldiagenesis in Late Quaternary sediments from theEquatorial Atlantic. In: Wefer, G., Mulitza, S. andRatmeyer, V. (eds) The South Atlantic in the LateQuaternary: Reconstruction of Material Budget andCurrent Systems. Springer, Berlin, pp. 237-260.

Funk, J.A., von Dobeneck, T., Wagner, T. and Kasten,S., 2003b. Late Quaternary sedimentation and earlydiagenesis in the equatorial Atlantic Ocean:

Patterns, trends and processes deduced from rockmagnetic and geochemical records. In: Wefer, G.,Mulitza, S. and Ratmeyer, V. (eds) The SouthAtlantic in the Late Quaternary: Reconstruction ofMaterial Budgets and Current Systems. Springer,Berlin, pp. 461-497.

Garrison, T., 1997. Oceanography. Wadsworth Publi-shing Company, Belmont, California.

Giblin, A.E. and Howarth, R.W., 1984. Porewaterevidence for a dynamic sedimentary iron cycle insalt marshes. Limnology and Oceanography, 29: 47-63 .

Gingele, F.X. and Dahmke, A., 1994. Discrete bariteparticles and barium as tracers of paleoproductivityin South Atlantic sediments. Paleoceanography, 9:151-168.

Gingele, F.X., Zabel, M., Kasten, S., Bonn, W.J. andNürnberg, C.C., 1999. Biogenic barium - methodsand constraints of application as a proxy forpaleoproductivity. In: Fischer G. and Wefer, G. (eds),Use of proxies in paleoceanography: examples fromthe South Atlantic. Springer, Berlin, pp. 345-364.

Goldhaber, M.B. and Kaplan, I.R., 1974. The sulfurcycle. In: Goldberg, E.D. (ed), The Sea, 5, Wiley, pp569-655.

Goodfellow, W.D. and Jonasson, I.R., 1984. Oceanstagnation and ventilation defined by 34S seculartrends in pyrite and barite, Selwyn Basin, Yukon.Geology, 12: 583-586.

Greinert, J., Bollwerk, S.M., Derkachev, A., Bohrmann,G. and Suess, E., 2002. Massive barite deposits andcarbonate mineralization in the Derugin Basin, Seaof Okhotsk: precipitation processes at cold seeps.Earth and Planetary Science Letters, 203: 165-180.

Hallam, S.J., Girguis, P.R., Preston, C.M., Richardson,P.M., and DeLong, E.F., 2003. Identification ofmethyl coenzyme M reductase A (mcrA) genes asso-ciated with methane oxidizing archaea. Applied andEnvironmental Microbiology, 69: 5483-5491.

Hedges, J.I., and Keil, R.G., 1995. Sedimentary organicmatter preservation: An assessment and speculativesynthesis. Marine Chemistry, 49: 81-115.

Henrichs, S.M. and Reeburgh, W.S., 1987. Anaerobicmineralization of marine sediment organic matter:rates and the role of anaerobic processes in theoceanic carbon economy. Geomicrobiology Journal,5: 191-237.

Hensen, C., Zabel, M., Pfeifer, K., Schwenk, T., Kasten,S., Riedinger, N., Schulz, H.D. and Boetius, A., 2003.Control of sulfate pore-water profiles by sedimen-tary events and the significance of anaerobic oxida-tion of methane for burial of sulfur in marine sedi-ments. Geochimica et Cosmochimica Acta, 67:2631-2647.

Hinrichs, K.-U., and Boetius, A., 2002. The anaerobicoxidation of methane: New insights in microbialecology and biogeochemistry. In: Wefer, G., Billett,D., Hebbeln, D., Jørgensen, B.B., Schlüter, M., andvan Weering, T. (eds), Ocean margin Systems.Springer, Berlin, pp. 457-477.

Hinrichs, K.-U., Hayes, J.M., Sylva, S.P., Brewer, P.G.,and DeLong, E.F., 1999. Methane-consumingarchaebacteria in marine sediments. Nature, 398:802-805.

Hinrichs, K.-U., Summons, R.E., Orphan, V., Sylva, S.P.,and Hayes, J.M., 2000. Molecular and isotopicanalyses of anaerobic methane-oxidizing communi-ties in marine sediments. Organic Geochemistry, 31:1685-1701.

