Iron-Mediated Anaerobic Oxidation of Methane in Brackish Coastal

Nov 20, 2014 - Department of Earth Sciences - Geochemistry, Faculty of Geosciences ... Institute for Marine and Atmospheric Research Utrecht (IMAU), U...
0 downloads 0 Views 5MB Size
Subscriber access provided by UNIV OF UTAH

Article

Iron-mediated anaerobic oxidation of methane in brackish coastal sediments Matthias Egger, Olivia Rasigraf, Célia J. Sapart, Tom Jilbert, Mike S. M. Jetten, Thomas Röckmann, Carina van der Veen, Narcisa Banda, Boran Kartal, Katharina F. Ettwig, and Caroline P. Slomp Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/es503663z • Publication Date (Web): 20 Nov 2014 Downloaded from http://pubs.acs.org on November 22, 2014

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Environmental Science & Technology

Feox

ca tio hi tro p

Feox

Feox SO4

2-

Eu

SO42-

CH4

CH4

Fe2+

CH4 + Feox

4

SO

FeS

Feox

CH4 + Feox

-A OM

Feox

H2S + Feox

ACS Paragon Plus Environment

CH4

CH4

SO4

FeS

-A OM

H2S + Feox

n

SO42-

Feox

2-

4

Feox Feox

CH4

SO4

2-

SO

SO4

2-

Feox Fe2+

Fe2+

CH4

Fe -A OM

Page 1 of 27

Environmental Science & Technology

Page 2 of 27

1

Iron-mediated anaerobic oxidation of methane in

2

brackish coastal sediments

3

Matthias Egger a*, Olivia Rasigraf b, Célia J. Sapart

4

Thomas Röckmann c, Carina van der Veen c, Narcisa Bândă c, Boran Kartal b,f, Katharina F.

5

Ettwig b, Caroline P. Slomp a

6

a

7

Budapestlaan 4, 3584 CD Utrecht, The Netherlands; b Department of Microbiology, Institute

8

for Water and Wetland Research, Faculty of Science, Radboud University Nijmegen,

9

Heyendaalseweg 135, 6525 AJ Nijmegen, The Netherlands;

c,d

, Tom Jilbert

a,e

, Mike S. M. Jetten b,

Department of Earth Sciences - Geochemistry, Faculty of Geosciences, Utrecht University,

c

Institute for Marine and

10

Atmospheric research Utrecht (IMAU), Utrecht University, Princetonplein 5, 3584 CC

11

Utrecht, The Netherlands;

12

Avenue F. D. Roosevelt, B-1050 Bruxelles, Belgium; e Now at: Department of Environmental

13

Sciences, University of Helsinki, Viikinkaari 2a, 00014 Helsinki, Finland;

14

Biochemistry and Microbiology, Laboratory of Microbiology, Ghent University, K. L.

15

Ledeganckstraat 35, 9000 Gent, Belgium

d

Laboratoire de Glaciologie, Université Libre de Bruxelles, 50

f

Department of

16

17

KEYWORDS

18

Anaerobic oxidation of methane; iron biogeochemistry; coastal sediments; eutrophication

19

1 ACS Paragon Plus Environment

Page 3 of 27

Environmental Science & Technology

20

ABSTRACT

21

Methane is a powerful greenhouse gas and its biological conversion in marine sediments,

22

largely controlled by anaerobic oxidation of methane (AOM), is a crucial part of the global

23

carbon cycle. However, little is known about the role of iron oxides as an oxidant for AOM.

24

Here we provide the first field evidence for iron-dependent AOM in brackish coastal surface

25

sediments and show that methane produced in Bothnian Sea sediments is oxidized in distinct

26

zones of iron- and sulfate-dependent AOM. At our study site, anthropogenic eutrophication

27

over recent decades has led to an upward migration of the sulfate/methane transition zone in

28

the sediment. Abundant iron oxides and high dissolved ferrous iron indicate iron reduction in

29

the methanogenic sediments below the newly established sulfate/methane transition.

30

Laboratory incubation studies of these sediments strongly suggest that the in-situ microbial

31

community is capable of linking methane oxidation to iron oxide reduction. Eutrophication of

32

coastal environments may therefore create geochemical conditions favorable for iron-

33

mediated AOM and thus increase the relevance of iron-dependent methane oxidation in the

34

future. Besides its role in mitigating methane emissions, iron-dependent AOM strongly

35

impacts sedimentary iron cycling and related biogeochemical processes through the reduction

36

of large quantities of iron oxides.

37

38

39

40

41

42 2 ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 27

43

INTRODUCTION

44

Methane (CH4) is a potent greenhouse gas in the Earth’s atmosphere with a global warming

45

potential 28 - 34 times that of carbon dioxide (CO2) on a centennial timescale 1. Anaerobic

46

oxidation of methane (AOM) efficiently controls the atmospheric CH4 efflux from the ocean

47

by consuming an estimated > 90 % of all the CH4 produced in marine sediments

48

this AOM is attributed to sulfate (SO42-) reduction

49

oxidized solid-phases such as iron (Fe)-oxides are thermodynamically favorable electron

50

acceptors for the biological oxidation of CH4 (equation (2)).

51

CH4 + SO42- → HS- + HCO3- + H2O

52

CH4 + 8Fe(OH)3 + 15H+ → HCO3- + 8Fe2+ + 21H2O

53

The co-occurrence of reactive Fe-oxides and CH4 in freshwater

54

sedimentary environments suggests that AOM coupled to Fe-oxide reduction (Fe-AOM) has

55

the potential to act as a CH4 removal mechanism in SO42--poor systems. Laboratory

56

incubation studies have indeed shown that AOM can be stimulated by Fe-oxide additions 15,17–

57

19

58

knowledge about its significance for global CH4 dynamics are still lacking.

59

A basic prerequisite for Fe-mediated AOM is the concurrent presence of porewater CH4 and

60

abundant reducible Fe-oxides. However, in most brackish and marine sediments the presence

61

of porewater SO42- will stimulate microbial SO42- reduction and generate dissolved sulfide.

62

This typically results in the reductive dissolution of sedimentary Fe-oxides and the conversion

63

of most reducible Fe to authigenic Fe-sulfides. Significant amounts of Fe-oxides below the

64

zone of SO42- reduction in brackish and marine sediments, therefore, are likely mainly found

65

in systems where the input of Fe-oxides is high or where the sediments are subject to transient

66

diagenesis. Examples of the latter are found in the Black Sea

4–10

2,3

. Most of

(SO4-AOM, equation (1)), but also

(1) (2) 11–14

and some brackish

15,16

. However, conclusive field evidence for Fe-AOM in brackish surface sediments and

20

and Baltic Sea 21, where the 3

ACS Paragon Plus Environment

Page 5 of 27

Environmental Science & Technology

67

transition from a freshwater lake to a marine system has led to the preservation of Fe-oxides

68

below sulfidic sediment layers. Other examples of non-steady state depositional marine

69

systems where conditions are likely favorable for Fe-AOM include the continental margin off

70

Argentina

71

flows contribute to rapid burial of reactive Fe (see

72

subsurface sediments with potentially Fe-AOM facilitating conditions). Burial of Fe-oxides

73

below the zone of SO42- reduction is also enhanced when Fe-oxides are relatively resistant to

74

reduction, e.g. because of their crystalline nature or changes in surface structure due to

75

adsorption of ions

76

diagenesis is anthropogenic fertilization of coastal environments.

