Deconstructing Methane Emissions from a Small ... - ACS Publications

Oct 3, 2016 - GEOMAR Helmholtz Centre for Ocean Research Kiel, RD2 Marine ... Institute for Environmental Sciences, Environmental Physics, University ...
0 downloads 0 Views 1MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Deconstructing methane emissions from a small NorthernEuropean river: Hydrodynamics and temperature as key drivers Daniel F. McGinnis, Nicole Bilsley, Mark Schmidt, Peer Fietzek, Pascal Bodmer, Katrin Premke, Andreas Lorke, and Sabine Flury Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b03268 • Publication Date (Web): 03 Oct 2016 Downloaded from http://pubs.acs.org on October 8, 2016

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 29

Environmental Science & Technology

1

Deconstructing methane emissions from a small Northern-

2

European river: Hydrodynamics and temperature as key drivers

3

Daniel F. McGinnis1,2,3.*, Nicole Bilsley4, Mark Schmidt3, Peer Fietzek3,5, Pascal Bodmer2,6,

4

Katrin Premke2, Andreas Lorke8, Sabine Flury2 1

5 6 7

University of Geneva, Aquatic Physics, Department F.-A. Forel, Section of Earth and Environment Sciences, Faculty of Science, CH-1211 Geneva 4, Switzerland

2

Leibniz-Institute of Freshwater Ecology and Inland Fisheries (IGB), Chemical Analytics and

8 9

Biogeochemistry, 12587 Berlin, Germany 3

GEOMAR Helmholtz Centre for Ocean Research Kiel, RD2 Marine Biogeochemistry, 24148

10

Kiel, Germany 4

11

5

12 6

13 14

Scripps Institution of Oceanography, La Jolla, California 92093, USA

8

Kongsberg Maritime Contros GmbH, 24148 Kiel, Germany

Institute of Biology, Freie Universität Berlin, 14195 Berlin, Germany

Institute for Environmental Sciences, Environmental Physics, University of Koblenz-Landau,

15

76829 Landau, Germany

16

AUTHOR INFORMATION

17

Corresponding Author

18

* Corresponding Author, [email protected], +41 22 379 0792

19

University of Geneva, Uni Carl Vogt, 66 Blvd. Carl-Vogt, 1211 Geneva 4, Switzerland

20 21

KEYWORDS

22

Methane emissions, wetlands, river, climate change, ebullition, water flow rate, lateral exchange

1 ACS Paragon Plus Environment

Environmental Science & Technology

23

ABSTRACT

24

Methane (CH4) emissions from small rivers and streams, particularly via ebullition, are currently

25

under-represented in the literature. Here, we quantify the methane effluxes and drivers in a small,

26

Northern European river. Methane fluxes are comparable to those from tropical aquatic systems,

27

with average emissions of 320 mg CH4 m-2 d-1. Two important drivers of methane flux variations

28

were identified in the studied system: 1) temperature-driven sediment methane ebullition and 2)

29

flow-dependent contribution suspected to be hydraulic exchange with adjacent wetlands and

30

small side-bays. This flow-dependent contribution to river methane loading is shown to be

31

negligible for flows less than 4 m3 s-1, and greater than 50% as flows exceed 7 m3 s-1. While the

32

temperature - ebullition relationship is comparable to other systems, the flow rate dependency

33

has not been previously demonstrated. In general, we found that about 80% of the total emissions

34

were due to methane bubbles. Applying ebullition rates to global estimates for fluvial systems,

35

which currently are not considered, could dramatically increase emission rates to ranges from

36

lakes or wetlands. This work illustrates that small rivers can emit significant methane, and

37

highlights the need for further studies, especially the link between hydrodynamics and connected

38

wetlands.

39

2 ACS Paragon Plus Environment

Page 2 of 29

Page 3 of 29

Environmental Science & Technology

40

TEXT

41

INTRODUCTION

42

Including both seasonal and permanent waterbodies (lakes, reservoirs, rivers and wetlands),

43

earth’s land surface is covered with between ~6 to 15% freshwater.1 Recently, it was recognized

44

that these freshwaters actively process and transform received organic carbon (C).2, 3 Currently,

45

freshwaters are estimated to receive 3 Pg of C per year, of which about half (1.4 Pg C yr-1) is

46

emitted to the atmosphere as carbon dioxide (CO2) and methane (CH4). The remainder is either

47

transported to coastal oceans (~30%) or buried (~20%) in the freshwater sediments.3 In

48

freshwaters, the degree of C sequestration, degradation pathways and contribution to the

49

atmospheric greenhouse gas (GHG) budget are significantly impacted by sedimentation rates,4

50

temperature,3, 5, 6 hydrodynamics,7 and waterbody alterations (i.e. impoundments).8, 9

51

Emissions of methane from freshwaters are generally much lower than CO2, however

52

methane has 28 times higher global warming potential on a per mass basis.10 As methane is a

53

sparingly soluble gas (~27 times less soluble than CO2),11 it is estimated that about 50% of

54

freshwater methane emissions (excluding wetlands) are due to methane bubbles, which accounts

55

for ~10% of all global emissions.12, 13 Despite their importance for the global balance, ebullitive

56

emissions from aquatic systems are still likely underestimated due to the stochastic nature of

57

ebullition events.12, 14

58

While methane emission from lakes and reservoirs are increasingly investigated,8, 12, 15, 16

59

methane emission studies from streams17 and rivers,18 particularly via ebullition, are under-

60

represented in the literature.19, 20 Streams and rivers can be significant sources of methane,

61

however, and are particularly susceptible to both autochthonous and allochthonous carbon input

62

and exchange with their bordering environments,21 including groundwater22 and adjacent

3 ACS Paragon Plus Environment

Environmental Science & Technology

63

wetlands.19, 23 Recent estimates of streams and rivers place emissions at ~27 Tg CH4 yr-1,20

64

almost 20 times higher than the previous estimates.12 These estimates, however, do not include

65

ebullitive fluxes.20 The key challenges in accurately assessing C budgets and methane emissions

66

from these waterbodies are both the lack of data and measurement limitations. Methane

67

emissions, particularly via ebullition, are heterogeneous and stochastic and thus require a large

68

temporal and spatial coverage,14, 24 which is both time consuming and expensive.