Page 36: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

305

Hoehler, T.M., Alperin, M.J. , Albert , D.B., andMartens, C.S. , 1994. Field and laboratory studiesof methane oxidation in an anoxic marine sedi-ment: Evidence for a methanogen-sulfate reducerconsortium. Global Biogeochemical Cycles, 8:451-463 .

Hoehler, T.M., Alperin, M.J., Albert, D.B., and Martens,C.S., 1998. Thermodynamic control on hydrogenconcentration in anoxic sediments. Geochimica etCosmochimica Acta, 62: 1745-1756.

Hoehler, T.M., Alperin, M.J., Albert, D.B., and Martens,C.S., 2001. Apparent minimum free energyrequirements for methanogenic Archaea and sulfate-reducing bacteria in an anoxic marine sediment.FEMS Microbiology Ecology, 38: 33-41.

Howarth, R.W., 1979. Pyrite: Its rapid formation in asalt marsh and its importance in ecosystemmetabolism. Science, 203: 49-51.

Huettel, M., Ziebis, W., Forster, S., and Luther III,G.W., 1998. Advective transport affecting metal andnutrient distributions and interfacial fluxes inpermeable sediments. Geochimica et CosmochimicaActa, 62: 613-631.

Ivanov, M.V., 1968. Microbiological processes in theformation of sulfur deposits. Israel Program forScientific Translations, Jerusalem.

Iversen, N., and Jørgensen, B.B., 1985. Anaerobicmethane oxidation rates at the sulfate-methanetransition in marine sediments from Kattegat andSkagerrak (Denmark). Limnology and Oceanogra-phy, 30: 944-955.

Jackson, B.E., and McInerney, M.J., 2002. Anaerobicmicrobial metabolism can proceed close to thermo-dynamic limits. Nature, 415: 454-456.

Jahnke, R.A., 1996. The global ocean flux of particulateorganic carbon: Areal distribution and magnitude.Global Biogeochemical Cycles, 10: 71-88.

Jannasch, H.W., Nelson, D.C., and Wirsen, C.O., 1989.Massive natural occurrence of unusually largebacteria (Beggiatoa sp.) at a hydrothermal deep-seavent site. Nature, 342: 834-836.

Jørgensen, B.B., 1977. The sulfur cycle of a coastalmarine sediment (Limfjorden, Denmark). Limnologyand Oceanography, 22: 814-832.

Jørgensen, B.B., 1978. A comparison of methods for thequantification of bacterial sulfate reduction incoastal marine sediments. I. Measurements withradiotracer techniques. Geomicrobiology Journal, 1:11-27.

Jørgensen, B.B., 1982. Mineralization of organic matterin the sea bed - The role of sulfate reduction.Nature, 296: 643-645.

Jørgensen, B.B., 1983. Processes at the sediment-waterinterface. In: Bolin, B. and Cook, R.C. (eds), Themajor biogeochemical cycles and their interactions.SCOPE, pp. 477-509.

Jørgensen, B.B., 1990. A thiosulfate shunt in the sulfurcycle of marine sediments. Science, 249: 152-154.

Jørgensen, B.B., 1996. Case Study: Aarhus Bay. In:Jørgensen, B.B., and Richardson, K. (eds), Eutrophi-cation in a coastal marine environment. Coastal andEstuarine Studies, American Geophysical Union,Washington, DC, pp. 137-154

Jørgensen, B.B., and Bak, F., 1991. Pathways andmicrobiology of thiosulfate transformations andsulfate reduction in a marine sediment (Kattegat,Denmark). Applied and Environmental Microbio-logy, 57: 847-856.

Jørgensen, B.B., and Gallardo, V.A., 1999. Thioplocaspp.: filamentous sulfur bacteria with nitratevacuoles. FEMS Microbiology Ecology, 28: 301-313.

Jørgensen, B.B., and Nelson D.C., 2004. Sulfideoxidation in marine sediments: Geochemistry meetsmicrobiology. In.: Amend, J.P., Edwards, K.J., andLyons, T.W. (eds), Sulfur Biogeochemistry - Pastand Present. Geological Society of America SpecialPaper 379, Boulder, Colorado, pp. 63-81.

Jørgensen, B.B., Bang, M., and Blackburn, T.H., 1990.Anaerobic mineralization in marine sediments fromthe Baltic Sea - North Sea transition. MarineEcology Progress Series, 59: 39-54.

Jørgensen, B.B., Weber, A., and Zopfi, J., 2001. Sulfatereduction and anaerobic methane oxidation in BlackSea sediments. Deep-Sea Research, 48: 2097-2120.