77

Here, we provide strong geochemical evidence for Fe-dependent CH4 oxidation below a

78

shallow sulfate/methane transition zone (SMTZ) in brackish coastal sediments from the

79

Bothnian Sea (salinity 5-6; Figure 1). Furthermore, we show that anthropogenic

80

eutrophication may induce non-steady state diagenesis, favoring a coupling between Fe-oxide

81

reduction and CH4 oxidation in coastal sediments. Besides acting as a CH4 sink, Fe-AOM

82

greatly impacts the biogeochemical cycling of Fe and thereby other tightly coupled element

83

cycles in marine and brackish environments.

84

MATERIALS AND METHODS

85

Sediment and porewater sampling. Sediment cores (~ 0-60 cm) were collected from site

86

US5B (62°35.17’ N, 19°58.13’ E) using a GEMAX corer (core diameter 10 cm) during a

87

cruise with R/V Aranda in August 2012. All cores were sliced under nitrogen at 1-4 cm

88

resolution. For each slice a sub-sample was placed in a pre-weighed glass vial for solid-phase

89

analysis and stored anoxically at -20 °C. Porewater was extracted by centrifugation, filtered

90

through 0.45 µm pore size disposable filters and sub-sampled under nitrogen for analysis of

22,23

and the Zambezi deep-sea fan sediments

24

23

where turbidites and other mass for further examples of marine

25,26

. An additional, but still poorly investigated trigger for transient

4 ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 27

91

dissolved Fe, SO42- and sulfide (∑H2S = H2S + HS- + S2-). Sub-samples for total dissolved Fe,

92

which we assume to be present as Fe2+, were acidified with 10 µl 35 % suprapur HCl per ml

93

of sub-sample and stored at 4 °C until analysis by ICP-OES (Perkin Elmer Optima 3000

94

Inductively Coupled Plasma - Optimal Emission Spectroscopy). Porewater SO42- was

95

analyzed with ion chromatography (IC) (detection limit of < 75 µmol/l) and compared well

96

with total sulfur (S) measured by ICP-OES. Another sub-sample of 0.5 ml was immediately

97

transferred into glass vials (4 ml) containing 2 ml of 2 % zinc (Zn)-acetate solution to

98

precipitate ZnS, and was stored at 4 °C. Sulfide concentrations were then determined

99

spectrophotometrically by complexion of the ZnS precipitate in an acidified solution of 27

100

phenylenediamine and ferric chloride

(detection limit of < 1 µmol/l). The sulfide standard

101

was validated by titration with thiosulfate to achieve improved accuracy.

102

CH4 samples were taken from a pre-drilled core directly upon core retrieval. Precisely 10 ml

103

wet sediment was extracted from each hole and immediately transferred into a 65 ml glass

104

bottle filled with saturated NaCl solution, sealed with a rubber stopper and a screw cap and

105

subsequently stored upside-down. A headspace of 10 ml nitrogen was injected and CH4

106

concentrations in the headspace were determined by injection of a subsample into a Thermo

107

Finnigan Trace GC gas chromatograph (Flame Ionization Detector). δ13C-CH4 and δD-CH4

108

(D, deuterium) were analyzed by a Continuous Flow Isotope Ratio Mass Spectrometry (CF-

109

IRMS) system at the Institute for Marine and Atmospheric Research Utrecht (IMAU). The

110

amount of CH4 per sample was often larger than the calibrated range of the isotope ratio mass

111

spectrometer. Therefore, very small amounts of sample were injected via a low-pressure inlet

112

and diluted in helium (He BIP 5.7, Air Products) in a 40 ml sample loop to reach CH4 mixing

113

ratios of about 2000 ppb. After dilution, the samples were measured as described in detail in

114

28

and 29.

5 ACS Paragon Plus Environment

Page 7 of 27

Environmental Science & Technology

115

Sediment samples were freeze-dried, powdered and ground in an agate mortar inside an

116

argon-filled glove box. Sediment porosity was determined from the weight loss upon freeze-

117

drying. A split of about 125 mg of freeze-dried sediment was dissolved in a 2.5 ml HF (40 %)

118

and 2.5 ml of HClO4/HNO3 mixture, in a closed Teflon bomb at 90 °C during one night. The

119

acids were then evaporated at 160 °C and the resulting gel was subsequently dissolved in 1 M

120

HNO3 at 90 °C during another night. Finally, total elemental concentrations were determined

121

by ICP-OES (precision and accuracy < 5 %, based on calibration to standard solutions and

122

checked against internal laboratory standard sediments). To correct for variations in

123

terrigenous input of non-reactive Fe, total Fe concentrations were normalized to aluminum

124

(Al), with the input of the latter element assumed to be exclusively from terrestrial sources.

125

To investigate the solid-phase partitioning of Fe, a second 50 mg aliquot of dried sediment

126

was subjected to a sequential extraction procedure for Fe after

127

fractionated into I) carbonate associated Fe (including siderite and ankerite, extracted by 1 M

128

Na-acetate brought to pH 4.5 with acetic acid, 24 h), II) easily reducible (amorphous) oxides

129

(Fe(ox1), including ferrihydrite and lepidocrocite, extracted by 1 M hydroxylamine-HCl, 48

130

h), III) reducible (crystalline) oxides (Fe(ox2), including goethite, hematite and akagenéite,

131

extracted by Na-dithionite buffer, pH 4.8, 2 h) and IV) Fe in recalcitrant oxides (mostly

132

magnetite, extracted by 0.2 M ammonium oxalate / 0.17 M oxalic acid solution, 2 h).

133

Replicate analyses had a relative error of < 5 %. A third 0.5 g aliquot of dried sediment was

134

used to determine the amount of FeS2 (chromium reducible sulfur, CRS, using acidic

135

chromous chloride solution) and FeS (acid volatile sulfur, AVS, using 6 M HCl) via the

136

diffusion-based approach described by 31 and analyzed by iodometric titration of the alkaline

137

Zn acetate trap. Replicate sample extractions yielded reproducibility within 8 % for CRS and

138

10 % for AVS. Results for the Fe(ox1) fraction suggest that most of the AVS present in the

139

SMTZ was extracted during the hydroxylamine-HCl step (SI Figure S1). Thus, the estimation

30

. Sediment Fe was

6 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 27

140

of the reducible sedimentary Fe-oxide content (Fe(ox1) + Fe(ox2)) shown in Figure 2 was

141

corrected for AVS dissolution within the SMTZ.