69

Small rivers are abundant worldwide and represent commonly found aquatic systems25-27, and

70

there is a need to better assess the methane emissions, pathways and drivers. The goals of this

71

study are therefore to 1) quantify the overall methane emissions and pathways from a small

72

variably-flowing river located in a largely agricultural watershed, and 2) elucidate the key

73

emission drivers (i.e. temperature and flow rate) with system analytical approaches. Using these

74

results, we then 3) discuss the potential for future emissions under changing climate scenarios.

75 76 77

MATERIAL AND METHODS Study site. The Schwentine River (Figure S1), located in eastern Schleswig Holstein

78

(Germany) is ~70 km long and connects several primarily eutrophic lake-systems

79

(www.schleswig-holstein.de). The river is the main drainage system of these lakes to the Baltic

80

Sea and is dammed at its coastal outlet to the Kieler Fjord.28 The watershed is 726 km2 and

81

consists of about 85% arable land.29 Flow rates during the study ranged from 1.3 to 16 m3 s-1.

82

Water level was regulated so that fluctuations were less than ±3 cm [flow and water level data

83

from Schleswig-Holstein Agency for Coastal Defense, National Park and Marine Conservation].

84

Two upstream hydropower plants (Raisdorf 1 and 2) control the discharge into the lower reach of

85

the Schwentine River, which then flows about ~6.5 km to the Kieler Fjord. In the furthest

4 ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29

Environmental Science & Technology

86

downstream ~3 km, the shoreline is extensively lined with inundated reed stands, wetlands,

87

vegetation (estimated to cover >80% of the shore line), aquatic macrophytes, drainage ditches,

88

small connected ponds and side-bays (from here on, collectively called wetlands; Figure S1).

89

The downstream ~3 km stretch has an average width of ~35 meters and average depth of ~1 m.

90

Campaign overview. Dissolved surface water methane was measured along the lower

91

Schwentine River reach (see Figure S1). Between 9 and 10 discrete water (methane) samples

92

(see SI Methods) were collected while floating downstream with a canoe (termed “longitudinal

93

survey”). Six longitudinal surveys were performed in 2010 (21 July – 14 Sept) and 10

94

longitudinal surveys in 2011 (5 July – 5 Aug).

95

CH4 and CTD Sensors. Between 12 July to 28 Sept, 2011, a dissolved CH4 sensor (HydroC,

96

KM Contros GmbH, Germany) and CTD (conductivity-temperature-depth; CTD-XR420, RBR

97

Ltd., Canada) were mounted on a dedicated moored lander system at ~75 cm above sediment in

98

~1.5 meter water depth (See Fig 1. in Fietzek et al.30), which was deployed laterally just before

99

the discharge at the dam. CH4, temperature and conductivity were recorded every 1 minute. The

100

CH4 sensor was calibrated before and after the deployment for a range of up to 3 µmol L-1. The

101

data were post-processed to ensure data quality (See SI Methods).

102

Water surface methane fluxes. Surface CH4 fluxes were measured using floating chambers

103

in 2011 and 2013. Specifically, in 2011 surface flux measurements were performed August 17

104

and August 18 using four anchored chambers located between 50 and 100 m upstream from the

105

dam. The chamber gases were discretely sampled at time zero and every ~1 hour for

106

approximately 2 hours to determine the total flux (ebullitive + diffusive).see 8 From 2 – 5 July,

107

2013 fluxes were measured with a drifting chamber connected to a portable GHG analyzer

108

(UGGA; Los Gatos Research, Inc.; see SI Methods) covering the downstream 2000 meters of the

5 ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 29

109

river and could separately resolve diffusive and ebullitive fluxes.31, 32 Briefly, fluxes were

110

measured 30 times, and slopes obtained by determining the increase of chamber headspace CH4

111

over approximately 10 minutes.

112

Methane porewater sampling. Sediment porewater methane concentrations and fluxes were

113

analyzed after the method described in Maeck et al..8 Push cores ~20 cm in length were taken 2

114

km and 1 km upstream from the dam, and in the forebay near (~20 m) the dam in August 2011.

115

Fluxes were estimated with Ficks’s First Law (See SI Methods).

116 117

Mass Balance. The bubble dissolution contribution to the downstream methane loading was calculated with a mass balance on the downstream river stretch shown on Figure 1 where: 

118



=  −   +  −   +  

(kg d-1)

(1)

119

The left hand term is the change in methane mass (m, kg) with time (t). The masses denote the

120

various sources (+) and sinks (-) in kg d-1 from the sediment (sed), surface diffusive flux (diff),

121

and bubble dissolution (bub) within the water column. The average mass inflow and outflow of

122

CH4 (in kg d-1) were calculated ( =   ) using the campaign averaged discrete sampled

123

CH4 concentrations at 2.5 km upstream and at the dam, respectively (See Table S1).