Jørgensen, B.B., Böttcher, M.E., Lüschen, H., Neretin,L., and Volkov, I., 2004. Anaerobic methane oxida-tion and a deep H2S sink generate isotopically heavysulfides in Black Sea sediments. Geochimica et Cos-mochimica Acta, 68: 2095-2118.

Judd, A.G., Hovland, M., Dimitrov, I.I., Garcia Gil, S.,and Jukes, V., 2002. The geological methane budgetat Continental Margins and its influence on climatechange. Geofluids, 2: 109-126.

Kallmeyer, J., Ferdelman, T.G., Weber, A., Fossing, H.,and Jørgensen, B. B., 2004. A cold chromiumdistillation procedure for radiolabeled sulfide appliedto sulfate reduction measurements. Limnology andOceanography Methods, 2: 171-180.

Karlin, R. and Levi, S., 1983. Diagenesis of magneticminerals in recent hemipelagic sediments. Nature,303: 327-330.

Karlin, R. and Levi, S., 1985. Geochemical andsedimentological control of the magnetic propertiesof hemipelagic sediments. Journal of GeophysicalResearch, 90: 10373-10392.

Kasten, S., Zabel, M., Heuer, V. and Hensen, C., 2003.Processes and signals of nonsteady-state diagenesisin deep-sea sediments and their pore waters. In:Wefer, G., Mulitza, S. and Ratmeyer, V. (eds), TheSouth Atlantic in the Late Quaternary: Recon-struction of Material Budget and Current Systems.Springer, Berlin, pp. 431-459.

Kasten, S., Freudenthal, T., Gingele, F.X., von Dobeneck,T. and Schulz, H.D., 1998. Simultaneous formation ofiron-rich layers at different redox boundaries insediments of the Amazon Deep-Sea Fan. Geochimicaet Cosmochimica Acta 62: 2253-2264.

Kelly, D.P., 1988. Oxidation of sulfur compounds. In:Cole, A.S. and Ferguson, S.J. (eds), The Nitrogen andSulfur Cycles. Soc. Gen. Microbiol., 42, pp. 65-98.

Kjær, T., 2000. Development and application of newbiosensors for microbial ecology. Ph.D. Thesis,University of Aarhus, Denmark, 324 p.

Knittel, K., Lösekann, T., Boetius, A., Kort, R., andAmann, R., 2005. Diversity and distribution ofmethanotrophic archaea at cold seeps. Applied andEnvironmental Microbiology, 71: 467-479.

Kölling, A., 1991. Frühdiagenetische Prozesse und Stoff-Flüsse in marinen und ästuarinen Sedimenten.Berichte, 15, Fachbereich Geowissenschaften, Uni-versität Bremen, 140 pp.

Krämer, M., and Cypionka, H., 1989. Sulfate formationvia ATP sulfurylase in thiosulfate- and sulfite-disproportionating bacteria. Archives of Microbio-logy, 122: 183-188.

References

Page 37: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

306

Krein, E.B. and Aizenshtat, Z., 1995. Proposed thermalpathways for sulfur transformations in organicmacromolecules: Laboratory simulation experi-ments. In: Vairavamurthy, M.A. and Schoonen,M.A.A. (eds), Geochemical Transformations of Sedi-mentary Sulfur, ACS symposium series, 612,Washington, DC, pp. 110-137.

Krüger, M., Meyerdierks, A., Gloeckner, F.O., Amann,R., Widdel, F., Kube, M., Reinhardt, R., Kahnt, J.,Boecher, R., Thauer, R.K., and Shima, S., 2003. Aconspicuous nickel protein in microbial mats thatoxidize methane anaerobically. Nature, 426: 878-881.

Lin, S., and Morse, J.W., 1991. Sulfate reduction andiron sulfide mineral formation in Gulf of Mexicoanoxic sediments. American Journal of Science, 291:55-89.

Lowson, R.T., 1982. Aqueous oxidation of pyrite bymolecular oxygen. Chemical Reviews, 82: 461-497.

Luther, G.W., III, 1987. Pyrite oxidation and reduction:Molecular orbital theory considerations. Geochimicaet Cosmochimica Acta, 51: 3193-3199.

Luther III, G.W., 1991. Pyrite synthesis via polysulfidecompounds. Geochimica et Cosmochimia Acta, 55:2839-2849.