142

Incubation experiments. Replicate sediment cores taken in 2012 were sealed immediately

143

and stored onboard at 4 °C. Back in the lab, they were sliced in sections of 5 cm under strictly

144

anaerobic conditions and stored anaerobically in the dark at 4 °C. The incubation experiments

145

started within half a year after sampling. Per culture bottle, 30 g of wet sediment (~ 25 ml)

146

was homogenized in 75 ml SO42--depleted medium mimicking in-situ bottom water conditions

147

within a helium-filled glove box (containing ~ 2 % hydrogen gas, H2). 85 ml of this

148

sediment/medium slurry was distributed over 150 ml culture bottles and an additional 11.4 ml

149

of medium was added to all bottles with the exception of the Fe treatments, where 11.4 ml of

150

a Fe-nanoparticle solution

151

was added instead. The approximate ratio of 1 part sediment to 3 parts medium was chosen in

152

agreement to reported incubation studies

153

with airtight red butyl rubber stoppers and secured with open-top Al screw caps. After being

154

sealed, 5 ml CO2 and either 45 ml nitrogen (“cntl”) or 45 ml 13CH4 (“13CH4” and “13CH4 &

155

Fe(III)”) were injected into the headspace of duplicate incubations to yield 1 bar overpressure

156

(volume headspace = 50 ml) and incubated in the dark at 20 °C under gentle shaking.

157

Dissolved sulfide and SO42- was sampled within the glove box by allowing the sediment to

158

settle out of suspension and taking a sub-sample (1.5 ml) of the supernatant water via a needle

159

syringe. Analysis of sulfide was performed as described above for sediment porewater

160

(detection limit of < 1 µmol/l). Samples for total dissolved S were measured by ICP-OES

161

after acidification with 10 µl 35 % suprapur HCl and assumed to represent only SO42- due to

162

the release of sulfide to the gas phase during acidification 33 (detection limit of < 82 µmol/l).

163

Headspace samples (30 µl) were analyzed by gas - chromatography (GC, Agilent 6890 series,

164

USA) using a Porapak Q column at 80 °C (5 min) with helium as the carrier gas (flow rate 24

32

(20 mmol/l ferric Fe (Fe(III)), resulting in 2 mmol/l per bottle)

15,18

. After mixing, the culture bottles were sealed

7 ACS Paragon Plus Environment

Page 9 of 27

Environmental Science & Technology

165

ml/min). The GC was coupled to a mass spectrometer (Agilent 5975C inert MSD, Agilent,

166

USA) to quantify the masses 44 and 45 (CO2). To account for the medium loss due to sub-

167

sampling of the solution and because of an observed leveling-off of measured headspace

168

13

169

nanoparticle solution (“13CH4 & Fe(III)”) was added to the culture bottles after 55 days

170

(indicated by a dashed line in Figure 3 and SI Figure S2). Fe-AOM rates were determined

171

from the linear slope of 13CO2 production in duplicate incubations from 20-30 cm and 30-35

172

cm depth, before and after second addition of Fe(III) (SI Figure S2), taking into account the

173

CO2 dissolved in the liquid and removal thereof during sampling. Thus, the statistical mean is

174

based on a total number of eight rate estimates.

175

Diffusion model simulations. A column diffusion model was used to test different

176

hypotheses for sources and sinks of CH4 within the sediment at station US5B. The model

177

represents the sediment column from the surface down to a depth of 36 cm, using 72 layers of

178

0.5 cm each. Concentrations of total CH4, δD-CH4 and δ13C-CH4 are simulated, accounting

179

for production, loss and diffusion, and their corresponding isotopic fractionation. The model

180

was run to steady state for each scenario. As the boundary condition at 36 cm depth we

181

constrain the model with the measured values of CH4, δD-CH4 and δ13C-CH4. At the surface,

182

the flux to the water column was set to zero. The diffusion coefficient for CH4 was corrected

183

for in-situ temperature, salinity, and porosity (see Supplementary Information). A diffusion

184

fractionation of εDiffusion = 3 ‰ for both 13C-CH4 and D-CH4 was assumed according to 34,35.

185

Two zones were defined within the sediment column, between 0-8.5 cm and 8.5-36 cm, with

186

constant production and loss parameters within each zone. The first zone represents SO4-

187

AOM, while the second represents the area where both CH4 loss through Fe-AOM and CH4

188

production through methanogenesis may occur. We performed three model simulations: ‘With

189

Fe-AOM’, ‘No Fe-AOM’ and ‘Diffusion only’. The rate constants and fractionation factors

CO2 concentrations, an additional 11.4 ml of medium (“cntl” and “13CH4”) and Fe-

8 ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 27

190

used in these simulations for the CH4 production and removal processes are presented in SI

191

Table S3.

192

RESULTS AND DISCUSSION

193

Geochemical profiles. Vertical porewater profiles for station US5B (Figure 2) reveal a

194

shallow SMTZ at a depth of ca. 4 to 8.5 cm, where SO4-dependent AOM results in the

195

depletion of porewater SO42- and CH4. Dissolved SO42- decreases from 4.9 mmol/l at the

196

sediment water interface to 0.3 mmol/l at the bottom of the SMTZ, and stayed below the

197

detection limit of 75 µmol/l throughout the rest of the core. Reductive dissolution of Fe-

198

oxides driven by sulfide production during SO4-AOM induces a distinct minimum in

199

sedimentary Fe-oxides and precipitation of Fe-sulfides (mostly FeS). Abundant reducible Fe-

200

oxides below the SMTZ are accompanied by very high dissolved ferrous Fe (Fe2+)

201

concentrations (> 1.8 mmol/l). The depth trend in total (Fe/Al) indicates that the Fe is not only

202

repartitioned between oxide and sulfide phases within the SMTZ, but that Fe-oxide reduction

203

below the SMTZ triggers upward migration of dissolved Fe2+ with an enrichment of total Fe

204

in the sulfidic zone. However, in contrast to the SMTZs observed at greater depth (> 1.5 m) in

205

sediments of the Baltic Sea

206

sequestered in the form of authigenic Fe-sulfides in the shallow SMTZ observed in this study

207

(Figure 2). Thus, no sulfide remains to diffuse downwards into the zone where Fe reduction is

208

occurring. Reductive Fe-oxide dissolution by dissolved sulfide in the deeper sediments is

209

therefore not possible. Although a cryptic sulfur cycle where SO42- is generated by sulfide

210

reacting with deeply buried Fe(III) species

211

measurements of SO42- reduction rates would be needed to completely exclude the possibility

212

of a cryptic sulfur cycle through re-oxidation of Fe sulfide minerals. High dissolved Fe2+

213

concentrations

214

disproportionation of elemental sulfur

further

21

and Black Sea

preclude

Fe-oxide

20

, sulfide generated by SO42- reduction is all

20,21

is unlikely to occur,

reduction

via

sulfide

35

S radiotracer

released

during

36,37

, as any sulfide produced locally would be 9

ACS Paragon Plus Environment

Page 11 of 27

Environmental Science & Technology

215

immediately scavenged to form Fe-sulfides. Thus, there are only two alternative mechanisms

216

that may explain the high dissolved Fe2+ concentration in the porewater below the SMTZ. The

217

first mechanism is organoclastic Fe reduction, i.e. Fe reduction coupled to organic matter

218

degradation. The second is Fe-AOM, i.e. Fe reduction coupled to AOM. In the following, we

219

demonstrate why the latter mechanism is the process most likely responsible for the high

220

dissolved Fe2+ in these sediments.

