Methane Bubble Growth and Migration in Aquatic Sediments

Jan 29, 2018 - Figure 2. Change of sediment mechanical properties due to bubble growth. (a) Measured θg depth profiles (thick black lines) at incubat...
1 downloads 9 Views 1MB Size
Subscriber access provided by MT ROYAL COLLEGE

Article

Methane bubble growth and migration in aquatic sediments observed by X-ray µCT Liu Liu, Tim De Kock, Jeremy Wilkinson, Veerle Cnudde, Shangbin Xiao, Christian Buchmann, Daniel Uteau, Stephan Peth, and Andreas Lorke Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b06061 • Publication Date (Web): 29 Jan 2018 Downloaded from http://pubs.acs.org on January 29, 2018

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.

Page 1 of 30

Environmental Science & Technology

1

Methane bubble growth and migration in aquatic sediments observed by X-ray µCT

2

Liu Liu,1,* Tim De Kock,2 Jeremy Wilkinson,1 Veerle Cnudde,2 Shangbin Xiao,3 Christian

3

Buchmann,1 Daniel Uteau,4 Stephan Peth,4 and Andreas Lorke1

4

1

Institute for Environmental Sciences, University of Koblenz-Landau, 76829 Landau, Germany

5

2

PProGRess-UGCT, Department of Geology, Ghent University, Krijgslaan 281/S8, 9000 Ghent,

6

Belgium

7

3

8

Yichang, China

9

4

10

College of Hydraulic & Environmental Engineering, China Three Gorges University, 443002

Department of Soil Science, University of Kassel, 37213 Witzenhausen, Germany

* Corresponding author - Email address: [email protected]; Tel: +49 (0)6341 280-31584

11

12 13 14

ABSTRACT

15

Methane bubble formation and transport is an important component of biogeochemical carbon

16

cycling in aquatic sediments. To improve understanding of how sediment mechanical properties

17

influence bubble growth and transport in freshwater sediments, a 20-day laboratory incubation

18

experiment using homogenized natural clay and sand was performed. Methane bubble

19

development at high-resolution was characterized by µCT. Initially, capillary invasion by 1 ACS Paragon Plus Environment

Environmental Science & Technology

20

microbubbles (< 0.1 mm) dominated bubble formation, with continued gas production (4 d for

21

clay; 8 d for sand), large bubbles formed by deforming the surrounding sediment, leading to

22

enhanced of macropore connectivity in both sediments. Growth of large bubbles (> 1 mm) was

23

possible in low shear yield strength sediments (< 100 Pa), where excess gas pressure was

24

sufficient to displace the sediment. Lower within the sand, higher shear yield strength (> 360 Pa),

25

resulted in a predominance of microbubbles where the required capillary entry pressure was low.

26

Enhanced bubble migration, triggered by a controlled reduction in hydrostatic head, was

27

observed throughout the clay column, while in sand mobile bubbles were restricted to the upper 6

28

cms. The observed macropore network was the dominant path for bubble movement and release

29

in both sediments.

30 31

INTRODUCTION

32

Anaerobic organic matter decomposition in aquatic sediments produces methane, a potent

33

greenhouse gas. Low solubility and slow diffusive transport cause sediment gas bubble

34

accumulation. Stationary gas voids can reduce vertical solute transport , but provide a shortcut for

35

gas transport.1 Once sediment gas storage capacity is exceeded, gas exits the sediment by

36

bubbling.2 In inland waters, ebullition is an important pathway for methane release to the

37

atmosphere.3-6 Bubble release can also enhance solute transport across the sediment-water

38

interface.7-9 Thus, understanding methane bubble development and movement in sediment

39

contributes to understanding biogeochemical cycling, emission dynamics and controlling factors

40

in aquatic systems.

41

The primary control of methane bubble growth is sediment gas production rate.10-12 Methane

42

production in-excess of diffusional transport, is necessary to cause porewater supersaturation

43

leading to bubble formation. In marine sediments, methane bubble accumulation only starts 2 ACS Paragon Plus Environment

Page 2 of 30

Page 3 of 30

Environmental Science & Technology

44

below the surface layer where sulfate is present.13, 14 In freshwater sediments methane production

45

in the superficial surface layer was greatest, and decreases exponentially with depth.15, 16 Since

46

methane production generally declines with increasing depth, and bubble formation requires

47

production to exceed diffusive transport away from the source, new bubble formation can

48

generally only be expected in the upper sediment layers. In a recent study ebullition correlated

49

well with sediment methane production in a riverine impoundment, where sediment depth > 1 m

50

was considered to contribute little to total ebullition.17

51

Where methane production supports bubble formation, sediment mechanical properties become

52

important.18-20 Experiments and modelling both demonstrate the dependence of bubble growth by

53

capillary invasion on grain size.18, 21 Microbubble (< grain size) formation is favored in coarse

54

sediments, and in fine-grained sediments, bubbles (gas voids) grow larger than sediment grain

55

size by elastic/plastic deformation.22-24 Bubble growth in cohesive marine sediments has been

56

explained by linear elastic fracture mechanics (LEFM).25-27 Both size and shape of such

57

bubbles/voids is predicted well by sediment tensile fracture toughness (KIC),20, 25, 28 which was

58

also considered important for bubble migration.

59

Efficient gas transport by bubble migration can dominate over diffusion in sediments,22, 29 and is

60

controlled mainly by sediment mechanics.30, 31 In weak slurry-like sediments bubble movement

61

can be driven by buoyancy.32 In strong fine-grained sediments, initial bubble was controlled by

62

elastic fracturing, and bubble release was facilitated by vertical/sub-vertical fracture

63

propagation.25, 31 Such fracture formation and propagation can be considered as sediment tensile

64

failure, in line with LEFM for bubble growth in cohesive sediments.25, 31 These fracture/conduit

65

structures were responsible for persistent bubble release in sediment.32-35

66

While sediment bubble growth and migration behavior have been well studied in marine

67

sediments and reconstituted artificial sediments, understanding of gas storage capacity and 3 ACS Paragon Plus Environment

Environmental Science & Technology

68

sediment gas bubbles responses to external disturbances in freshwater sediments is still limited.

69

In freshwater sediments low ionic strength makes flocculation and coagulation of fine sediment

70

particles more difficult than in marine sediments.36,

71

riverine impoundments and reservoirs, may also be strongly affected by hydrologic and

72

hydrodynamic conditions,38 thus sediment grain size distribution, organic matter (OM) content

73

and sediment compaction may vary widely within and between locations, which may limit the

74

applicability of sediment mechanical properties from one system to another.

75

Here, we focus on natural sediments (sand and silty-clay) from impounded rivers often with rapid

76

sedimentation and particulate OM input derived from predominantly forested catchments.