Luther III, G.W. and Church, T.M., 1991. An overviewof the environment chemistry of sulfur in wetlandsystems. In: Howarth, R.W. et al (eds), Sulfurcycling on the continents. John Wiley, pp: 125-144.

Luther III, G.W., Giblin, A., Howarth, R.W. and Ryans,R.A., 1982. Pyrite and oxidized iron mineral phasesformed from pyrite oxidation in salt marsh andestuarine sediments. Geochimica et CosmochimicaActa, 46: 2665-2669.

Martens, C.S., and Berner, R.A., 1974. Methaneproduction in the interstitial waters of sulfate-depleted marine sediments. Science, 185: 1167-1169 .

McCollom, T.M., and Shock, E.L., 1997. Geochemicalconstraints on chemolithoautotrophic metabolismby microorganisms in seafloor hydrothermalsystems. Geochimica et Cosmochimica Acta, 61:4375-4391.

Mearon, S., Paytan, A. and Bralower, T.J., 2003.Cretaceous strontium isotope stratigraphy usingmarine barite. Geology, 31: 15-18.

Michaelis, W., Seifert, R., Nauhaus, K., Treude, T.,Thiel, V., Blumenberg, M., Knittel, K., Gieseke, A.,Peterknecht, K., Pape, T., Boetius, A., Amann, R.,Jørgensen, B.B., Widdel, F., Peckmann, J., Pimenov,N.V., and Gulin, M.B., 2002. Microbial reefs in theBlack Sea fueled by anaerobic oxidation of methane.Science, 297: 1013-1015.

Middelburg, J.B.M., 1990. Early diagenesis andauthigenic mineral formation in anoxic sediments ofKau Bay, Indonesia. PhD Thesis, Universitiy ofUtrecht, Utrecht, 177 pp.

Middelburg, J.J., Soetaert, K., and Herman, P.M.J.,1997. Empirical relationships for use in globaldiagenetic models. Deep-Sea Research I, 44: 327-344.

Moore, T.S., Murray, R.W., Kurtz, A.C. and Schrag,D.P., 2004. Anaerobic methane oxidation and theformation of dolomite. Earth and Planetary ScienceLetters, 229: 141-154.

Morse, J.W., 1991. Oxidation kinetics of sedimentarypyrite in seawater. Geochimica et CosmochimicaActa, 55: 3665-3667.

Morse, J.W., 2002. Sedimentary geochemistry of thecarbonate and sulphide systems and their potentialinfluence on toxic metal bioavailability. In. Gian-guzza, A., Pelizzetti, E. and Sammartano, S. (eds),Chemistry of Marine Water and Sediments. Springer,pp. 165-189.

Moses, C.O., and Herman, J.S., 1991. Pyrite oxidationat circumneutral pH. Geochimica et CosmochimicaActa, 55: 471-482.

Murray, J.W., Grundmanis, V. and Smethie, W.M. Jr,1978. Interstitial water chemistry in sediments ofSaanich Inlet. Geochimica et Cosmochimica Acta,42: 1011-1026.

Nauhaus, K., Boetius, A., Krüger, M., and Widdel, F.,2002. In vitro demonstration of anaerobic oxidationof methane coupled to sulphate reduction insediment from a marine gas hydrate area. Environ-mental Microbiology, 4: 296-305.

Neretin, L., Böttcher, M.E., Jorgensen, B.B., Volkov,I.I., Lüschen, H. and Hilgenfeldt, K., 2004. Pyriti-zation processes and greigite formation in theadvancing sulfidization front in the upper Pleisto-cene sediments of the Black Sea. Geochimica etCosmochimica Acta, 68: 2081-2093.

Neumann, T., Rausch, N., Leipe, T., Dellwig, O., Berner,Z. and Böttcher, M.E., 2005. Intense pyriteformation under low-sulfate conditions in the Ach-terwasser lagoon, SW Baltic Sea. Geochimica etCosmochimica Acta, 69: 3619-3630.

Niewöhner, C., Hensen, C., Kasten, S., Zabel, M. andSchulz, H.D., 1998. Deep sulfate reduction com-pletely mediated by anaerobic methane oxidation insediments of the upwelling area off Namibia.Geochimica et Cosmochimica Acta, 62: 455-464.

Nüster, J., 2005. New methods in biogeochemistry: Thedevelopment of electrochemical tools for themeasurement of dissolved and solid state compoundsin natural systems. PhD Thesis, University ofBremen, 216 pp.