124 125

Statistics. Statistical approaches are described in the Supporting Information (Methods). RESULTS

126

Results I – discrete sampling.

127

Sediment and surface fluxes. The average methane fluxes at the sediment-water interface

128

(SWI) from the three cores is FSED = 2.0 ± 1.1 mmol m-2 d-1 (32.7 ± 18 mg m-2 d-1) at ~19 oC

129

(Figure S2). Note that bubbles were observed in all cores, indicating gas oversaturation.

6 ACS Paragon Plus Environment

Page 7 of 29

Environmental Science & Technology

Figure 1. Average dissolved methane concentration (white dots) in the river as a function of distance from the dam (3,200 m) from the 17 longitudinal surveys (error bars represent standard deviation; see Figure S3). The dashed grey line shows the linear regression, which shows on average an increase of 0.16 nmol L-1 per meter of river distance (R2 = 0.95). Also shown are the measured mass inflow and outflow (Table S1), measured sediment fluxes, water surface diffusive (2013) and average ebullitive fluxes (2011 and 2013), and estimated bubble dissolution to close the mass balance. Where appropriate, values are reported in both kg d-1 for the mass balance and as fluxes (mmol m-2 d-1). 130 131

Methane emission measured with the anchored chamber (2011: ebullitive + diffusive) was

132

14.5 ± 20.4 mmol m-2 d-1 (232 ± 327 mg m-2 d-1; n = 28; average water temperature of 17.8 °C).

133

With the drifting chamber (2013) the ebullitive methane flux was 15.8 ± 17.7 mmol m-2 d-1 (253

134

± 283 mg m-2 d-1; n = 5; average water temperature of 18.2 °C), while the surface diffusive

135

methane flux was 1.4 ± 1.0 mmol m-2 d-1 (22.4 ± 16 mg m-2 d-1; n = 8). As the water velocity was

136

estimated to be 10 – 12 cm s-1, the anchored chambers may slightly bias the diffusive component

137

of the fluxes due to turbulence induced artifacts.33 The combined techniques, however, suggest

138

that diffusive fluxes are less than 10% of the total (ebullitive + diffusive) flux and the methods 7 ACS Paragon Plus Environment

Environmental Science & Technology

139

are comparable for determining total emissions. Using the average of the two methods gives an

140

ebullition estimate of ~15 mmol m-2 d-1.

141

Methane concentration along the river. The average (and standard deviation) of all 17

142

longitudinal surveys are shown on Figure 1. The average water discharge and temperature for the

143

surveys was 4.7 ± 1.6 m3 s-1 and 19.0 ± 2.2 oC. Individual survey results are shown on Fig S3

144

and are summarized Table S1 (Sample locations on Figure S1). In general, the methane increased

145

almost linearly approaching the dam and increased by about 100% along the measured river

146

stretch. Linear fits to the concentrations measured along an individual survey resulted in R2 of

147

0.9 – 0.99 (except transect 1 R2 = 0.38). The river was oxic during the measurements, with O2

148

values of 68-103% saturation (mean 91%) during longitudinal surveys. The mass methane inflow

149

(mCH4in = 3.8 kg d-1) and outflow (mCH4out = 6.5 kg d-1) are the averages from the longitudinal

150

transects and calculated as m = Q x CCH4 (See Table S1).

Figure 2. Methane concentration at the dam as a function of discharge (a) and temperature (b) and mass discharge of methane as a function of discharge (and temperature) (c). Plot colors indicate the water temperature (or discharge Q) when each data point was obtained. Lines on plot c) show linear fits to data for selected temperature ranges.

8 ACS Paragon Plus Environment

Page 8 of 29

Page 9 of 29

Environmental Science & Technology

151

Results II – time series. The average methane concentration at the dam discharge was 860

152

nmol L-1 (~260 times over saturated relative to atmospheric equilibrium) and covered a 6-fold

153

range (300 – 1800 nmol L-1) (12 July – 28 Sept, 2011; Figure 2a and 2b and Figure 3a). While

154

the methane concentration at the dam decreased with increasing flow, the rate of concentration

155

decrease was proportionately slower than the increase of the flow rate (Fig 2a). For example, as

156

the flow increased from 2 to 16 m3 s-1 (a factor of 8), the methane concentrations only decreased

157

from about 1800 to 500 nmol L-1 (a factor of 3.6). The water temperature ranged from ~14 – 22

158

o

159

factor of 10, from ~2 to 20 kg d-1 (125 – 1250 mol d-1) (Figure 2c).

C and was correlated with discharge (discussed below). Methane mass discharge varied by a

2000

P1 P2

Power-spectral density (-)

CH4 (nmol L-1)

(a) 1500 1000 500 22

0.03 (b) T Q CH4

0.02 0.01 0.00

10-1

18

Co-spectral density (-)

T (oC)

20

16 14

Q (m3 s-1)

16 12 8

0.8

(c)

Q - CH4

100

T-Q T - CH4

0.4 0.0 -0.4 -0.8

4

10-1

0 7/12/2011 7/26/2011 8/9/2011 8/23/2011 9/6/2011 9/20/2011

100

Frequency (day-1)

Figure 3 a) Time series of dissolved methane concentration (black is estimated sensor response time τ63= 15 min (flag 1), blue is estimated τ63= 25 min (flag 2); see SI Methods), temperature (T) and discharge (Q). P1 and P2 indicate the two contrasting flow regimes (see text). b) Variancepreserving power spectral density of temperature, discharge and dissolved CH4 concentration. c) Cospectral density of paired variables. For comparison, the spectra and co-spectra are normalized using the total variance and covariance, respectively.