221

Fe-AOM activity below the SMTZ. The potential of the microbial community present in the

222

sediments below the SMTZ to perform Fe-AOM was experimentally tested with slurry

223

incubation studies. Fresh sediment samples from several depth layers were incubated with

224

13

225

water conditions. Duplicate incubations were amended with either only

226

20 mmol/l Fe hydroxide nanoparticles 32. In the control slurries, nitrogen was used instead of

227

13

228

product of 13CH4 oxidation. The addition of Fe hydroxide nanoparticles almost doubled AOM

229

activity (1.7 fold increase) in the sediment samples between 20 and 35 cm depth compared to

230

the slurries where no additional Fe(III) was added (Figure 3). The increase in δ13C-CO2 as a

231

response to Fe(III) addition thus suggest that the microbial community present in the sediment

232

below the SMTZ is capable of coupling AOM to Fe reduction. Throughout the whole

233

experiment, sulfide stayed below detection limit (< 1 µmol/l) and SO42- concentrations stayed

234

below 350 µmol/l. A slight decrease in background SO42- during the incubation period might

235

indicate low levels of SO42- reduction of ~ 1.9 pmol SO42- cm-3 day-1. Similar rates of SO42-

236

reduction (~ 1 pmol SO42- cm-3 day-1) were reported for methanogenic Baltic Sea sediments 21.

237

Taking into account indirect Fe stimulated SO42--driven AOM 19 through a cryptic sulfur cycle

238

20,21

239

to SO42- in a 17:1 stoichiometric ratio, we estimate a gross rate of SO42- reduction of ~ 2 pmol

C-labeled CH4 (13CH4), CO2 and a SO42--depleted medium mimicking Bothnian Sea bottom

CH4. AOM rates were then determined by measuring production of

13

CH4 or

13

CH4 and

13

CO2, i.e. the end

, where redox reactions between sulfide and Fe-oxides result in the re-oxidation of sulfide

10 ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 27

240

SO42- cm-3 day-1. SO4-AOM is thus unlikely to contribute more than 0.1 % to the total 13CO2

241

production observed in our incubations. These findings support our hypothesis that the

242

accumulation of dissolved Fe2+ in the porewater below the SMTZ is, at least partly, a result of

243

Fe-AOM. The potential rate of Fe-AOM in our incubations is 1.32 ± 0.09 µmol cm-3 year-1

244

(see SI Figure S2), which compares well to recent estimates of potential Fe-AOM rates in

245

slurry incubations of brackish wetland (1.42 ± 0.11 µmol cm-3 year-1; 15) and Fe(III)-amended

246

mesocosm studies of intact deep lake sediment cores (1.26 ± 0.63 µmol cm-3 year-1;

247

should be noted, however, that these rates all derive from stimulated microbial communities

248

and thus could be lower under in situ conditions.

249

Isotopic evidence for concurrent methanogenesis. Porewater profiles of δD-CH4 and δ13C-

250

CH4 (Figure 4a) show the common features observed in marine sediments where

251

methanogenesis deep in the sediment results in the build up of isotopically depleted CH4 in

252

the porewater 38. The preferential oxidation of isotopically light CH4 during AOM 39,40 results

253

in a progressive enrichment of the residual CH4 in 13C-CH4 and D-CH4 towards the sediment

254

surface. Application of the Rayleigh distillation function

255

kinetic isotope fractionation factors for hydrogen (εH) and carbon (εC) that are spatially

256

separated (Figure 4b). The sympathetic change in the

257

depth is driven by SO4-AOM within the SMTZ. Corresponding isotope fractionations (εH = 98

258

± 0.3 ‰, εC = 9 ± 0.2 ‰, see Figure 4b) are at the low end of reported values from near

259

coastal sediments in the Baltic Sea (εH = 100-140 ‰, εC = 11-13 ‰)40 and SO4-AOM in

260

general

261

enrichment in D isotopes but depletion in 13C (εH = 24 ± 0.1 ‰, εC = -1 ± 0.04 ‰, see Figure

262

4b). Such antipathetic changes are usually attributed to CH4 production

263

hydrogenotrophic methanogenesis (CO2 reduction by H2) as the most likely methanogenic

264

pathway for the range of CH4 isotope ratios observed at station US5B 38.

12,38,41

13

). It

reveals two distinct sets of

13

C and D content from 4 to 8.5 cm

38,39,42

. The isotopic composition of CH4 below the SMTZ, on the other hand, shows

38,43

, with

11 ACS Paragon Plus Environment

Page 13 of 27

Environmental Science & Technology

265

Fe-AOM as the main contributor to Fe2+ production and CH4 removal would be expected to

266

isotopically enrich porewater CH4 in both heavy isotopes below the SMTZ, in contrast to the

267

observations. We hypothesized that production of isotopically light CH4 during

268

methanogenesis counteracts the progressive enrichment of isotopically heavy CH4 during Fe-

269

AOM. Indeed, simulations of the isotopic composition of porewater CH4 at station US5B

270

using a column diffusion model could successfully reproduce the observed patterns in δD-

271

CH4 and δ13C-CH4 when assuming the concurrent presence of Fe-AOM and hydrogenotrophic

272

methanogenesis (Figure 5). Due to the slow rate of CH4 oxidation, Fe-AOM only removes a

273

small amount of porewater CH4. Hence, modeled profiles of δD-CH4 and δ13C-CH4 are highly

274

sensitive to the rate of methanogenesis and the assumed values of εH and εC, which have a

275

wide reported range

276

approximated without Fe-AOM, when altered CH4 production rates and methanogenic

277

fractionation factors are used in the model. Nevertheless, the model does demonstrate that

278

methanogenesis is required to reproduce the porewater concentration and isotope profiles of

279

CH4; removal in the SMTZ only and diffusion from depth cannot explain the observed trends.

280

This provides conclusive evidence for methanogenesis within the zone of Fe reduction.