Orphan, V.J., House, C.H. Hinrichs, K.-U., McKeegan,K.D., DeLong, E.F., 2001. Methane-consumingarchaea revealed by directly coupled isotopic andphylogenetic analysis. Science, 293: 484-487.

Orphan, V.J., House, C.H., Hinrichs, K.-U., McKeegan,K.D., DeLong, E.F., 2002. Multiple archaeal groupsmediate methane oxidation in anoxic cold seepsediments. Proceedings of the National Academy ofSciences USA, 99: 7663-7668.

Orr, W.L. and White, C.M. (eds), 1990. Geochemistryof sulfur in fossil fuels. ACS Symposium Series, 429,Washington, DC.

Pancost, R.D., Sinninghe Damsté, J.S., Lint, S.D., vander Marel, M.J.E.C., Gottschal, J.C., and ShipboardScientific Party, 2000. Biomarker evidence forwidespread anaerobic methane oxidation inMediterranean sediments by a consortium ofmethanogenic archaea and bacteria. Applied andEnvironmental Microbiology, 66: 1126-1132.

Parkes, R.J., Webster, G., Cragg, B.A., Weightman, A.J.,Newberry, C.J., Ferdelman, T.G., Kallmeyer, J.,Jørgensen, B.B., Aiello, I.W., and Fry, J.C., 2005.Deep sub-seafloor prokaryotes stimulated atinterfaces over geological time. Nature, 436: 390-394.

Passier, H.F., Middelburg, J.J., Os, B.J.H.v. and Lange,G.J.de, 1996. Diagenetic pyritisation under easternMediterranean sapropels caused by downward sul-phide diffusion. Geochimica et Cosmochimica Acta,60: 751-763.

Page 38: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

307

Paytan, A., Kastner, M., Martin, E.E., MacDougall,J.D., and Herbert, T., 1993. Marine barite as a moni-tor of seawater strontium isotope composition.Nature, 366: 445-449.

Paytan, A., Kastner, M. and Chavez, F.P., 1996a.Glacial to interglacial fluctuations in productivity inthe equatorial Pacific as indicated by marine barite.Science, 274: 1355-1357.

Paytan, A., Moore, W.S. and Kastner, M., 1996b. Sedi-mentation rate as determined by 226Ra activity inmarine barite. Geochimica et Cosmochi-mica Acta,60: 4313-4319.

Paytan, A., Kastner, M., Campbell, D. and Thiemens,M.H., 1998. Sulfur isotope composition of Cenozoicseawater sulfate. Science, 282: 1459-1462.

Paytan, A., Mearon, S., Cobb, K. and Kastner, M., 2002.Origin of marine barite deposits: Sr and S isotopecharacterization. Geology, 30: 747-750.

Paytan, A., Martinez-Ruiz, F., Eagle, M., Ivy, A. andWankel, S.D., 2004. Using sulfur isotopes toelucidate the origin of barite associated with highorganic matter accumulation events in marinesediments. In: Amend, J.P., Edwards, K.J. and Lyons,T.W. (eds), Sulfur Biogeochemistry – Past andPresent. Geological Society of America SpecialPaper 379, pp. 151-160.

Poulton, S.W., Krom, M.D. and Raiswell, R., 2004. Arevised scheme for the reactivity of iron(oxyhydr)oxides towards dissolved sulfide.Geochimica et Cosmochimica Acta, 68: 3703-3715.

Pyzik, A.J. and Sommer, S.E., 1981. Sedimentary ironmonosulfides: kinetics and mechanism of formation.Geochimica et Cosmochimica Acta, 45: 687-698.

Rabus, R., Hansen, T., and Widdel, F., 2004.Dissimilatory Sulfate- and Sulfur Reducing Proka-ryotes. In: Dworkin, M. et al. (eds), The Proka-ryotes: An Evolving Electronic Resource for theMicrobiological Community, 3rd edition, Springer-Verlag, New York, (http://link.springer-ny.com/link/service/books/10125/).

Raiswell, R., 1982. Pyrite texture, isotopic compositionand the availability of iron. American Journal ofScience, 282: 1244-1265.

Raiswell, R., 1988. Chemical model for the origin ofminor limestone-shale cycles by anaerobic methaneoxidation. Geology, 16: 641-644.

Raiswell, R., and Canfield, D.E., 1998. Sources of ironfor pyrite formation in marine sediments. AmericanJournal of Science, 298: 219-245.