9 ACS Paragon Plus Environment

Environmental Science & Technology

160 161

DISCUSSION Bubble dissolution and mass balance. Discrete measurements are summarized on Figure 1,

162

the only term to solve for is the internal loading due to the rising bubble dissolution.

163

Using equation 1 and assuming steady state, the bubble contributions, mbub is solved for as:

164

  = − +   −  +   = −3.8 + 6.5 − 3.4 + 1.9 = 1.2 kg d%& . The

165

mass balance indicates that 1.2 kg d-1 (Fbub = 0.7 mmol m-2 d-1) of internal methane loading is

166

due to bubble dissolution, or about 5% of the measured ebullitive surface flux (26 kg d-1). The

167

5% dissolution would result from a bubble diameter of 5 mm released from 1 m depth.34

168

Emissions – first estimate. Summing the total measured emissions presented on Figure 1,

169

this lower reach of the Schwentine river releases 34.4 kg d-1 of CH4 (~19 mmol m-2 d-1 or 310 mg

170

CH4 m-2 d-1) in total. Of this amount, about 8.4 kg d-1 (24%) escapes by diffusion and release at

171

the dam, and 26 kg d-1 (76%) escape as bubbles. While the above only provides a snapshot of

172

emissions, below we investigate temporal variability.

173

Methane drivers: temperature and flow rate. A time-series analysis was performed to

174

identify the roles of temperature and flow as drivers for internal methane loading and emissions.

175

The analysis covered two distinct discharge regimes named P1 and P2 before and after 13

176

August 2013, respectively (Figure 3a). This separation was chosen because P1 exhibited a lower

177

average discharge (6.2 ± 1.5 m3 s-1) with well-pronounced periodic variations in all parameters

178

(Q, T, and CH4) while P2 had a mean discharge nearly twice as large (11.2 ± 1.3 m3 s-1) with

179

more irregular and lower amplitude variations.

180

Considering both discharge regimes, CH4 concentration was positively linearly correlated

181

with temperature (R2 = 0.56, p < 0.001) and negatively correlated with discharge (R2 = -0.67, p
10 days which may reflect the time

193

scales of temperature change necessary to affect the methane oversaturated zone in the sediment

194

(diffusive length scale for 10 days for heat is ~34 cm).

195

System analysis. The dissolved methane concentrations in the downstream river stretch

196

increase in a linear manner (Figure 1). A similar trend was observed in run-of-river Lake

197

Wohlen16 and allows us to simplify a system-analysis approach by assuming spatially-constant,

198

zero-order inputs of the various methane sources (Figure 1). Using the longitudinal survey data,

199

the flow-dependent methane contribution (R) is separated from the temperature-driven ebullitive

200

fluxes and subsequent bubble dissolution FbubA (A is the sediment surface area) for each survey.

201

Equation 2 is solved to seek the methane loading FbubA + R (CH4loading, Table S1). In this

202

analysis, only the lower 2500 m stretch is consider as all surveys covered this range. Tracking a

203

1 x 1 m water parcel traveling down the river (see Figure S6 for model concept), the methane

204

concentration can be expressed as

205

' 

( = ) * − )  * + +)  * + ,-,

(mmol d-1, kg d-1)

11 ACS Paragon Plus Environment

(2)

Environmental Science & Technology

Page 12 of 29

206

where V is the downstream water volume (35 m3). On average, the total CH4loading from

207

equation 2 [FbubA + R] totals 2.8 kg d-1 (Table S1). The methane loading generally follows a

208

linear trend with flow rate (R2 = 0.77, Figure 4a). Two outliers (not included in the regression)

209

were observed (longitudinal surveys 1 and 2) when the flow was low (2 m3 s-1) and temperatures

210

high (>20°C) suggesting that the river was perhaps more resembling a (hot) lake under these

211

conditions. The outliers on Figure 4b generally followed periods of rapidly increasing flow rates

212

(Figure S5), potentially indicating more extreme flushing of side-bays and adjacent wetlands.

213

Additionally, we cannot rule out that high and rapidly accelerating flow rates may also disturb

214

the sediments. The temperature also predicts the loading well, but only for flows < 6 m3 s-1

215

(Figure 4b), and under-predicts the methane loading by a factor ~2 for flows > 6 m3 s-1. 6 (a)

(b)

11

14

14 2

2

2

10

10

1 0

R2 = 0.68

R2 = 0.77

3

-1

Q ≤ 6.0 m s 3 -1 Q > 6.0 m s

Surveys 3 - 17 Surveys 1 - 2

(c)

90 1

1

4 3

17

% Methane loading

Methane loading (kg d-1)

17

5

100

11

80

R² = 0.84

70 60 Ebullition R

50 40 30 20 10

1 2 3 4 5 6 7 8 910 15 20 25 3 -1 Discharge, Q (m s ) Temperature, T (°C)

0 2

3

4

5

6

7 3

Discharge, Q (m s-1)

Figure 4. Methane loading (FbubA + R) as a function of Q (a) and T (b). Red symbols were omitted from the regression analysis. Grey line on (b) is regression on data for Q ≤ 4.1 m3 s-1. (c) Fraction of loading estimated from bubble dissolution contribution and adjacent wetland fluxes. Curves are fits of 3rd degree polynomials. 12 ACS Paragon Plus Environment

Page 13 of 29

216

Environmental Science & Technology

Deconstructing methane sources. The flow-related contribution, R, and the bubble

217

contribution, FbubA, to the internal methane loading are estimated in the following steps:

218

1) The ebullition can be considered as a measure of methane production rate in the

219

sediment,35 and therefore we assume it can be described as a function of temperature. The

220

dissolution fraction of a single bubble however does not significantly vary over the

221

temperatures34 and was determined to be 5% for this system. The flow-driven methane

222

contribution appears minimal when flow rates are low, and a linear relationship between methane

223

loading and temperature shows an excellent fit (FbubA (kg d-1) = 0.39T – 5.08; R2 = 0.97, Fig 4b)

224

when Q < 4.1 m3 s-1, while including data with Q > 4.1 m3 s-1 decreases the correlation. The

225

temperature-ebullition loading equation is therefore used to estimate the temperature-driven

226

ebullition loading over all of the measured longitudinal surveys (Table S1, Fbub).