281

However, hydrogenotrophic Fe-reducers coupling H2 oxidation to Fe(III) reduction are known

282

to be able to outcompete methanogens for H2 by maintaining the H2 concentrations at levels

283

that are too low for methanogens to grow

284

methanogenesis by Fe(III) has been reported 47–49. The high demand for H2 needed to explain

285

the dissolved Fe2+ concentrations (2:1 stoichiometry)

286

competitive inhibition of hydrogenotrophic methanogenesis. Here we clearly demonstrate the

287

co-occurrence of Fe reduction and methanogenesis, indicating that reduction of Fe(III) by H2

288

is unlikely to be the main reason for the high dissolved Fe2+ below the SMTZ. Consequently,

289

we conclude that CH4 might be a plausible electron donor for the reduction of relatively

38

. Our simulations show that the measurements can also be

44–46

. In addition, direct inhibition of

44,50

could therefore result in a

12 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 27

290

refractory Fe-oxides under methanogenic conditions. Our field data and model results further

291

imply that the imprint of Fe-AOM on the isotopic composition of porewater CH4 may be

292

masked when co-occurring with methanogenesis.

293

Fe-AOM and coastal eutrophication. In the Bothnian Sea, the shallow position of the

294

SMTZ is linked to anthropogenic eutrophication over recent decades. Long-term monitoring

295

studies based on water transparency, inorganic nutrients and chlorophyll a in the Bothnian Sea

296

51,52

297

land has enhanced primary productivity and subsequent export production of organic matter.

298

As a consequence, rates of SO42- reduction and methanogenesis likely increased, initiating an

299

upward migration of the SMTZ. Since the early 2000s, eutrophication of the Bothnian Sea has

300

started to decline again

301

position. Evidence for a rapid upward shift of the SMTZ and its recent fixation is provided by

302

the distinct enrichment of authigenic Fe-sulfides around the present day SMTZ

303

The diffusive influx of SO42- into the SMTZ (~ 1 mol m-2 year-1; Figure 2) and the S content

304

of the Fe-sulfide enrichment (~ 10 mol m-2; Figure 2) suggest that the fixation of the SMTZ in

305

its current position occurred approximately 10 years before the time of sampling, consistent

306

with earlier findings at site US5B

307

response to eutrophication likely reduced the exposure time of sedimentary Fe-oxides to

308

dissolved sulfide, allowing a portion of Fe-oxides to remain preserved below the newly

309

established SMTZ and facilitating Fe-AOM. Since many coastal environments have

310

experienced eutrophication in recent decades

311

may be widespread, creating similar zones of Fe-AOM in Fe-oxide rich brackish coastal

312

sediments worldwide.

313

Biogeochemical implications. Although Fe-oxides are abundant in the sediments below the

314

SMTZ, their bioavailability may be limiting Fe-dependent AOM. This is implied by the

indicate that, throughout the last century, anthropogenic loading of nutrients derived from

51,52

, halting the upward movement of the SMTZ at its current

16,22,24,53,54

.

16

. This rapid upward displacement of the SMTZ in

55

, we postulate that such transient diagenesis

13 ACS Paragon Plus Environment

Page 15 of 27

Environmental Science & Technology

315

coexistence of both CH4 and Fe-oxides in the same zone and the clear stimulation of CH4

316

oxidation after addition of Fe nanoparticles. Our incubation and modeling results suggest that

317

Fe-AOM accounts for ~ 3 % of total CH4 removal in Bothnian Sea sediments (see

318

Supplementary Information for calculations), contributing strongly to Fe2+ production due to

319

the 8:1 Fe-CH4 stoichiometry (equation (2)). The remaining CH4 is likely removed by SO4-

320

AOM and aerobic CH4 oxidation at the sediment surface. The large amount of reactive Fe-

321

oxides needed to oxidize significant amounts of CH4 and the low measured reaction rates

322

(Figure 3), imply that Fe-AOM probably does not contribute more than a few percent to CH4

323

oxidation in modern marine and brackish surface sediments. However, Fe-AOM could

324

significantly impact the biogeochemical cycles of other elements. For example, high

325

porewater Fe2+ may be especially conducive to in-situ formation of ferrous-phosphate

326

minerals, thus increasing the sequestration of phosphorus in the sediment and providing a

327

negative feedback on coastal eutrophication

328

aquatic sediments, future climate change and eutrophication may increase the global

329

importance of Fe-AOM and its impact on other biogeochemical cycles.

330

ACKNOWLEDGMENTS

331

We thank the captain, crew and scientific participants aboard R/V Aranda in 2012 and P.

332

Kotilainen for their assistance with the fieldwork and D. van de Meent, H. de Waard and T.

333

Zalm for analytical assistance in Utrecht. This work was funded by ERC Starting Grant

334

278364 (to C. P. S.). O. R. was supported by ERC Advanced Grant 2322937. M. S. M. J. was

335

supported by ERC Advanced Grant 339880 and OCW/NWO Gravitation Grant SIAM

336

024002002. K. F. E. was supported by the Darwin Center for Biogeology (project 3071) and

337

VENI grant 863.13.007 from the Dutch Science Foundation (NWO), and B. K. by VENI grant

338

863.11.003 by NWO.

16

. By inducing non-steady state diagenesis in

14 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 27

339

340

ASSOCIATED CONTENT

341

Supporting Information Available. Description of porewater calculations and model details,

342

correction of AVS dissolution during Fe oxide extraction (Figure S1), determination of Fe-

343

AOM rate (Figure S2), measured porewater (Table S1) and solid phase concentrations (Table

344

S2), and model parameters used for the different simulations (Table S3). This information is

345

available free of charge via the Internet at http://pubs.acs.org/.

346

AUTHOR INFORMATION

347

Corresponding Author. *Department of Earth Sciences - Geochemistry, Faculty of

348

Geosciences, Utrecht University, Budapestlaan 4, 3584 CD Utrecht, The Netherlands, Phone:

349

+31 30 253 3264, Email: [email protected]

350

Authors Contributions. C. P. S. conceived and designed the field study. M. E. and T. J.

351

carried out the fieldwork. M. E., C. P. S., O. R., K. F. E., B. K., and M. S. M. J. designed

352

incubation experiments. M. E., O. R. and K. F. E. performed and analyzed experiments. C. J.

353

S., C.v.d. V. and T. R. performed methane isotope analysis. M. E. analyzed fieldwork samples

354

and compiled all data. N. B. designed the diffusion model applied for methane isotopes. M. E.

355

and C. P. S. wrote the manuscript with contributions of all co-authors.

356

REFERENCES

357 358 359 360 361

(1)

Myhre, G.; Schindell, D.; Bréon, F. M.; Collins, W.; Fuglestvedt, J.; Huang, J.; Koch, D.; Lamarque, J. F.; Lee, D.; Mendoza, B.; et al. Anthropogenic and natural radiative forcing. In Climate Change 2013: The physical science basis.; Stocker, T. F.; Qin, D.; Plattner, G. K.; Tignor, M.; Allen, S. K.; Boschung, J.; Nauels, A.; Xia, Y.; Bex, V.; Midgley, P. M., Eds.; Cambridge University Press: Cambridge United Kingdom and New York, NY, USA, 2013.