Raiswell, R., Canfield, D.E. and Berner, R.A., 1994. Acomparison of iron extraction methods for thedetermination of degree of pyritization and the re-cognition of iron-limited pyrite formation. Chemi-cal Geology, 111: 101-110.

Reeburgh, W.S., 1969. Observations of gases inChesapeake Bay sediments. Limnology and Oceano-graphy, 14: 368-375.

Reeburgh, W.S., 1976. Methane consumption in CariacoTrench waters and sediments. Earth and PlanetaryScience Letters, 47: 345-352.

Reeburgh, W.S., 1982. A major sink and flux control formethane in sediments: Anaerobic consumption. In:Fanning, K.A. and Manheim, F.T. (eds), Thedynamic environment. Heath, Lexington, MA, pp.203-217.

Reeburgh, W.S., Whalen, S.C., and Alperin, M.J., 1993.The role of methylotrophy in the global methanebudget. In: Murrell, J.C., and Kelly, D.P. (eds),

Microbial growth on C1 compounds. Intercept,Andover, UK, pp. 1-14.

Reitz, A., Hensen, C., Kasten, S., Funk, J. and de Lange,G., 2004. A combined geochemical and rock-magnetic investigation of a redox horizon at the lastglacial/interglacial transition. Physics and Chemistryof the Earth, 29: 921-931.

Rickard, D.T., 1975. Kinetics and mechanisms of pyriteformation at low temperatures. American Journal ofScience, 275: 636-652.

Rickard, D. and Luther III, G.W., 1997. Kinetics ofpyrite formation by the H2S oxidation of iron(II)monosulfide in aqueous solutions between 25 and125°C: The rate equation. Geochimica et Cosmo-chimica Acta, 61: 115-134.

Rickard, D., Schoonen, M.A.A., Luther, G.W., 1995.Chemistry of iron sulfides in sedimentaryenvironments. In: Vairavamurthy, M.A. andSchoonen, M.A.A. (eds), Geochemical Transfor-mations of Sedimentary Sulfur. ACS SymposiumSeries 612, Washington DC, pp. 168-193.

Riedinger, N., 2005. Preservation and diageneticoverprint of geochemical and geophysical signalsin ocean margin sediments related to depositionaldynamics. Berichte, 242, Fachbereich Geowis-senschaften, Universität Bremen, 91 pp.

Riedinger, N., Pfeifer, K., Kasten, S., Garming, J.F.L.,Vogt, C. and Hensen, C., 2005. Diagenetic al te-ration of magnetic signals by anaerobic oxidationof methane related to a change in sedimentationrate. Geochimica et Cosmochimica Acta, 69:4 1 1 7 - 4 1 2 6 .

Roden, E.E., and Tuttle, J.H., 1992. Sulfide release fromestuarine sediments underlying anoxic bottom water.Limnology and Oceanograpy 37: 725-738.

Rusch, A., Töpken, H., Böttcher, M.E., and Höpner, T.,1998. Recovery from black spots: results of aloading experiment in the Wadden Sea. Journal ofSea Research, 40: 205-219.

Schink, B., 1997. Energetics of syntrophic cooperationin methanogenic degradation. Microbiological andMolecular Biological Reviews, 61: 262-280.

Schinzel, U., 1993. Laboratory experiments on early dia-genetic reactions of iron(III) oxyhydroxides in marinesediments (in German). Berichte, 36, FachbereichGeowissenschaften, Universität Bremen, 189 pp.

Schippers, A., 2004. Biogeochemistry of metal sulfideoxidation in mining environments, sediments, andsoils. In: Amend, J.P., Edwards, K.J., and Lyons, T.W.(eds), Sulfur biogeochemistry - Past and Present.Geological Society of America Special Paper 379,Boulder, Colorado, pp. 49-62

Schippers, A., and Jørgensen, B.B., 2001. Oxidation ofpyrite and iron sulfide by manganese in marinesediments. Geochimica et Cosmochimica Acta, 65:915-922.

Schippers, A., and Jørgensen, B.B., 2002. Biogeo-chemistry of pyrite and iron sulfide oxidation inmarine sediments. Geochimica et CosmochimicaActa, 66: 85-92.

Schoonen, M.A.A., 2004. Mechanisms of sedimentarypyrite formation. In: Amend, J.P., Edwards, K.J. andLyons, T.W. (eds), Sulfur Biogeochemistry – Past andPresent. Geological Society of America Special Paper379, pp. 117-134.