77

Laboratory incubated sediment were characterized for bubble growth dynamics by high-

78

resolution X-ray computed microtomography (µCT),39, and time-lapse µCT scanning on

79

completion of incubations enabled characterization of bubble movement. Sediment mechanical

80

properties, shear yield strength (SYS) and compressibility, were characterized to explain

81

observed patterns of bubble formation and migration. The experiments provide a basis for

82

mechanistic modelling of freshwater sediments methane bubble storage and release.

37

Freshwater sediments, particularly in

83 84

METHODS

85

Instead of intact cores, homogenized sediments were used for two purposes: 1) to mimic methane

86

bubble growth in riverine impoundments with rapid sedimentation (30 cm yr-1) and efficient OM

87

burial such as River Saar, Germany;17 2) to avoid substrate limitation on methane production and

88

hence bubble formation, given that gas production in freshwater sediments often declines sharply

89

with depth and age.15, 16

90 91

Definitions 4 ACS Paragon Plus Environment

Page 4 of 30

Page 5 of 30

Environmental Science & Technology

92

To simplify terminology relating to the use of “bubble” and “gas void”, we define “bubble” as a

93

volume of free gas enclosed in a liquid or solid. A void can be a quasi-static feature within the

94

sediment, a cavity that may contain gas and/or water at any given time. In soils, macropores are

95

defined as structures that provide preferential pathways for water flow (bypassing the soil matrix)

96

irrespective of their size, and the measured size strongly depends on the observational methods.40,

97

41

98

spatial resolution of CT scans, here macropores are defined as pores > 100 µm in equivalent

99

diameter.

The reported equivalent diameter of macropores in soil is usually > 50 µm.40 Limited by the

100 101

Sediment Collection, Processing and Characterization

102

In June 2016, clayey and sandy sediments were sampled from a sidearm of the Rhine River in

103

Germersheim (49.221735°N, 8.382457°E) and a stream in Hochstadt (49.24678°N, 8.22675°E),

104

respectively. Riverine sediments can have rapid sedimentation and thick homogenized layers

105

(e.g., Figure S1a, b). Natural freshwater sediments generally include particulate OM, comprised

106

of leaf or woody debris, and its presence changes the particle size distribution by increasing the

107

fraction of coarse material in our clayey sediment (Figure S2). To promote homogeneous

108

enhanced gas production, we removed large pieces of OM (by sieving (> 2 mm) Figure S1c), and

109

amended sediments with powdered air-dried alder leaf (10 g L-1 wet sediment). Leaf matter

110

amendment raised sediment OM content (estimated by loss on ignition at 550 ºC) to 12.5% in

111

clay, and 2.1% in sand (increases of 1.3% and 0.7%, respectively). Sediment particle size

112

distribution was determined by laser diffraction with a particle size analyzer (Mastersizer 3000,

113

Malvern, UK) (Figure S2): the median particle size (D50) of clay was 21 µm and 352 µm for sand.

114 115

Experimental Setup 5 ACS Paragon Plus Environment

Environmental Science & Technology

116

For each sediment, paired incubation experiments in 60 cm tall, 2 mm thick transparent 6 cm

117

(inner) diameter Plexiglas tubes filled with approximately 30 cm well-mixed sediment and

118

topped-up with 15 cm tap water, sealed with rubber stoppers, were set-up. A 1.5 L inflatable gas

119

bag was fitted to the top each tube to measure total gas volume produced (P, mL). P, water level,

120

hw, and the position of the sediment-water interface (SWI) hs, in each tube were monitored daily.

121

The difference between P and the daily change in total sediment gas storage gives the ebullition

122

Eb. We express sediment gas content (θg), not as a volume, but based on the height and height

123

change within the tubes, assuming that the mass (and volume) of (gas free) sediment and water

124

remain relatively constant throughout the experiments. The θg relates to the hw and hs in the tube,

125

θg = (∑∆hw)/hs; gas storage by capillary invasion θcap = θg - (∑∆hs)/hs. ∆hw and ∆hs are the daily

126

change in hw and hs, respectively. 10 mL gas was extracted daily from gas bags to track methane

127

and CO2 concentrations (measured with a greenhouse gas analyzer (Los Gatos, US)). The

128

methane and CO2 flux was calculated from P and concentration measurements. 2 mL sediment

129

porewater samples were taken at the end of the experiment using Rhizon tubes and vacuumed 10

130

mL glass vials at different depths of the sediment columns. Porewater dissolved methane and

131

CO2 were estimated from headspace concentrations.

132

One of each paired incubated columns was used for µCT scanning and the other for excess

133

dissolved gas pressure (EDGP) measurements. Porewater EDGP was measured with 3 vented

134

pressure sensors (resolution 0.001 kPa, SENECT, Germany; 10 s sampling interval) in contact to

135

the porewater through gas-permeable membranes (Contros, Germany) via the tube side wall. Two

136

EDGP sensors were mounted in clay, 15 and 25 cm below the SWI, respectively; the third

137

measured at 20 cm below the SWI in sand.

6 ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30

Environmental Science & Technology

138

To avoid transportation disturbance of the columns, they were stored dark adjacent to the µCT

139

scanner at constant temperature (24.5 ± 0.8 °C). After incubation, cores were cut into 2 cm slices

140

and the material stored in 50 mL water-tight plastic vials awaiting sediment water content (θw)

141

and dry particle density measurement (ρdry: clay: 2553.7 kg m-3; sand: 2557.3 kg m-3).

142 143

X-ray Computed Microtomography (µCT)

144

A custom-made high-energy CT system HECTOR27 (Centre for X-ray Tomography, Ghent

145

University: www.ugct.ugent.be) was used to scan the sediment columns. Using 7 vertical

146

overlapping scans a voxel (volumetric pixel) size of 64.1 µm for the full column scan was

147

achieved. A 1 mm Cu filter reduced beam hardening and the X-ray tube was operated at 60 W,

148

190 kV for clay and 55 W, 190 kV for sand. Each full column scan was followed by a high-

149

resolution scan (voxel size of 19.7 µm), in a central cylindrical region of interest (ROI) (19.7 mm

150

diameter, 19.7 mm height), without a filter (15-16 W, 190 kV). With scans at incubation day 1, 4,

151

8, 14 and 20, the progression of bubble growth was followed. The sand had one additional ROI

152

scan in the upper 5 cm layer at day 20, but was skipped for clay because bubbles were adequately

153

captured by column scans. Followed the day 20 scans, sediment column water level was dropped

154

8 cm to trigger bubbling, and a further column scan performed after 8 hours. In between scans,

155

the columns were left undisturbed to allow for accurate spatial tracing of bubble movement.

156 157

CT Data Analysis

158

Following image processing and segmentation (Text S1), gas bubble and pore parameters

159

(bubble/pore volume, equivalent spherical diameter (Deq), connectivity, bubble shape, orientation

160

and θg) were extracted from the final images using Octopus Analysis (formerly Morpho+).42

161

Macropore density (ρpore) is simply the number of macropores (N) per unit analyzed sediment 7 ACS Paragon Plus Environment

Environmental Science & Technology

162

volume. The overall macropore connectivity (at voxel size 64.1 µm) was quantified using Euler–

163

Poincaré characteristic (E)43, and decreasing E/N indicates increasing macropore connectivity.