227

2) With the bubble dissolution contribution from step 1, the remaining loading is therefore

228

flow-driven R, which increases by ~1.35 kg d-1 per 1 m3 s-1 of flow increase above 5 m3 s-1 (R2 =

229

0.75).

230 231

With this approach the relative internal CH4 loading contribution is determined to reach a 50/50 split (bubble dissolution vs. flow-driven) at around 7 m3 s-1 (Figure 4c).

232

Flow dependency on methane loading. Methane loading and discharge were tightly coupled

233

with water flow rate (Figure 2c, 3c & 4b). To our knowledge, this has not been demonstrated for

234

similar systems. The sediment flux, Fsed, was similar to fluxes reported for other systems.8, 16

235

Note that the temperature differences measured at our site (range 15.8 – 24.4 °C) should only

236

affect the diffusive fluxes across the sediment-water interface by about ± 15% around the mean

237

(19°C).34, 36 However, as the water column is oxic, methane oxidizers at the sediment surface

238

likely suppress diffusive methane fluxes to the water column.37 Bubble release could be

13 ACS Paragon Plus Environment

Environmental Science & Technology

239

exacerbated by high or variable flows, as it has been shown to be affected by bottom shear and

240

pressure changes due to water level fluctionations.38 While we do not exclude this possibility, the

241

net bubble and the sediment flux should not increase substantially with increasing flow rate,

242

especially over longer timescales, and the hydrodynamics should not affect the rate of sediment

243

methane production in cohesive sediments.

244

The most likely explanation for this flow-dependent R term is the connection to the adjacent

245

wetlands (and submerged macrophytes, side-bays, etc.). The Schwentine River shoreline in the

246

final ~3 km reach is predominately submerged vegetation, stagnant side-bays as well as small,

247

quiescent streams (see photos, Figure S1). Melack et al.39 showed an increase of methane

248

emissions with increasing water level in areas of aquatic macrophytes in the Amazon basin.

249

Marin-Muniz et al.40 reported significantly higher methane emissions (more than double) during

250

the higher water levels (wet season) than during the dry season in tropical wetlands and swamps.

251

The authors attribute this to more reduced conditions in the sediment due to the increase in water

252

coverage of the land.

253

In the Schwentine River, the internal loading of methane due to bubble dissolution vs. the

254

presumed wetland exchange is ~50/50% at flows of 7 m3 s-1, with the wetland exchange as the

255

major contributor to internal methane loading at higher flows. Carbon sources fueling CO2 and

256

CH4 emissions from floodplain lakes and river channels have been attributed to input from

257

adjacent wetlands,25, 39, 41 and interaction between the hydrology, wetlands and rivers.42 Borges et

258

al.25 for example also observed an increase in CH4 concentrations with increasing wetlands

259

inundation in the catchment of sub-Saharan African rivers. Teodoru et al.43 report that the spatial

260

variability of GHGs on the Zambezi River (and tributaries) was related to the connectivity with

14 ACS Paragon Plus Environment

Page 14 of 29

Page 15 of 29

Environmental Science & Technology

261

wetlands and floodplains, with highest main-channel CH4 concentrations found just downstream

262

of extensive floodplains and wetland.

263

In addition to receiving suspended sediments from the main river, the autochthonous C input

264

into wetlands is reported to be high as these systems can be very productive.44 Quiescent areas

265

along rivers and impoundments are more likely to be productive, accumulate settled particulate

266

matter45 and are known to be methane ebullition hotspots.24 While the methane loading

267

dependence on flow rate appears robust in this study, it is unclear if there exists a hysteresis or

268

how different flow oscillation periods and amplitudes or sustained high flow rates may affect the

269

methane loading.

270

Temperature dependency on methane loading. The methane bubble emissions fit well with

271

the emission curve presented by DelSontro et al. (Figure 5).16 We report a Q10 of ~10 (the

272

relative change in a process over a 10 °C temperature change), which is generally higher than

273

those reported by others (e.g. ~4 by Yvon-Durocher et al.5), but the effect of temperature on

274

methanogenesis is still not well constrained.46 DelSontro et al. 47 defines an “ecosystem-level”

275

Q10 which accounts for additional in situ factors driving ebullition and is better representative of

276

actual emissions from freshwater ecosystems. The authors report Q10 for ebullition from shallow

277

ponds of 13, which is in the range of those measured in subarctic lakes (Q10 = 14).46, 47

278

Total CH4 emissions. The total emissions and internal loadings are summarized in Figure 1

279

for the average case for the longitudinal transects from measured data. Because the average flow

280

rate was low (4.8 ± 1.6 m3 s-1), the flow-dependent contribution R was likely near negligible for

281

that first estimate. However, the fraction of the internal methane loading due to dissolution of

282

rising methane bubbles has important implications for the overall emissions from shallow water

283

bodies. From the mass balance in Figure 1, only about 5% will dissolve in the water column,

15 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 29

284

leaving the remaining ~95% of the bubble methane to be emitted to the atmosphere. This rate of

285

dissolution corresponds to a 5 mm diameter bubble released from 1 m depth, and is in the size

286

range reported in the literature.8,14,34 Therefore, by solving for the amount of methane internal

287

loading due to released sediment bubble dissolution combined with bubble modeling16 allows

288

estimating the atmospheric bubble emissions.