362

(2)

Reeburgh, W. S. Oceanic methane biogeochemistry. Chem. Rev. 2007, 107, 486–513.

363 364

(3)

Knittel, K.; Boetius, A. Anaerobic oxidation of methane: progress with an unknown process. Annu. Rev. Microbiol. 2009, 63, 311–334. 15 ACS Paragon Plus Environment

Page 17 of 27

Environmental Science & Technology

365 366

(4)

Barnes, R. O.; Goldberg, E. D. Methane production and consumption in anoxic marine sediments. Geology 1976, 4, 297–300.

367 368

(5)

Reeburgh, W. S. Methane consumption in Cariaco Trench waters and sediments. Earth Planet. Sci. Lett. 1976, 28, 337–344.

369 370 371

(6)

Martens, C. S.; Berner, R. A.; Rosenfeld, J. K. Interstitial water chemistry of anoxic Long Island Sound sediments. 2. Nutrient regeneration and phosphate removal. Limnol. Oceanogr. 1978, 23, 605–617.

372 373 374

(7)

Iversen, N.; Jørgensen, B. B. Anaerobic methane oxidation rates at the sulfate-methane transition in marine sediments from Kattegat and Skagerrak (Denmark). Limnol. Oceanogr. 1985, 30, 944–955.

375 376 377

(8)

Hoehler, T. M.; Alperin, M. J.; Albert, D. B.; Martens, C. S. Field and laboratory studies of methane oxidation in an anoxic marine sediment: Evidence for a methanogen-sulfate reducer consortium. Global Biogeochem. Cycles 1994, 8, 451–463.

378 379 380

(9)

Boetius, A.; Ravenschlag, K.; Schubert, C. J.; Rickert, D.; Widdel, F.; Gieseke, A.; Amann, R.; Jørgensen, B. B.; Witte, U.; Pfannkuche, O. A marine microbial consortium apparently mediating anaerobic oxidation of methane. Nature 2000, 407, 623–626.

381 382 383

(10)

Milucka, J.; Ferdelman, T. G.; Polerecky, L.; Franzke, D.; Wegener, G.; Schmid, M.; Lieberwirth, I.; Wagner, M.; Widdel, F.; Kuypers, M. M. M. Zero-valent sulphur is a key intermediate in marine methane oxidation. Nature 2012, 491, 541–546.

384 385

(11)

Wersin, P.; Hoehener, P.; Giovanoli, R.; Stumm, W. Early diagenetic influences on iron transformations in a freshwater lake sediment. Chem. Geol. 1991, 90, 233–252.

386 387 388

(12)

Crowe, S. A.; Katsev, S.; Leslie, K.; Sturm, A.; Magen, C.; Nomosatryo, S.; Pack, M. A.; Kessler, J. D.; Reeburgh, W. S.; Roberts, J. A.; et al. The methane cycle in ferruginous Lake Matano. Geobiology 2011, 9, 61–78.

389 390 391

(13)

Sivan, O.; Adler, M.; Pearson, A.; Gelman, F.; Bar-Or, I.; John, S. G.; Eckert, W. Geochemical evidence for iron-mediated anaerobic oxidation of methane. Limnol. Oceanogr. 2011, 56, 1536–1544.

392 393

(14)

Nordi, K. Á.; Thamdrup, B.; Schubert, C. J. Anaerobic oxidation of methane in an iron-rich Danish freshwater lake sediment. Limnol. Oceanogr. 2013, 58, 546–554.

394 395 396

(15)

Segarra, K. E. A.; Comerford, C.; Slaughter, J.; Joye, S. B. Impact of electron acceptor availability on the anaerobic oxidation of methane in coastal freshwater and brackish wetland sediments. Geochim. Cosmochim. Acta 2013, 115, 15–30.

397 398 399

(16)

Slomp, C. P.; Mort, H. P.; Jilbert, T.; Reed, D. C.; Gustafsson, B. G.; Wolthers, M. Coupled dynamics of iron and phosphorus in sediments of an oligotrophic coastal basin and the impact of anaerobic oxidation of methane. PLoS One 2013, 8, e62386.

400 401 402

(17)

Adler, M.; Eckert, W.; Sivan, O. Quantifying rates of methanogenesis and methanotrophy in Lake Kinneret sediments (Israel) using pore-water profiles. Limnol. Oceanogr. 2011, 56, 1525– 1535.

16 ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 27

403 404

(18)

Beal, E. J.; House, C. H.; Orphan, V. J. Manganese- and iron-dependent marine methane oxidation. Science 2009, 325, 184–187.

405 406

(19)

Sivan, O.; Antler, G.; Turchyn, A. V.; Marlow, J. J.; Orphan, V. J. Iron oxides stimulate sulfate-driven anaerobic methane oxidation in seeps. Proc. Natl. Acad. Sci. 2014, 1–9.

407 408 409

(20)

Holmkvist, L.; Kamyshny, A.; Vogt, C.; Vamvakopoulos, K.; Ferdelman, T. G.; Jørgensen, B. B. Sulfate reduction below the sulfate–methane transition in Black Sea sediments. Deep Sea Res. Part I Oceanogr. Res. Pap. 2011, 58, 493–504.

410 411 412

(21)

Holmkvist, L.; Ferdelman, T. G.; Jørgensen, B. B. A cryptic sulfur cycle driven by iron in the methane zone of marine sediment (Aarhus Bay, Denmark). Geochim. Cosmochim. Acta 2011, 75, 3581–3599.

413 414 415

(22)

Riedinger, N.; Pfeifer, K.; Kasten, S.; Garming, J. F. L.; Vogt, C.; Hensen, C. Diagenetic alteration of magnetic signals by anaerobic oxidation of methane related to a change in sedimentation rate. Geochim. Cosmochim. Acta 2005, 69, 4117–4126.

416 417 418

(23)

Riedinger, N.; Formolo, M. J.; Lyons, T. W.; Henkel, S.; Beck, A.; Kasten, S. An inorganic geochemical argument for coupled anaerobic oxidation of methane and iron reduction in marine sediments. Geobiology 2014.

419 420 421

(24)

März, C.; Hoffmann, J.; Bleil, U.; de Lange, G. J.; Kasten, S. Diagenetic changes of magnetic and geochemical signals by anaerobic methane oxidation in sediments of the Zambezi deep-sea fan (SW Indian Ocean). Mar. Geol. 2008, 255, 118–130.

422 423

(25)

Lovley, D. R.; Phillips, E. J. P. Availability of ferric iron for microbial reduction in bottom sediments of the freshwater tidal Potomac River. Appl. Environ. Microbiol. 1986, 52, 751–757.