164

Bubble sphericity (0-1) was characterized as the ratio of the largest inscribed sphere diameter to

165

Deq. Bubble orientation (0-180o) was characterised by the angle of the principal axis of the

166

equivalent ellipsoid to the vertical (z-axis).

167

Sediment column gas bubble migration was demonstrated by generating difference images for the

168

last two day 20 µCT scan images (processed using DataViewer: Bruker microCT, Belgium). The

169

resulting images were analyzed in Octopus applying column-specific threshold values; a low

170

threshold for newly formed bubbles originally filled by water, a high threshold for those that

171

disappeared due to movement.

172 173

Sediment Rheology and Compressibility

174

Using measured θw at day 20 (Figure S4) and ρdry, gas-free sediment wet bulk density (ρwet) was

175

calculated. For each scan, volumetric fraction of solid sediment particles (θs) was computed from

176

measured CT scan θg profiles and θcap data (14.9-22.8% for clay, and 44.2-70.4% for sand).

177

Profiles of ρwet including gas phase for all scans were also calculated (Figure S5).

178

Shear yield strength was measured using a rheometer (Anton Paar, MCR 102, Austria) involving

179

a four-blade vane (22 mm diameter) inserted into a cup (29 mm diameter) containing gas free

180

homogenized sediment samples (with θs ranging from 9.0-24.7% for clay and 37.4-61.6% for

181

sand; Figure S6). Strain was measured while increasing shear stress from 0 Pa in logarithmically

182

distributed increments until sediment failure; the SYS value was taken at the break point from

183

initial linearity.44 The clear exponential dependence of SYS on θs enabled back calculation of

8 ACS Paragon Plus Environment

Page 8 of 30

Page 9 of 30

Environmental Science & Technology

184

SYS depth profiles for each CT scan (assuming θs changes due to θg development resulted in

185

similar SYS change).

186

Sediment compressibility was tested using an oedometer (model 08.67, Eijkelkamp®, the

187

Netherlands), on samples with a range of θs (clay, 20.1-26.2%, and sand 52.6-68.7%), under

188

static incremental loads (clay 5, 10, 20 … 60 kPa, and sand 5, 10, 50 … 600 kPa). From this the

189

pre-compression stress of the sediments could be related to their θs (for details refer to Text S2

190

and Figure S7). At vertical stress 5 and 10 kPa (lying in the range of measured sediment EDGP),

191

sediment volumetric deformation and θs had a strong linear correlation (Figure S8) enabling the

192

prediction of θg depth profiles (at 5 and 10 kPa EDGP) from the initial depth θs for both

193

sediments.

194 195

RESULTS

196

During incubation, θg in clay increased sharply, reaching a maximum ~20% at day 4 (Figure S9a),

197

and stabilized thereafter ~18.4%. In sand, θg development was slower and stabilized at ~15.2%

198

after 8 days. In accordance with θg development, ebullition was less intense in the first 2 days.

199

The occurrence of steady state, i.e. gas production equals ebullition (Eb/P = 100%), after day 5

200

for clay and day 9 for sand, indicated that sediment gas storage capacity had been achieved

201

(Figure S9b). The observed methane flux was minimum (< 0.01 mmol day-1) at the initial stage of

202

θg development (until day 3 in clay and day 5 in sand), and then was enhanced dramatically at the

203

steady state (~2 orders and 1 order higher for clay and sand, respectively) (Figure S9d), which

204

highlights the importance of ebullition in methane transport. Compared to the change of

205

methane:CO2 ratio in headspace (which was consistent to the change of Eb/P) (Figure S10a), this

206

ratio experienced an initial peak at the early stage and was stabilized at the steady state (~1.7 for 9 ACS Paragon Plus Environment

Environmental Science & Technology

207

clay and ~0.3 for sand) (Figure S10b. This can be explained by the greater solubility of CO2

208

relative to methane and more CO2 went into solution at the early stage before saturation was

209

reached. By the end of incubation, ~3% and ~13% methane was found in solution in clay and

210

sand, respectively, while ~50% for CO2. The estimated methane:CO2 ratio of the total production

211

during the incubation experiment was ~1 and ~0.3 in clay and sand, respectively.

212 213

Bubble growth by capillary invasion

214

Capillary invasion dominated gas content development in both sediments (Figure S9c), although

215

more so for sand than clay. In clay, θcap/θg dropped sharply from the initial 100% to 47.6% in the

216

first 4 days finally stabilizing at ~67%; in sand, θcap/θg was 100% in the first 5 days decreasing

217

linearly to ~80% by day 12. Capillary invasion was also evidenced by EDGP dynamics (Figure

218

S11), decreasing sharply by 10.0 kPa in the surface layer, and 8.0 kPa at lower depths during the

219

first 4 days; in sand, EDGP decreased by 5.0 kPa from an initial 3.6 kPa during day 2-8. This

220

sediment EDGP reduction was in accordance with bubble growth dynamics by capillary invasion

221

(Figure S9c).

222

The sand microbubble size distribution (from µCT scans) peaked at 60 µm diameter (Figure S12).

223

Indirect evidence from EDGP measurements suggest microbubbles in clay with a diameter of 30

224

µm (estimated from the 10.0 kPa EDGP decrease), fall into the size range of capillary pores. This

225

confirms the initial occurrence of capillary invasion for bubble formation in both sediments and

226

is consistent with Reed et al.45 whose µCT scans found a similar size range (diameter > 60 µm) of

227

bubbles in reconstituted clay.

228 229

Dependence of large bubble growth on sediment mechanical properties

10 ACS Paragon Plus Environment

Page 10 of 30

Page 11 of 30

Environmental Science & Technology

230

Both gas content profiles (Figure S9e and f) and 2D vertical CT slices (Figure S13) showed the

231

development of large depth gradients of gas content, and bubble size distribution, in clay and

232

sand, these are attributable to sediment mechanical properties.

233

Significant changes in bubble size distribution in both sediments were shown by the µCT scans

234

(Figure 1), these can be categorized to two general modes of bubble growth: 1) bubble density

235

(ρbub, number of bubbles per sediment volume) decrease associated with increasing bubble

236

volume; 2) ρbub increasing over time without bubble volume increase. The bubble growth mode 1

237

was observed throughout the clay column, but only in the surface layer of sand, where ρbub

238

decreased by two orders of magnitude. Initially, the bubble size distribution was log-normal at all

239

depths, becoming bimodal towards the end. Bubble growth in the sand mid-layer was

240

characterised mode 2: the log-normal bubble size distribution persisted over time with an

241

increasing ρbub of microbubbles, while the peak bubble size was decreasing, suggesting bubble

242

growth by invading smaller pores. These modes of bubble growth were confirmed by 3D

243

visualization (Figure 1b), which also enabled detection of a change not apparent from bubble size

244

distribution statistics alone: the bubbles in the clay surface layer were significantly less at day 20

245

compared to day 8 in both density and volume.