Bubble emissions (mg m-2 d-1)

900 800 700 600 500 400

1

300 200

Surveys DelSontro 1 - mass balance 2 - 2013 3 - 2011

2 3

100 15

20

25 o

Water temperature ( C) Figure 5. Sediment bubble emissions from Table S1 as a function of temperature, and fit of ebullition vs. temperature from DelSontro et al. 16. The open symbols are (1) average emissions estimated from the longitudinal surveys, and (2) measured chamber fluxes from 2013 and (3) 2011. 289 290

With this assumption, the total CH4 ebullitive emission for the investigated stretch can be

291

estimated using the temperature relationship from Fig 4b. We estimate that the average bubble

292

emissions (Table S1, Bub.Emiss.) are ~33 kg d-1 (24 mmol m-2 d-1; 384 mg m-2 d-1; Tavg = 19 °C), 16 ACS Paragon Plus Environment

Page 17 of 29

Environmental Science & Technology

293

which is well within the range of our surface ebullitive flux measurements and fit well on the

294

emissions vs. temperature curve reported in DelSontro et al. (see Figure 5).16

295

The average methane exported at the dam is 6.6 kg d-1 (Table S1; Dam.CH4Diss.). Combining

296

the estimated ebullition emissions with the dam methane release implies a total mean emission of

297

40 kg d-1 from the investigated stretch, close to the initial 35 kg d-1 estimated from the measured

298

data (Figure 1). In percentage, about 17% is emitted through the dam, and 83% as bubbles to the

299

atmosphere.

300

The temperature equation (Figure 4b) can also be applied to the time series to estimate the

301

average internal methane loading due to bubble dissolution, and subsequent bubble emissions

302

(Figure 3). This results in an average internal bubble dissolution loading of 2.1 kg d-1 and a

303

subsequent ebullitive emission of 28 kg d-1 (Tavg =17.9 °C). For the same time series, it is known

304

that the mean dam methane discharge is 10 kg d-1, giving a total emissions of 38 kg d-1. In this

305

case the dam discharge is about 26% of the total and the bubble emissions are 74%. Nevertheless

306

the total emissions are close to the previous estimates.

307

Using the average of the three total emission estimates (flux chamber, system analysis, and

308

time series) gives an emission of 38 ± 3 kg d-1, or 340 mg m-2 d-1 expressed as an aerial rate over

309

the downstream 3.2 km (~112,000 m2). Average emissions for this small, northern-European

310

river stretch are comparable to tropical reservoirs.48 Though this estimate does not consider the

311

lower emissions in winter, even halving this emission rate is in the same range as tropical and

312

subtropical wetlands and swamps40 and those reported for Amazonian rivers in areas with

313

aquatic macrophytes (92 – 243 mg m-2 d-1 for low and high water, respectively).39 Expressing

314

half of our above average as a yearly value (3.8 mol m-2 yr-1), the emissions from the Schwentine

315

are on the upper-end of those reported by Bastviken et al..12 What is clear in this analysis is the

17 ACS Paragon Plus Environment

Environmental Science & Technology

316

importance of including ebullition in the river methane estimates. In our case, the total methane

317

emissions is ~4 times higher than the diffusive flux and outflow combined.

318

Finally, expressed as CO2 equivalents, the total methane emissions including all sources along

319

the investigated stretch are ~1000 kg CO2 d-1. This demonstrates that this small river, located in a

320

northern temperate zone, has areal methane emissions rates rivaling those from tropical rivers

321

and reservoirs.16, 52 Given the high abundance of small river systems worldwide, such rivers may

322

be much more significant components of the atmospheric carbon and methane budgets than

323

previously suspected.

324

An uncertain future. The Schwentine methane emissions are sensitive to climate parameters,

325

namely temperature and flow rate. The temperature influence on the methane concentration and

326

ebullition is not surprising and has been previously demonstrated.e.g. 49 DelSontro et al.16 reports

327

an exponential increase in bubble emission with increasing temperature in a run-of-river

328

reservoir, though their data were limited to temperatures less than 17oC. The exponential fit of

329

emissions vs. temperature provided by DelSontro et al., however, are in the range of the

330

ebullitive emissions reported in this study (Figure 5).

331

The increases in methane loading with both increasing temperature and flow rate suggests

332

that the emissions can be sensitive to both global warming and precipitation changes, though this

333

also depends on future nutrient and carbon loading that drive methane production. This is similar

334

to Campeau and Del Giorgio,49 who report that for their boreal stream network, GHG emissions

335

may increase by 13 – 68% over estimated climate scenarios spanning the next 50 years.