Schouten, S., Eglington, T.I., Sinninghe Damsté, J.S. andde Leeuw, J.W., 1995. Influence of sulfur cross-linking on the molecular size distribution of sulfur-rich macromolecules in bitumen. In: Vairavamurthy,

References

Page 39: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

308

M.A. and Schoonen, M.A.A. (eds), Geochemicaltransformation of sedimentary sulfur. ACS Sympo-sium 612, Washington, DC, pp. 80-92.

Schulz H.D., Dahmke A., Schinzel U., Wallmann K., andZabel M., 1994. Early diagenetic processes, fluxesand reaction rates in sediments of the South-Atlantic.Geochimica et Cosmochimica Acta, 58: 2041-2060.

Schulz, H.N., and Jørgensen, B.B., 2001. Big bacteria.Annual Reviews in Microbiology, 55: 105-137.

Schulz, H.N., Brinkhoff, T., Ferdelman, T.G., HernandezMarine, M., Teske, A., and Jørgensen, B.B., 1999.Dense populations of a giant sulfur bacterium inNamibian shelf sediments. Science, 284: 493-495.

Smith, S.V., and Hollibaugh, J.T., 1993. Coastal meta-bolism and the oceanic organic carbon balance.Reviews in Geophysics, 31: 75-89.

Sørensen, K.B., Finster, K., and Ramsing, N.B., 2001.Thermodynamic and kinetic requirements inanaerobic methane oxidizing consortia exclude hydro-gen, acetate, and methanol as possible electronshuttles. Microbial Ecology, 42: 1-10.

Sorokin, Yu.L., 1962. Experimental investigation ofbacterial sulfate reduction in the Black Sea using S35.Microbiology, 31: 329-335.

Stetter, K.O., Huber, R., Blöchl, E., Kurr, M., Eden, R.D.,Fielder, M., Cash, H., and Vance, I., 1993. Hyper-thermophilic archaea are thriving in deep North Seaand Alaskan oil reservoirs. Nature, 365: 743-745.

Sweeney, R.E. and Kaplan, I.R., 1973. Pyrite FramboidFormation: Laboratory Synthesis and Marine Sedi-ments. Economic Geology, 68: 618-634.

Thamdrup, B., and Canfield, D.E., 1996. Pathways ofcarbon oxidation in continental margin sediments offcentral Chile. Limnology and Oceanography, 41:1629-1650.

Thamdrup, B., Finster, K, Hansen, J.W., and Bak, F.,1993. Bacterial disproportionation of elementalsulfur coupled to chemical reduction of iron ormanganese. Applied and Environmental Microbio-logy, 59: 101-108.

Thamdrup, B., Fossing, H., and Jørgensen, B.B., 1994a.Manganese, iron, and sulfur cycling in a coastalmarine sediment, Aarhus Bay, Denmark. Geochimicaet Cosmochimica Acta, 58: 5115-5129.

Thamdrup, B., Finster, K., Fossing, H., Hansen, J.W., andJørgensen, B.B., 1994b. Thiosulfate and sulfitedistributions in porewater of marine sediments relatedto manganese, iron and sulfur geochemistry. Geochi-mica et Cosmochimica Acta, 58: 67-73.

Thiel, V., Peckmann, J., Richnow, H.H., Luth, U., Reit-ner, J., and Michaelis, W., 2001. Molecular signals foranaerobic methane oxidation in Black Sea seepcarbonates and a microbial mat. Marine Chemistry,73: 97-112.

Thode-Andersen, S., and Jørgensen, B.B., 1989. Sulfatereduction and the formation of 35S-labeled FeS, FeS2,and S° in coastal marine sediments. Limnology andOceanography, 34: 793-806.

Torres, M.E., Brumsack, H.J., Bohrmann, G. and Emeis,K.C., 1996. Barite fronts in continental marginsediments: A new look at barium remobilization in thezone of sulfate reduction and formation of heavybarites in diagenetic fronts. Chemical Geology, 127:125-139.

Treude, T., Niggemann, J., Kallmeyer, J., Wintersteller,P., Schubert, C.J., Boetius, A., and Jørgensen, B.B.,2005. Anaerobic oxidation of methane and sulfatereduction along the Chilean continental margin.

Geochimica et Cosmochimica Acta, 69: 2767-2779.Troelsen, H., and Jørgensen, B.B., 1982. Seasonal

dynamics of elemental sulfur in two coastal sedi-ments. Estuarine and Coastal Shelf Science, 15: 255-266.