246

The maximum gas storage capacity of the sediments could be related to sediment compressibility.

247

Expected clay θg at 5 kPa EDGP was 19-24.9% (Figure 2a), comparable to θg at steady state

248

bubble growth (except for the upper 4 cm). The measured maximum clay θg was in the range

249

predicted for 10 kPa EDGP. In sand, predicted θg at 5 kPa load only explained half the gas

250

content formation. Increasing sand EDGP from 5 to 10 kPa led to < 1% volume expansion,

251

demonstrating poor sand compressibility. Porewater drainage from sand was much faster (5 min)

11 ACS Paragon Plus Environment

Environmental Science & Technology

252

than for clay (90 min) (see compression curves, Figure S7), indirectly confirming that capillary

253

invasion in sand is easier compared to water displacement in clay.

254

SYS increased exponentially with θs for both sediments (Figure S6), and was affected by the

255

counteracting processes, sediment compaction and gas content development, the former

256

increasing shear strength, and the latter weakening it. Both effects were both minor in clay. All

257

clay SYS were < 15 Pa, whereas, in sand, compaction enhanced SYS over the entire depth (day

258

1-4), and gas content development decreased sediment strength (from day 4); most apparent in

259

the surface layer (SYS ranging between 99-376 Pa at day 4, compared to 25-94 Pa at day 20).

260

As with gas content, sediment compressibility explained bubble size distribution change over

261

depth in clay well (e.g. the clay surface layer bubble size distribution peak decreased from 1.7

262

mm to 1 mm in the bottom layer, Figure S14). Bubble sphericity and orientation (in addition to

263

gas content and bubble size) were closely related to depth gradients in sediment mechanical

264

properties. In surface clay, bubbles were near spherical (sphericity = 0.6), but at lower depths

265

were elongated (sphericity < 0.4) and ~50% of bubbles were horizontally-oriented. This pattern

266

was stronger in sand where bubble shape changed from spherical (sphericity = 0.6) above 3.5 cm

267

to being horizontally-oriented elongated bubbles (sphericity = 0.3, ~34% bubbles horizontally-

268

oriented) at 3.5-6 cm depth.

12 ACS Paragon Plus Environment

Page 12 of 30

Page 13 of 30

Environmental Science & Technology

269 270

Figure 1. (a) Bubble size distribution in surface (0-6 cm) and mid (13-19 cm) layers of sediment

271

columns at different days of incubation. The probability density of bubble volume was

272

normalized by the analyzed sediment volume, i.e., the integration of each distribution equals the

273

total volumetric gas content of the analyzed sediment layer. (b) 3D visualization of gas voids

274

(golden color with shading) in a control volume (1 cm3) at selected sediment column depths.

275 276

13 ACS Paragon Plus Environment

Environmental Science & Technology

277 278

Figure 2. Change of sediment mechanical properties due to bubble growth. (a) measured θg depth

279

profiles (thick black lines) at incubation day 20 and predicted θg for initial sediment at different

280

EDGP (5 and 10 kPa; represented by red solid and dotted lines, respectively) from sediment

281

compressibility test; (b) sediment shear yield strength (SYS) for clay (left side, red lines) and

282

sand (right side, black lines) at different incubation days calculated from θs depth profiles.

283 284

Development of sediment macropores

285

Sediment gas content development not only changed sediment strength, but also altered

286

macropore structure due to plastic sediment volumetric deformation (Table 1). In response to

287

methane bubble growth, ρpore decreased at all depths in both sediments, and the total volume of

288

macropores (macro-porosity) increased. In clay, macro-porosity doubled over the entire depth 14 ACS Paragon Plus Environment

Page 14 of 30

Page 15 of 30

Environmental Science & Technology

289

during the period of incubation; in sand, initial macro-porosity was high and the increase in pore

290

volume less pronounced (6% in the mid layer, 10.6% in the surface layer) due to the dominance

291

of bubble growth by capillary invasion. The reduction in ρpore and increase in macro-porosity,

292

enhanced macropore connectivity in both sediments, was reflected by changes in E/N. Despite

293

high macropore connectivity (E/N < 1) in both sediments at steady state gas content development

294

(suggesting a well-developed macropore network) E/N was initially more variable. A large

295

decrease in clay E/N (0.6 to -0.6 in the surface layer, and 1.1 to -3.6 in the mid layer) was

296

observed over the first 4 incubation days. Conversely, an increase in E/N was seen in sand (+5.0

297

surface, and +2.3 mid layer) during the first 4 days, this however reversed from day 4-8.

298

Following the initial changes in E/N (day 4 clay, and day 8 sand), a steady increase was observed

299

in both sediments.

300

Macropore connectivity development was closely related to large bubble growth (Figure 3). In

301

surface clay, isolated small bubbles with relatively weak pore connections were observed at day 1

302

(Figure 3a). As bubbles grew larger and maximum sediment gas storage reached (day 4, Figure

303

1b and Figure S9a), macropores (mainly gas-filled) were enlarged and their connectivity was

304

significantly enhanced. From day 4 in clay, intense bubble release (Figure S9b) led to the

305

formation of water-filled macropores (Figure 3a). This was also observed in sand but with a

306

further 4 day delay. Macropore connectivity decrease in days 1-4 was associated with slight

307

initial settling, and microbubble formation by capillary invasion did not contribute to the change

308

in sediment macropore connectivity (Figure 1b and Figure S9c). The great increase in macropore

309

connectivity during day 4-8 coincided with large bubble growth by sediment matrix deformation

310

(Figure S9c).

311 312 15 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 30

313

Table 1. Macropore connectivity (E/N), macropore density (ρpore) and macro-porosity at two

314

selected depths in two sediment columns over time. Macropore connectivity increases with

315

decreasing E/N (macropores were isolated when E/N ≥ 1); ρpore is the number of pores

316

normalized by the analyzed sediment volume. Sediment type

Clay

Sand

317 318

Day 1 4 8 14 20 1 4 8 14 20

Surface 6 cm layer Macroρpore E/N -1 porosity % (mL ) 0.60 -0.60 -0.30 0.10 0.10 -5.00 0.02 -3.60 -3.10 -2.10

314 63 164 203 192 491 698 292 354 349

13.9 37.2 34.7 30.0 30.2 29.0 22.7 41.6 39.9 39.6

Bu

16 ACS Paragon Plus Environment

Mid 13-19 cm layer Macroρpore E/N -1 porosity % (mL ) 1.10 -3.60 0.10 0.60 0.60 -1.80 0.50 -16.70 -14.70 -11.60

365 25 114 160 172 645 716 150 179 195

15.0 41.0 33.4 29.0 28.6 26.7 20.2 32.8 31.8 32.7

Page 17 of 30

Environmental Science & Technology

319 320

Figure 3. (a) 3D visualization of macropore and gas voids in a control volume (1 cm3) in the

321

surface sediment layer; (b) 3D visualization of macropore and gas voids in surface layer of clay

322

and sand at day 20. Water-filled macropores are colored blue/green and gas voids red/yellow.