336

Temperatures in Northern Germany are forecasted to climb by 0.5 – 1.5 oC between 2021 –

337

2050.50 This has the potential to increase methane ebullitive emissions by about 3 – 8 mmol m-2

338

d-1 (15 – 40%). Precipitation is not expected to vary considerably (by about 10%) in the next

18 ACS Paragon Plus Environment

Page 18 of 29

Page 19 of 29

Environmental Science & Technology

339

decades, but forecasts suggest more extreme events, particularly during the crucial summer

340

period.51

341

Our findings underline the importance of ‘river-connected’ wetlands/floodplains and their

342

role in carbon turnover. Adjacent wetlands, side-bays, etc. create important sedimentation zones

343

in rivers, with the adjacent wetlands in the Schwentine River essentially trapping organic matter

344

and nutrients during higher flows, where it is then processed under probably more reduced,

345

anoxic conditions under low flow rates. The result is a system where the carbon turnover and

346

flux pathways are sensitive to both temperature and flow regimes. These results, while by no

347

means providing complete coverage of annual variations, illustrate the importance of combined

348

measurements and system analytical approaches to help determine and deconstruct the different

349

methane sources, and their drivers. Using these approaches, we could paint a picture of a water

350

body that is sensitive to future changes in climate in terms of methane emissions.

351 352

Acknowledgements

353

The authors would like to thank Mr. H. Kühl and Schwentientalfahrt for providing the boats and

354

infrastructure. The authors also thank the Geomar laboratory for their assistance with the

355

samples, particularly Ms. B. Domeyer. Special thanks to O. Ghamraoui for processing the sensor

356

data, A. Mäck and S. Geissler for assisting with the 2011 flux data. Mr. H.-J. Weber and the

357

Schleswig-Holstein Agency for Coastal Defense, National Park and Marine Conservation kindly

358

provided the flow and water level data.

359 360 361

19 ACS Paragon Plus Environment

Environmental Science & Technology

362

ASSOCIATED CONTENT

363

Supporting information

364

The supporting information includes supplementary methods, 1 table, and 6 ancillary figures,

365

including the map of the study site (Figure S1) and additional figures (Figures S2 – S6)

366

demonstrating the background conditions and results. This information is available free of charge

367

via the Internet at http://pubs.acs.org/.

368 369

Author Contributions

370

DFM conceived the study. NB measured transect data in 2010, 2011. PF installed and treated

371

methane and CTD time series data. SF, PB, KP, DFM organized and conducted field campaigns.

372

MS assisted with 2010, 2011 field campaigns, supervised chemical measurements, and provided

373

partial funding for 2010 and 2011 field work. PB, SF, AL, DFM analyzed flux and time series

374

data. DFM and SF wrote the manuscript and developed figures. All authors contributed to

375

discussion and input to the manuscript. All photographs were taken by SF, MS, and DFM.

376 377

Competing financial interests

378

There are no competing financial interests except for P. Fietzek, who is affiliated with

379

Kongsberg Maritime Contros, the manufacturer of the methane sensor deployed in 2011. We

380

independently validated the sensor measurements with gas chromatography analysis

381

at GEOMAR, and therefore we are confident of the objectivity of the results.

382 383 384

20 ACS Paragon Plus Environment

Page 20 of 29

Page 21 of 29

Environmental Science & Technology

385

Funding Sources

386

DFM was partially supported by the Leibniz-Institute of Freshwater Ecology and Inland

387

Fisheries (IGB) Fellowship Program in Freshwater Science. SF was founded by the Swiss

388

National Science Foundation Mobility Stipends (PBEZP2-129527 and PAOOP2-142041).

389

Further funding for the study was provided by the German Research Foundation (LO 1150/5-1).

390 391

References

392

1. Aufdenkampe, A. K.; Mayorga, E.; Raymond, P. A.; Melack, J. M.; Doney, S. C.; Alin, S. R.;

393

Aalto, R. E.; Yoo, K., Riverine coupling of biogeochemical cycles between land, oceans, and

394

atmosphere. Front. Ecol. Environ. 2011, 9, (1), 53-60.

395

2. Cole, J. J.; Prairie, Y. T.; Caraco, N. F.; McDowell, W. H.; Tranvik, L. J.; Striegl, R. G.;

396

Duarte, C. M.; Kortelainen, P.; Downing, J. A.; Middelburg, J. J.; Melack, J., Plumbing the

397

global carbon cycle: Integrating inland waters into the terrestrial carbon budget. Ecosystems

398

2007, 10, (1), 171-184.

399

3. Tranvik, L. J.; Downing, J. A.; Cotner, J. B.; Loiselle, S. A.; Striegl, R. G.; Ballatore, T. J.;

400

Dillon, P.; Finlay, K.; Fortino, K.; Knoll, L. B.; Kortelainen, P. L.; Kutser, T.; Larsen, S.;

401

Laurion, I.; Leech, D. M.; McCallister, S. L.; McKnight, D. M.; Melack, J. M.; Overholt, E.;

402

Porter, J. A.; Prairie, Y.; Renwick, W. H.; Roland, F.; Sherman, B. S.; Schindler, D. W.; Sobek,

403

S.; Tremblay, A.; Vanni, M. J.; Verschoor, A. M.; von Wachenfeldt, E.; Weyhenmeyer, G. A.,

404

Lakes and reservoirs as regulators of carbon cycling and climate. Limnol. Oceanogr. 2009, 54,

405

(6), 2298-2314.

406

4. Sobek, S.; DelSontro, T.; Wongfun, N.; Wehrli, B., Extreme organic carbon burial fuels

407

intense methane bubbling in a temperate reservoir. Geophys. Res. Lett. 2012, 39, L01401.

21 ACS Paragon Plus Environment

Environmental Science & Technology

408

5. Yvon-Durocher, G.; Allen, A. P.; Bastviken, D.; Conrad, R.; Gudasz, C.; St-Pierre, A.;

409

Nguyen, T.-D.; del Giorgio, P. A., Methane fluxes show consistent temperature dependence

410

across microbial to ecosystem scales. Nature 2014, 507, (7493), 488-491.