424 425 426

(26)

Roden, E. E. Diversion of electron flow from methanogenesis to crystalline Fe(III) oxide reduction in carbon-limited cultures of wetland sediment microorganisms. Appl. Environ. Microbiol. 2003, 69, 5702–5706.

427 428

(27)

Cline, J. D. Spectrophotometric determination of hydrogen sulfide in natural waters. Limnol. Oceanogr. 1969, 14, 454–458.

429 430 431

(28)

Brass, M.; Röckmann, T. Continuous-flow isotope ratio mass spectrometry method for carbon and hydrogen isotope measurements on atmospheric methane. Atmos. Meas. Tech. 2010, 3, 1707–1721.

432 433 434

(29)

Sapart, C. J.; van der Veen, C.; Vigano, I.; Brass,, M.; van de Wal, R. S. W.; Bock, M.; Fischer, H.; Sowers, T.; Buizert, C.; Sperlich, P.; et al. Simultaneous stable isotope analysis of methane and nitrous oxide on ice core samples. Atmos. Meas. Tech. 2011, 4, 2607–2618.

435 436 437

(30)

Poulton, S.; Canfield, D. Development of a sequential extraction procedure for iron: implications for iron partitioning in continentally derived particulates. Chem. Geol. 2005, 214, 209–221.

438 439 440

(31)

Burton, E. D.; Sullivan, L. A.; Bush, R. T.; Johnston, S. G.; Keene, A. F. A simple and inexpensive chromium-reducible sulfur method for acid-sulfate soils. Appl. Geochemistry 2008, 23, 2759–2766.

17 ACS Paragon Plus Environment

Page 19 of 27

Environmental Science & Technology

441 442

(32)

Bosch, J.; Heister, K.; Hofmann, T.; Meckenstock, R. U. Nanosized iron oxide colloids strongly enhance microbial iron reduction. Appl. Environ. Microbiol. 2010, 76, 184–190.

443 444 445

(33)

Jilbert, T.; Slomp, C. P. Iron and manganese shuttles control the formation of authigenic phosphorus minerals in the euxinic basins of the Baltic Sea. Geochim. Cosmochim. Acta 2013, 107, 155–169.

446 447 448

(34)

Happell, J. D.; Chanton, J. P.; Showers, W. J. Methane transfer across the water-air interface in stagnant wooded swamps of Florida : Evaluation of mass-transfer coefficients and isotopic fractionation. 1995, 40, 290–298.

449 450

(35)

Chanton, J. P. The effect of gas transport on the isotope signature of methane in wetlands. Org. Geochem. 2005, 36, 753–768.

451 452

(36)

Bak, F.; Pfennig, N. Chemolithotrophic growth of Desulfovibrio sulfodismutans sp. nov. by disproportionation of inorganic sulfur compounds. Arch. Microbiol. 1987, 147, 184–189.

453 454 455

(37)

Thamdrup, B.; Finster, K.; Hansen, J. W.; Bak, F. Bacterial disproportionation of elemental sulfur coupled to chemical reduction of iron or manganese. Appl. Environ. Microbiol. 1993, 59, 101–108.

456 457

(38)

Whiticar, M. J. Carbon and hydrogen isotope systematics of bacterial formation and oxidation of methane. Chem. Geol. 1999, 161, 291–314.

458 459

(39)

Alperin, M. J.; Reeburgh, W. S.; Whiticar, M. J. Carbon and hydrogen isotope fractionation resulting from anaerobic methane oxidation. Global Biogeochem. Cycles 1988, 2, 279–288.

460 461 462

(40)

Martens, C. S.; Albert, D. B.; Alperin, M. J. Stable isotope tracing of anaerobic methane oxidation in the gassy sediments of Eckernforde Bay, German Baltic Sea. Am. J. Sci. 1999, 299, 589–610.

463 464

(41)

Rayleigh, J. W. S. Theoretical considerations respecting the separation of gases by diffusion and similar processes. Philos. Mag. 1896, 42, 493–499.

465 466 467

(42)

Holler, T.; Wegener, G.; Knittel, K.; Boetius, A.; Brunner, B.; Kuypers, M. M. M.; Widdel, F. Substantial (13) C/(12) C and D/H fractionation during anaerobic oxidation of methane by marine consortia enriched in vitro. Environ. Microbiol. Rep. 2009, 1, 370–376.

468 469 470

(43)

Chanton, J.; Chaser, L.; Glasser, P.; Siegel, D. Carbon and hydrogen isotopic effects in microbial methane from terrestrial environments. In Stable isotopes and biosphere-atmosphere interactions; Flanagan, L., Ed.; Elsevier-Academic Press.: Amsterdam, 2005; pp. 85–105.

471 472

(44)

Lovley, D. R.; Phillips, E. J. P.; Lonergan, D. J. Hydrogen and formate oxidation coupled to dissimilatory reduction of iron or manganese by Alteromonas putrefaciens. 1989, 55.

473 474 475

(45)

Achtnich, C.; Bak, F.; Conrad, R. Competition for electron donors among nitrate reducers, ferric iron reducers, sulfate reducers, and methanogens in anoxic paddy soil. Biol. Fertil. Soils 1995, 19, 65–72.

476 477

(46)

Reiche, M.; Torburg, G.; Küsel, K. Competition of Fe(III) reduction and methanogenesis in an acidic fen. FEMS Microbiol. Ecol. 2008, 65, 88–101.

18 ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 27

478 479

(47)

Bond, D. R.; Lovley, D. R. Reduction of Fe(III) oxide by methanogens in the presence and absence of extracellular quinones. Environ. Microbiol. 2002, 4, 115–124.

480 481

(48)

Bodegom, P. M.; Scholten, J. C. M.; Stams, A. J. M. Direct inhibition of methanogenesis by ferric iron. FEMS Microbiol. Ecol. 2004, 49, 261–268.

482 483

(49)

Liu, D.; Wang, H.; Dong, H.; Qiu, X.; Dong, X.; Cravotta, C. A. Mineral transformations associated with goethite reduction by Methanosarcina barkeri. Chem. Geol. 2011, 288, 53–60.

484 485

(50)

Thauer, R. K.; Jungermann, K.; Decker, K. Energy conservation in chemotrophic anaerobic bacteria. Bacteriol. Rev. 1977, 41, 809.

486 487 488

(51)

Fleming-Lehtinen, V.; Laamanen, M.; Kuosa, H.; Haahti, H.; Olsonen, R. Long-term development of inorganic nutrients and chlorophyll alpha in the open northern Baltic Sea. Ambio 2008, 37, 86–92.

489 490 491

(52)

Fleming-Lehtinen, V.; Laamanen, M. Long-term changes in secchi depth and the role of phytoplankton in explaining light attenuation in the Baltic Sea. Estuar. Coast. Shelf Sci. 2012, 102-103, 1–10.