323 324

Bubble mobility

325

At incubation day 20, 8 cm water-level drawdown experiments were conducted. Each sediment

326

column was scanned before and after the water-level change enabling bubble mobility

327

characterization. Sediment pores in the clay upper layer remained stable over 8 h and vigorous

328

bubble movement was observed (Figure 4a and b), with gas movement clearly apparent in the

329

difference image of the two scans (Figure 4c). While gas movement occurred over the entire clay

330

column, it was restricted to the uppermost 6 cm in sand (Figure 4d). The estimated gas bubble

331

flux from clay (125.2 mL d-1) was ~5.8 times greater than that from sand (21.6 mL d-1). 17 ACS Paragon Plus Environment

Environmental Science & Technology

332

These bubble mobility differences in clay and sand are related to macropores, which enhance

333

pore connectivity and decrease capillary pressure. In sand, this was observed in the surface layer,

334

but not the mid layer. Yet, the diameter of all mobile bubbles was in the millimeter range (0.2-6.1

335

mm in clay, 0.2-3.5 mm in sand), comparable to the pore sizes (Figure 5).

336

The shape of mobile bubbles in both sediments showed no strong pattern. In clay mean mobile

337

bubble sphericity (Figure 4(c)) was 0.4, smaller but comparable to bubbles before triggering

338

movement (sphericity = 0.6, e.g. Figure 4(a)). In addition, no preferential orientation of mobile

339

bubbles was observed in either sediment.

340 341

Figure 4. (a) - (b) time-lapse scans of the surface 6 cm of clay. (c) difference between images (a)

342

and (b), bubble and porewater movement are shown as white and black patches, respectively. The

343

black circles identify a newly formed bubble at 10pm; the red circles highlight the disappearance

344

of a bubble originally trapped in a pore. The height and diameter of analyzed volume are marked

18 ACS Paragon Plus Environment

Page 18 of 30

Page 19 of 30

Environmental Science & Technology

345

by yellow axes. (d) Mean vertical profiles of bubble movement. The black lines show the changes

346

in volume fraction of gas phase.

347 348

Figure 5. Sediment bubble and pore size distributions. The probability density of bubble volume

349

was normalized by the analyzed sediment volume. Black and red lines show pore and bubble size

350

distributions in the upper 6 cm of the sediments before water-level change, respectively; and blue

351

lines are for mobile bubbles.

352 353

DISCUSSION

354

The role of sediment mechanical properties in bubble growth

355

Methane bubbles in soft marine sediments have been characterized as disk-like in shape and with

356

a vertical/sub-vertical orientation.26, 46 Their growth can be well explained by LEFM theory,19, 25

357

i.e., these bubbles grow by elastic fracture of the sediment. Disk-like bubble development was

358

not observed in this study, instead, sediment displacing bubbles both in clay and sand were either 19 ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 30

359

close to spherical or were elongated bubbles with a horizontal orientation, suggesting the control

360

of bubble growth by elastoplastic deformation of sediment matrix rather than elastic fracturing.

361

In clay, compressibility, explains the growth of larger bubbles, due to sediment matrix

362

deformation, although a quantitative compressibility to bubble size relationship remains elusive.

363

The predicted θg at 5 and 10 kPa EDGP was consistent with observed θg despite a few percent

364

overestimation (Figure 2); whereas in sand this discrepancy was > 10%. In clay, where pore

365

drainage is slow, applied compression load is effectively taken by the porewater due to the high

366

water content and the visco-plastic fluid-like sediment can be easily deformed (Figure S7).

367

Conversely, in sand, the free drainage of porewater means the applied stress is taken by the

368

sediment matrix, in accordance with the dominance of bubble formation by capillary invasion

369

(θcap/θg > 80%), hence the underestimation of θg in sand is likely due to porewater displacing

370

microbubbles.

371

In clay, the underestimation of θg by sediment compressibility (in response to measured EDGP)

372

may be due to sediment gas storage capacity and controlled by buoyancy-induced instability. In

373

clay, the measured EDGP ranged from 8.1-10 kPa, corresponding to an expected θg of 29.6%, 9.9%

374

higher than the θg measured by µCT. High gas content, however, results in ρwet < water density,

375

leading to buoyancy-induced instability where the excessive gas content is released upwards.32

376

The critical θg for the initial gas free ρwet for clay (1270 kg m-3) was 25.5%. Depth profiles of ρwet

377

indicate the presence thin gas-charged layers below 5 cm depth with ρwet < 1000 kg m-3 (Figure

378

S5), and hence low mechanical stability zones where buoyancy can limit gas storage capacity.

379

Such a limitation on maximum gas storage did not apply in sand because ρwet > 1200 kg m-3.

380

One study found that bubble size in weak sediment (SYS < 10 Pa) was limited to 9 mm

381

(equivalent spherical diameter) above which bubbles leave the sediment, but in strong sediment 20 ACS Paragon Plus Environment

Page 21 of 30

Environmental Science & Technology

382

(SYS of a few hundred Pa) gas release was facilitated by fractures.47 In the present experiment,

383

sand SYS (~102-104 Pa) would be sufficient to stabilize fractures,32 and clay was weak (SYS < 15

384

Pa) and continuous long fractures could not form, whereas short fractures/macropores that

385

connect macropores can be stabilized (Figure 3b). Depth gradients in bubble shape in sand

386

correspond with strong vertical SYS gradient; bubbles changed from spherical to horizontal

387

elongated shapes below 3.5 cm depth where SYS was ~94 Pa and doubled to ~192 Pa at 6 cm

388

depth, as observed elsewhere.47 In clay, the large SYS depth gradient was absent, but similar

389

change in bubble shape was observed. The formation of horizontal bubbles may be explained by

390

sediment compactness increase with depth (Figure S4), i.e., sediment was more compressible

391

horizontally than vertically, as previously reported for gas dome formation in cohesive

392

sediments.48

393 394

Macropore networks: a framework for bubble migration

395

Macropore transport appeared as the dominant form of gas bubble movement in this study. The

396

development of connected macropore structures, where bubbles could accumulate and move, was

397

revealed by µCT scans. These structures are analogous to macropores in soil, which serve as

398

preferential pathways for air and water movement.40, 41, 49 The most apparent and intense bubble

399

migration (Figure 4, 5) occurred in sediment layers containing large macropores, in the mm to cm

400

scale, produced by sediment deformation during bubble formation. The pre-existing sediment

401

macropores in sub-mm range (e.g. in sand mid layer) were largely occupied by microbubbles