Vairavamurthy, A., Manowitz, B., Luther III, G.W., Jeon,Y., 1993. Oxidation state of sulfur in thiosulfate andimplications for anaerobic energy metabolism. Geo-chimica et Cosmochimica Acta, 57: 1619-1623.

Vairavamurthy, M.A., Orr, W.L. and Manowitz, B., 1995.Geochemical transformation of sedimentary sulfur: anintroduction. In: Vairavamurthy, M.A. and Schoonen,M.A.A. (eds), Geochemical tranformation of sedi-mentary sulfur. ACS Symposium, 612, Washing-ton,DC, pp. 1-17.

Valentine, D.L., and Reeburgh, W.S., 2000. Newperspectives on anaerobic methane oxidation - Mini-review. Environmental Microbiology, 2: 477-484.

Van Beek, P. and Reyss, J.-L., 2001. 226Ra in marinebarite: New constraints on supported 226Ra. EarthPlanetary Science Letters, 187: 147-161.

Van Beek, P., Reyss, J.-L., Paterne, M., Gersonde, R.,Rutgers van der Loeff, M. and Kuhn, G., 2002. 226Rain barite: Absolute dating of Holocene SouthernOcean sediments and reconstruction of sea-surfacereservoir ages. Geology, 30: 731-734.

van Cappellen, P., and Wang, Y., 1996. Cycling of ironand manganese in surface sediments: A general theoryfor the coupled transport and reaction of carbon,oxygen, nitrogen, sulfur, iron, and manganese. Ame-rican Journal of Science, 296: 197-243.

Von Breymann, M.T.K., Emeis, K.C. and Suess, E., 1992.Water depth and diagenetic constraints on the use ofbarium as a paleoproductivity indicator. In: Summer-hayes, C.P. (ed) Upwelling Systems: Evolution sincethe Early Miocene. Geological Society SpecialPublication 64, pp 273-284.

Weber, A., and Jørgensen, B.B., 2002. Bacterial sulfatereduction in hydrothermal sediments of the GuaymasBasin, Gulf of California, Mexico. Deep-Sea ResearchI, 49: 827-841.

Werne, J.P., Hollander, D.J., Lyons, T.W. and SinningheDamsté, J.S., 2004. Organic sulfur biogeochemistry:Recent advances and future research directions. In:Amend, J.P., Edwards, K.J. and Lyons, T.W. (eds),Sulfur Biogeochemistry – Past and Present. Geo-logical Society of America Special Paper 379, pp.135-150.

Westrich, J.T., and Berner, R.A., 1984. The role ofsedimentary organic matter in bacterial sulfate reduc-tion: The G model tested. Limnology and Oceano-graphy, 29: 236-249.

Whiticar, M.J., 1999. Carbon and hydrogen isotopesystematics of bacterial formation and oxidation ofmethane. Chemical Geology, 161: 291-314.

Whitman, W.B., Bowen, T.L., and Boone, D.R., 1999.The methanogenic bacteria. In: Dworkin, M., Balows,A., Trüper, H.G., Harder, W., and Schleifer, K.-H.(eds), The Prokaryotes, 3rd. Ed. Springer, New York.

Widdel, F., 1988. Microbiology and ecology of sulfate-and sulfur-reduction bacteria. In: Zehnder, A.J.B. (ed),Biology of anaerobic microorganisms. Wiley & Sons,NY, pp. 469-585.

Widdel, F. and Hansen, T.A., 1991. The dissimilatorysulfate- and sulfur-reducing bacteria. In: Balows, H. etal. (eds), The Procaryotes. Springer, pp. 583-624.

Wilkin, R.T. and Barnes, H.L., 1996. Pyrite formationby reactions of iron monosulfides with dissolved

Page 40: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

309

inorganic and organic sulfur species. Geochimica etCosmochimica Acta, 60: 4167-4179.

Wollast, R., 1998. Evaluation and comparison of theglobal carbon cycle in the coastal zone and in theopen ocean. In: Brink, K.H., and Robinson, A.R.(eds), The Sea, Vol. 10, pp. 213-252.

References

Page 41: Sulfur Cycling and Methane Oxidation · For an overview of the processes and pathways involved in the incorporation of sulfur into organic matter, we refer to Orr and White (1990),

8 Sulfur Cycling and Methane Oxidation

310