492 493 494

(53)

Kasten, S.; Freudenthal, T.; Gingele, F. X.; Schulz, H. D. Simultaneous formation of iron-rich layers at different redox boundaries in sediments of the Amazon deep-sea fan. Geochim. Cosmochim. Acta 1998, 62, 2253–2264.

495 496 497 498

(54)

Hensen, C.; Zabel, M.; Pfeifer, K.; Schwenk, T.; Kasten, S.; Riedinger, N.; Schulz, H.; Boetius, A. Control of sulfate pore-water profiles by sedimentary events and the significance of anaerobic oxidation of methane for the burial of sulfur in marine sediments. Geochim. Cosmochim. Acta 2003, 67, 2631–2647.

499 500

(55)

Diaz, R. J.; Rosenberg, R. Spreading dead zones and consequences for marine ecosystems. Science (80-. ). 2008, 321, 926–929.

501 502

19 ACS Paragon Plus Environment

Page 21 of 27

Environmental Science & Technology

503 504

FIGURE CAPTIONS

505

Figure 1. Location map of the study area. Site US5B (62°35.17’ N, 19°58.13’ E) is located in

506

the deepest part of the Bothnian Sea at 214 m depth with an average bottom water salinity of

507

around 6.

508

Figure 2. Geochemical profiles for site US5B. a, Porewater profiles of SO42-, CH4, sulfide

509

(∑H2S = H2S + HS- + S2-) and Fe2+. Grey bar indicates the sulfate/methane transition zone

510

(SMTZ) b, Sediment profiles of total sulfur (Stot), FeS (acid volatile sulfide, AVS), Fe-oxides

511

(see SI Figure S1 for the calculation of the Fe-oxide fraction) and total (Fe/Al).

512

Figure 3. Incubation experiment with

513

cm depth. SO42--depleted slurry incubations showed an increasing enrichment of headspace

514

CO2 in 13C after addition of Fe(III) (“13CH4 & Fe(III)”, red triangles) compared to treatments

515

where no additional Fe(III) was added (“13CH4”, green circles), suggesting stimulation of Fe-

516

AOM. Elevated δ13C-CO2 values ([13CO2] / ([12CO2]+[13CO2]) of the 13CH4-treatment without

517

additional Fe(III) compared to the control (“cntl”, blue squares) indicate Fe-AOM with

518

remaining Fe-oxides present in the sediment. 20 mmol/l Fe(III) were added at t0 (0 days) and

519

after 55 days (dashed red line). Error bars are based on duplicates for 20-30 cm and 30-35 cm

520

(i.e. n = 4; see SI Figure S2 for the calculation of the Fe-AOM rate).

521

Figure 4. Isotopic composition of porewater methane. a, Vertical porewater profiles for total

522

CH4, δD-CH4 (in ‰ vs. Standard Mean Ocean Water; SMOW) and δ13C-CH4 (in ‰ vs.

523

Vienna Pee Dee Belemnite; VPDB). b, Rayleigh fractionation plots for δD-CH4 (left) and

524

δ13C-CH4 (right). The yellow area (I) indicates the depth zone of SO4-AOM, i.e. the SMTZ,

525

while the green area (II) shows all measurements below the SMTZ. Different slopes between

526

(I, yellow) and (II, green) reveal two distinct kinetic isotope fractionation factors for hydrogen

13

C-labeled CH4 conducted on sediments from 20-35

20 ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 27

527

(εH) and carbon (εC) that are spatially separated. SO4-AOM drives a sympathetic change in C

528

and D isotopes (area (I)), whereas the antipathetic change in area (II) indicates

529

hydrogenotrophic CH4 production.

530

Figure 5. Column diffusion model simulations of the total concentration and isotopic

531

composition of porewater CH4 at station US5B. SO4-AOM (between 0-8.5 cm) is represented

532

in all model simulations (a-c), whereas different scenarios for the fate of CH4 below the

533

SMTZ (8.5-36 cm) were tested. a, CH4 loss through Fe-AOM and concurrent CH4 production

534

through methanogenesis. b, only CH4 production but no removal through Fe-AOM. c, upward

535

diffusing CH4 with no further alteration. SMOW = Standard Mean Ocean Water and VPDB =

536

Vienna Pee Dee Belemnite. See SI Table S3 for details about rate constants and fractionation

537

factors used in the model simulations.

21 ACS Paragon Plus Environment

15 °E 20 °E 25 °E PageEnvironmental 23 of 27 Science & Technology 65 °N

Finland

Sweden

US5B

Åland 60 °N

ACS Paragon Plus Environment

15 °E

20 °E Figure 1

25 °E

a 0

SO 4

2-Environmental

6000 0

[µmol/l]

Science & Technology Page 24 2+of 27

CH 4

ΣH2S

6000 0

200 0

Fe

2000

0 SMTZ

Depth [cm]

10 20 30 40 50 60

Depth [cm]

b

[µmol/g] FeS (AVS)

Stot 0

800 0

Fe-oxides

800 0

350 0.3

0 SMTZ

10 20 30 40

ACS Paragon Plus Environment

50 60

Figure 2

[mol/mol] Fe/Al

0.5

45

2nd Fe(III)-injection

Page Environmental 25 of 27 Science(after & 55 Technology days) 40 Estimated Fe-AOM rate:

1.32 ± 0.09 μmol/cm3/year

CH4 & Fe(III)

13

δ13CO2 [‰]

35 30 25

CH4

13

20 15

ACS Paragon Plus Environment 10

0

Figure 3

20

40

60

Days

80

100

cntl 120

a 0

δDCH4 [‰ ] CHEnvironmental [mM] δ13C-CH4 [‰] Page 26 of 27 Science & Technology 4 6 -300

Depth [cm]

0 10

0 -70

(I)

(I)

(II)

(II)

-50

20 30 40 50 60

0

ln(δ13C-CH4+1) [‰]

ln(δD-CH4+1) [‰]

b

εH = 98 ‰

-100

(R = 0.993) 2

(I)

-200

-300

εH = 24 ‰

(II)

-5

(R2 = 0.880)

-4

-3

-55

εC = 9.0 ‰ (R2 = 0.999)

-60 -65

εC = -1.0 ‰

(I)

(R2 = 0.795)

-70

ACS Paragon Plus Environment

-2

-1

ln(CH4(z)/CH4(max))

0

Figure 4

-4

-3

(II)

-2

-1

ln(CH4(z)/CH4(max))

0

0

1

Depth [cm]

0

CH4 [mmol/l] δD-CH4 [‰ vs. SMOW] Environmental Science & Technology -200 -100 2 3 4 5 6 0

10

30 a b c

0

10 20 30

20

δ13C-CH4 [‰ vs. VPDB]

measured

Depth [cm]

Depth [cm]

Page 27 of 27

0

-68

-64

10 20

Paragon Plus Environment Fe-AOM andACS methanogenesis 30 methanogenesis (no Fe-AOM) diffusion only Figure 5

-60

-56

-52