402

(~50%) (Figure S11). Gas transport in this layer was facilitated by direct breakthrough due to

403

high inter-connectivity, which was evidenced by the sharp increase of EDGP in sand prior to day

404

8. The extent of bubble movement depended on the stage of macropore development, which, with

405

concurrent ebullition, increasing during sediment expansion in both clay and sand (Figure S1b 21 ACS Paragon Plus Environment

Environmental Science & Technology

406

and c). Bubbles occurred with a relatively uniform spatial pattern at discrete distance intervals

407

within the existing macropore network, this general pattern persisted, and was similar to

408

observations of rising bubbles in glass bead columns.50 This is also supported by the videos in

409

previous experiments2, 25 where bubble chain movement in pores was observed. Macropores are

410

important for solute transport in soils,51, 52 and for bubbles in this study - where macropores were

411

connected by fractures. The upward moving bubbles can be facilitated by vertically/sub-vertically

412

oriented fractures opened by the passage of previously migrated bubbles (Figure 3b and Figure

413

4a/b), which was previously explained by viscoelastic fracturing.31

414

Potential alternative bubble migration mechanisms to macropore transport include, buoyant

415

migration (as discussed above) and fluidization32,

416

transport are very small (< 0.1 mm), and immobile bubbles in our experiments had Deq > 2 mm at

417

steady state, and since smaller bubbles moved freely within macropores our observations do not

418

support fluidization as a relevant bubble transport mechanism. Indeed, Johnson et al.47 observed

419

no fluidization, even in extremely low-strength sediments (SYS = 7 Pa). Regarding buoyancy, as

420

already discussed above, only large bubbles Deq > 9 mm have been found to migrate, and only in

421

very weak sediments.47 In our study, larger bubbles (Deq > 2 mm), tended to be stationary (Figure

422

5), only small bubbles (Deq < 2 mm) were dominantly mobile, and this mobility was strongly

423

influenced by the relative size of bubbles to pores. So, despite low sediment strength in our clay

424

column (SYS < 15 Pa), we found no evidence to suggest that buoyant migration was an important

425

bubble movement mechanism, and we propose that macropore transport is dominant in feeding

426

ebullition in highly methane productive freshwater sediment systems.

47

. Critical bubble diameters in fluidized

427 428

Implications

22 ACS Paragon Plus Environment

Page 22 of 30

Page 23 of 30

Environmental Science & Technology

429

In our experiments, small scale, but high resolution examination of bubble formation and

430

transport mechanisms in freshwater sediment was achieved. The small diameter columns were a

431

limitation, and to investigate larger-scale structure development, incubation vessels should be

432

several meters wide.

433

structure in the surface layer of clay (Figure 4a and b). Although initially homogenized natural

434

sediments were used, sediment porosity in clay was 80.8-81.6% at incubation day 20 when

435

steady-state macropore structure was reached. This is consistent with porosity of freshwater

436

sediments previously reported in lakes and river impoundments (e.g., Lake Kinneret, Israel - 70-

437

87%55 and Saar River, Germany 80 ± 0.4%5). The addition of leaf matter was also reasonable,

438

and stimulated methane production consistent with levels observed in natural river cores.17 The

439

relatively high methane:CO2 ratio (~1.7) in gas bubbles produced from incubated natural river

440

sediments suggests the potential for greenhouse gas emission mitigation by harvesting methane

441

and burning to CO2, as proposed elsewhere for large tropical reservoirs. 56, 57

442

We found that the depth range, where gas bubbles can be mobilized by changing hydrostatic

443

pressure differed considerably between sediments, highlighting the importance of the pore

444

structure for ebullition dynamics in sediments. We believe that the observations shown here are

445

applicable to natural systems with commonly occurring, strong flood depositional events,

446

resulting in thick well-mixed sediment layers rich in OM (e.g. Figure S1), and that extrapolation

447

to freshwater systems with gradual sedimentation (e.g. lakes) is not appropriate.

448

Direct high-resolution observations from µCT revealed the importance of macropores in methane

449

bubble migration in freshwater sediment, and sediment SYS was a key physical property in

450

determining the depth of the surface zone exhibiting macropore network development. Sediment

451

gas content could be predicted by sediment compressibility, which is an estimation of sediment

452

elastic-plastic deformation under applied loads. An improved model, which incorporates these

53, 54

Despite this limitation, µCT captured a well-developed macropore

23 ACS Paragon Plus Environment

Environmental Science & Technology

453

physical characteristics, such as depth-dependent macro-porosity and pore size distribution, can

454

be expected to better capture bubble storage and release dynamics in aquatic sediments.

455 456

ACKNOWLEDGEMENTS

457

The authors would like to thank Marijn A. Boone (XRE) for his contribution to X-ray CT 3D

458

image reconstructions and Inka Meyer for helping with sediment grain size analysis. Thanks to

459

Jeroen Van Stappen for his help during experiment preparation and to Björn Krüger for his

460

assistance with sediment compressibility test. This study was financially supported by the

461

German Research Foundation (grant LO 1150/5). Tim De Kock is a postdoctoral fellow of the

462

Research Foundation - Flanders (FWO) and acknowledges its support.

463 464

Supporting Information

465

Additional figures on examples of natural sediments taken from River Rhine, sediment grain size

466

distributions, examples of filtered µCT images, depth profiles of sediment volumetric water

467

content (θw), solid fraction (θs) and gas content (θg), sediment wet bulk density (ρwet) depth

468

profiles, sediment yield shear strength (SYS) at different solid volume fraction (θs), examples of

469

compression test using the odometer, volumetric deformation in response to 5 and 10 kPa vertical

470

stress, respectively at different solid volume fraction (θs), overview of gas content development,

471

CH4:CO2 ratio in headspace and gas bubbles, excess dissolved gas pressure (EDGP) in porewater,

472

pore and bubble size distributions in region of interest scans, vertical µCT slices of gas bubble

473

growth, bubble size distribution at different depths of clay at day 20; additional text on CT image

474

processing and segmentation, sediment compressibility test, sediment methane (CH4) and CO2

475

production.

476 24 ACS Paragon Plus Environment

Page 24 of 30

Page 25 of 30

Environmental Science & Technology

477

References

478

(1) Flury, S.; Glud, R. N.; Premke, K.; McGinnis, D. F. Effect of Sediment Gas Voids and

479

Ebullition on Benthic Solute Exchange. Environ. Sci. Technol. 2015, 49 (17), 10413-10420.

480

(2) Liu, L.; Wilkinson, J.; Koca, K.; Buchmann, C.; Lorke, A. The role of sediment structure in

481

gas bubble storage and release. J. Geophys. Res. Biogeosci. 2016, 121 (7), 1992-2005.

482

(3) DelSontro, T.; McGinnis, D. F.; Sobek, S.; Ostrovsky, I.; Wehrli, B. Extreme methane

483

emissions from a Swiss hydropower reservoir: contribution from bubbling sediments. Environ.

484

Sci. Technol. 2010, 44 (7), 2419-2425.

485

(4) Baulch, H. M.; Dillon, P. J.; Maranger, R.; Schiff, S. L. Diffusive and ebullitive transport of

486

methane and nitrous oxide from streams: Are bubble‐mediated fluxes important? J. Geophys.

487

Res. Biogeosci. 2011, 116 (G04028); doi:10.1029/2011JG001656.

488

(5) Maeck, A.; DelSontro, T.; McGinnis, D. F.; Fischer, H.; Flury, S.; Schmidt, M.; Fietzek, P.;

489

Lorke, A. Sediment trapping by dams creates methane emission hot spots. Environ. Sci. Technol.

490

2013, 47 (15), 8130-8137.

491

(6) Xiao, S.; Yang, H.; Liu, D.; Zhang, C.; Lei, D.; Wang, Y.; Peng, F.; Li, Y.; Wang, C.; Li, X.

492

Gas transfer velocities of methane and carbon dioxide in a subtropical shallow pond. Tellus B

493

2014, 66 (1), 23795; DOI: 10.3402/tellusb.v66.23795.

494

(7) Klein, S. Sediment porewater exchange and solute release during ebullition. Mar. Chem. 2006,

495

102 (1), 60-71.

496

(8) Viana, P. Z.; Yin, K.; Rockne, K. J. Field measurements and modeling of ebullition-facilitated

497

flux of heavy metals and polycyclic aromatic hydrocarbons from sediments to the water column.

498

Environ. Sci. Technol. 2012, 46 (21), 12046-12054.

25 ACS Paragon Plus Environment

Environmental Science & Technology

499

(9) Cheng, C. H.; Huettel, M.; Wildman, R. A. Ebullition-enhanced solute transport in coarse-

500

grained sediments. Limnol. Oceanogr. 2014, 59 (5), 1733-1748.

501

(10) Lubetkin, S. The fundamentals of bubble evolution. Chem. Soc. Rev. 1995, 24, (4), 243-250.

502

(11) Jones, S.; Evans, G.; Galvin, K. Bubble nucleation from gas cavities - a review. Adv. Colloid

503

Interface Sci. 1999, 80 (1), 27-50.

504

(12) Boudreau, B. P.; Gardiner, B. S.; Johnson, B. D. Rate of growth of isolated bubbles in

505

sediments with a diagenetic source of methane. Limnol. Oceanogr. 2001, 46 (3), 616-622.

506

(13) Martens, C. S.; Berner, R. A. Methane production in the interstitial waters of sulfate-

507

depleted marine sediments. Science 1974, 185 (4157), 1167-1169.

508

(14) Flury, S.; Røy, H.; Dale, A. W.; Fossing, H.; Tóth, Z.; Spiess, V.; Jensen, J. B.; Jørgensen, B.

509

B. Controls on subsurface methane fluxes and shallow gas formation in Baltic Sea sediment

510

(Aarhus Bay, Denmark). Geochim. Cosmochim. Acta 2016, 188, 297-309.

511

(15) Falz, K. Z.; Holliger, C.; Grosskopf, R.; Liesack, W.; Nozhevnikova, A.; Müller, B.; Wehrli,

512

B.; Hahn, D. Vertical distribution of methanogens in the anoxic sediment of Rotsee (Switzerland).

513

Appl. Environ. Microbiol. 1999, 65 (6), 2402-2408.

514

(16) Nüsslein, B.; Eckert, W.; Conrad, R. Stable isotope biogeochemistry of methane formation

515

in profundal sediments of Lake Kinneret (Israel). Limnol. Oceanogr. 2003, 48 (4), 1439-1446.

516

(17) Wilkinson, J.; Maeck, A.; Alshboul, Z.; Lorke, A. Continuous seasonal river ebullition

517

measurements linked to sediment methane formation. Environ. Sci. Technol. 2015, 49 (22),

518

13121-13129.

519

(18) Jain, A.; Juanes, R. Preferential mode of gas invasion in sediments: Grain‐scale mechanistic

520

model of coupled multiphase fluid flow and sediment mechanics. J. Geophys. Res. Solid Earth

521

2009, 114, B08101; doi:10.1029/2008JB006002.

26 ACS Paragon Plus Environment

Page 26 of 30

Page 27 of 30

Environmental Science & Technology

522

(19) Boudreau, B. P. The physics of bubbles in surficial, soft, cohesive sediments. Mar. Pet. Geol.

523

2012, 38 (1), 1-18.

524

(20) Katsman, R.; Ostrovsky, I.; Makovsky, Y. Methane bubble growth in fine-grained muddy

525

aquatic sediment: Insight from modeling. Earth Planet. Sci. Lett. 2013, 377, 336-346.

526

(21) Choi, J. H.; Seol, Y.; Boswell, R.; Juanes, R. X‐ray computed‐tomography imaging of gas

527

migration in water‐saturated sediments: From capillary invasion to conduit opening. Geophys.

528

Res. Lett. 2011, 38, L17310; doi:10.1029/2011GL048513.

529

(22) Wheeler, S. A conceptual model for soils containing large gas bubbles. Geotechnique 1988,

530

38 (3), 389-397.

531

(23) Sills, G.; Wheeler, S.; Thomas, S.; Gardner, T. Behaviour of offshore soils containing gas

532

bubbles. Geotechnique 1991, 41 (2), 227-241.

533

(24) Wheeler, S.; Gardner, T. Elastic moduli of soils containing large gas bubbles. Geotechnique

534

1989, 39 (2), 333-342.

535

(25) Boudreau, B. P.; Algar, C.; Johnson, B. D.; Croudace, I.; Reed, A.; Furukawa, Y.; Dorgan, K.

536

M.; Jumars, P. A.; Grader, A. S.; Gardiner, B. S. Bubble growth and rise in soft sediments.

537

Geology 2005, 33 (6), 517-520.

538

(26) Gardiner, B.; Boudreau, B.; Johnson, B. Growth of disk-shaped bubbles in sediments.

539

Geochim. Cosmochim. Acta 2003, 67 (8), 1485-1494.

540

(27) Johnson, B. D.; Boudreau, B. P.; Gardiner, B. S.; Maass, R. Mechanical response of

541

sediments to bubble growth. Mar. Geol. 2002, 187 (3), 347-363.

542

(28) Katsman, R. Correlation of shape and size of methane bubbles in fine-grained muddy

543

aquatic sediments with sediment fracture toughness. J. Struct. Geol. 2015, 70, 56-64.

544

(29) Wheeler, S. Movement of large gas bubbles in unsaturated fine‐grained sediments. Mar.

545

Georesour. Geotec. 1990, 9 (2), 113-129. 27 ACS Paragon Plus Environment

Environmental Science & Technology

546

(30) Algar, C. K.; Boudreau, B. P.; Barry, M. A. Release of multiple bubbles from cohesive

547

sediments. Geophys. Res. Lett. 2011, 38, L08606; doi:10.1029/2011GL046870.

548

(31) Algar, C. K.; Boudreau, B. P.; Barry, M. A. Initial rise of bubbles in cohesive sediments by a

549

process of viscoelastic fracture. J. Geophys. Res. Solid Earth 2011, 116, B04207;

550

doi:10.1029/2010JB008133.

551

(32) Van Kessel, T.; Van Kesteren, W. Gas production and transport in artificial sludge depots.

552

Waste Manag. 2002, 22 (1), 19-28.

553

(33) Scandella, B. P.; Delwiche, K.; Hemond, H.; Juanes, R. Persistence of bubble outlets in soft,

554

methane‐generating sediments. J. Geophys. Res. Biogeosci. 2017, 122 (6), 1298-1320.

555

(34) Scandella, B. P.; Varadharajan, C.; Hemond, H. F.; Ruppel, C.; Juanes, R. A conduit dilation

556

model of methane venting from lake sediments. Geophys. Res. Lett. 2011, 38, L06408;

557

doi:10.1029/2011GL046768.

558

(35) Bussmann, I.; Damm, E.; Schlüter, M.; Wessels, M. Fate of methane bubbles released by

559

pockmarks in Lake Constance. Biogeochemistry 2013, 112 (1-3), 613-623.

560

(36) Droppo, I.; Leppard, G.; Flannigan, D.; Liss, S. The freshwater floc: a functional

561

relationship of water and organic and inorganic floc constituents affecting suspended sediment

562

properties. Water Air Soil Pollut. 1997, 99 (1-4), 43-54.

563

(37) Aberle, J.; Nikora, V.; Walters, R. Effects of bed material properties on cohesive sediment

564

erosion. Mar. Geol. 2004, 207 (1), 83-93.

565

(38) Mouri, G.; Shiiba, M.; Hori, T.; Oki, T. Modeling reservoir sedimentation associated with an

566

extreme flood and sediment flux in a mountainous granitoid catchment, Japan. Geomorphology

567

2011, 125 (2), 263-270.

568

(39) Cnudde, V.; Boone, M. N. High-resolution X-ray computed tomography in geosciences: A

569

review of the current technology and applications. Earth-Sci. Rev. 2013, 123, 1-17. 28 ACS Paragon Plus Environment

Page 28 of 30

Page 29 of 30

Environmental Science & Technology

570

(40) Beven, K.; Germann, P. Macropores and water flow in soils. Water Resour. Res. 1982, 18

571

(5), 1311-1325.

572

(41) Jarvis, N. A review of non‐equilibrium water flow and solute transport in soil macropores:

573

Principles, controlling factors and consequences for water quality. Eur. J. Soil Sci. 2007, 58 (3),

574

523-546.

575

(42) Vlassenbroeck, J.; Dierick, M.; Masschaele, B.; Cnudde, V.; Van Hoorebeke, L.; Jacobs, P.

576

Software tools for quantification of X-ray microtomography at the UGCT. Nuclear Instruments

577

and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and

578

Associated Equipment 2007, 580 (1), 442-445.

579

(43) Vogel, H.-J.; Roth, K. Quantitative morphology and network representation of soil pore

580

structure. Adv. Water Resour. 2001, 24 (3), 233-242.

581

(44) Barnes, H. A.; Nguyen, Q. D. Rotating vane rheometry - a review. J. Nonnewton Fluid Mech.

582

2001, 98 (1), 1-14.

583

(45) Reed, A. H.; Boudreau, B. P.; Algar, C.; Furukawa, Y. Morphology of gas bubbles in mud: A

584

microcomputed tomographic evaluation; DTIC Document: 2005.

585

(46) Anderson, A.; Abegg, F.; Hawkins, J.; Duncan, M.; Lyons, A. Bubble populations and

586

acoustic interaction with the gassy floor of Eckernförde Bay. Cont Shelf Res. 1998, 18 (14),

587

1807-1838.

588

(47) Johnson, M.; Fairweather, M.; Harbottle, D.; Hunter, T. N.; Peakall, J.; Biggs, S. Yield stress

589

dependency on the evolution of bubble populations generated in consolidated soft sediments.

590

AIChE J. 2017, 63, 3728-3742. doi:10.1002/aic.15731.

591

(48) Barry, M. A.; Boudreau, B. P.; Johnson, B. D. Gas domes in soft cohesive sediments.

592

Geology 2012, 40 (4), 379-382.

29 ACS Paragon Plus Environment

Environmental Science & Technology

593

(49) Lin, H.; Bouma, J.; Wilding, L.; Richardson, J.; Kutilek, M.; Nielsen, D. Advances in

594

hydropedology. Adv. Agron. 2005, 85, 1-89.

595

(50) Roosevelt, S. E.; Corapcioglu, M. Y. Air bubble migration in a granular porous medium:

596

Experimental studies. Water Resour. Res. 1998, 34 (5), 1131-1142.

597

(51) Pierret, A.; Capowiez, Y.; Belzunces, L.; Moran, C. 3D reconstruction and quantification of

598

macropores using X-ray computed tomography and image analysis. Geoderma 2002, 106 (3),

599

247-271.

600

(52) Peth, S.; Horn, R.; Beckmann, F.; Donath, T.; Fischer, J.; Smucker, A. Three-dimensional

601

quantification of intra-aggregate pore-space features using synchrotron-radiation-based

602

microtomography. Soil Sci. Soc. Am. J. 2008, 72 (4), 897-907.

603

(53) Winterwerp, J. C.; Van Kesteren, W. G. Introduction to the physics of cohesive sediment

604

dynamics in the marine environment. Elsevier: 2004; Vol. 56.

605

(54) Gauglitz, P. A.; Buchmiller, W. C.; Probert, S. G.; Owen, A. T.; Brockman, F. J. Strong-

606

Sludge Gas Retention and Release Mechanisms in Clay Simulants; Pacific Northwest National

607

Laboratory (PNNL), Richland, WA (US): 2012.

608

(55) Sobek, S.; Zurbrügg, R.; Ostrovsky, I. The burial efficiency of organic carbon in the

609

sediments of Lake Kinneret. Aquat. Sci. 2011, 73 (3), 355-364.

610

(56) Lima, I. B.; Ramos, F. M.; Bambace, L. A.; Rosa, R. R. Methane emissions from large dams

611

as renewable energy resources: a developing nation perspective. Mitig. Aapt. Strat. Gl. 2008, 13

612

(2), 193-206.

613

(57) Ramos, F.; Bambace, L.; Lima, I.; Rosa, R.; Mazzi, E.; Fearnside, P. Methane stocks in

614

tropical hydropower reservoirs as a potential energy source. Clim. Change 2009, 93 (1-2), 1.

30 ACS Paragon Plus Environment

Page 30 of 30