411

6. Yvon-Durocher, G.; Montoya, J. M.; Woodward, G.; Jones, J. I.; Trimmer, M., Warming

412

increases the proportion of primary production emitted as methane from freshwater mesocosms.

413

Global Change Biol. 2011, 17, (2), 1225-1234.

414

7. Battin, T. J.; Kaplan, L. A.; Findlay, S.; Hopkinson, C. S.; Marti, E.; Packman, A. I.;

415

Newbold, J. D.; Sabater, F., Biophysical controls on organic carbon fluxes in fluvial networks.

416

Nature Geoscience 2008, 1, (2), 95-100.

417

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

418

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

419

Technol. 2013, 47, (15), 8130-8137.

420

9. Weise, L.; Ulrich, A.; Moreano, M.; Gessler, A.; Kayler, Z.; Steger, K.; Zeller, B.; Rudolph,

421

K.; Knezevic-Jaric, J.; Premke, K., Water level changes affect carbon turnover and microbial

422

community composition in lake sediments. FEMS Microbiol. Ecol. 2016, 92, (5).

423

10.

424

to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge,

425

United Kingdom and New York, NY, USA, 2013; p 1535.

426

11.

427

Chem. Phys. 2015, 15, (8), 4399-4981.

428

12.

429

Methane Emissions Offset the Continental Carbon Sink. Science 2011, 331, (6013), 50.

IPCC Climate Change 2013: The Physical Science Basis. Contribution of Working Group I

Sander, R., Compilation of Henry's law constants (version 4.0) for water as solvent. Atmos.

Bastviken, D.; Tranvik, L. J.; Downing, J. A.; Crill, P. M.; Enrich-Prast, A., Freshwater

22 ACS Paragon Plus Environment

Page 22 of 29

Page 23 of 29

Environmental Science & Technology

430

13.

Dlugokencky, E. J.; Nisbet, E. G.; Fisher, R.; Lowry, D., Global atmospheric methane:

431

budget, changes and dangers. Phil. Trans. R. Soc. A 2011, 369, (1943), 2058-2072.

432

14.

433

large bubbles and small-scale hot spots for methane transport. Environ. Sci. Technol. 2015, 49,

434

(3), 1268-76.

435

15.

436

Methane Emissions from a Midlatitude Reservoir Draining an Agricultural Watershed. Environ.

437

Sci. Technol. 2014, 48, (19), 11100-11108.

438

16.

439

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

440

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

441

17.

442

stream (Sitka, Czech Republic). Limnologica - Ecology and Management of Inland Waters 2000,

443

30, (4), 359-366.

444

18.

445

potentials, pathways, and communities of methanogens in vertical sediment profiles of river

446

Sitka. Frontiers in Microbiology 2015, 6, 10.3389/fmicb.2015.00506.

447

19.

448

Ebullitive methane emissions from oxygenated wetland streams. Global Change Biol. 2014, 20,

449

(11), 3408-3422.

450

20.

451

The ecology of methane in streams and rivers: patterns, controls, and global significance. Ecol.

452

Monogr. 2016, 86, (2), 146 - 171.

DelSontro, T.; McGinnis, D. F.; Wehrli, B.; Ostrovsky, I., Size does matter: importance of

Beaulieu, J. J.; Smolenski, R. L.; Nietch, C. T.; Townsend-Small, A.; Elovitz, M. S., High

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

Rulík, M.; Čáp, L.; Hlaváčová, E., Methane in the hyporheic zone of a small lowland

Mach, V.; Blaser, M. B.; Claus, P.; Chaudhary, P. P.; Rulik, M., Methane production

Crawford, J. T.; Stanley, E. H.; Spawn, S. A.; Finlay, J. C.; Loken, L. C.; Striegl, R. G.,

Stanley, E. H.; Casson, N. J.; Christel, S. T.; Crawford, J. T.; Loken, L. C.; Oliver, S. K.,

23 ACS Paragon Plus Environment

Environmental Science & Technology

453

21.

Abril, G.; Deborde, J.; Savoye, N.; Mathieu, F.; Moreira-Turcq, P.; Artigas, F.; Meziane,

454

T.; Takiyama, L. R.; de Souza, M. S.; Seyler, P., Export of C-13-depleted dissolved inorganic

455

carbon from a tidal forest bordering the Amazon estuary. Estuarine Coastal and Shelf Science

456

2013, 129, 23-27.

457

22.

458

Identification of Methanogenic archaea in the Hyporheic Sediment of Sitka Stream. Plos One

459

2013, 8, (11), 10.1371/journal.pone.0080804.

460

23.

461

evasion from a temperate peatland stream. Limnol. Oceanogr. 2001, 46, (4), 847-857.

462

24.

463

Heterogeneity of Methane Ebullition in a Large Tropical Reservoir. Environ. Sci. Technol. 2011,

464

45, (23), 9866-9873.

465

25.

466

N.; Omengo, F. O.; Guerin, F.; Lambert, T.; Morana, C.; Okuku, E.; Bouillon, S., Globally

467

significant greenhouse-gas emissions from African inland waters. Nature Geosci 2015, 8, (8),

468

637-642.

469

26.

470

Kortelainen, P.; Striegl, R. G.; McDowell, W. H.; Tranvik, L. J., Global abundance and size

471

distribution of streams and rivers. Inland Waters 2012, 2, (4), 229-236.

472

27.

473

Outgassing from Amazonian rivers and wetlands as a large tropical source of atmospheric CO2.

474

Nature 2002, 416, (6881), 617-620.

Buriankova, I.; Brablcova, L.; Mach, V.; Dvorak, P.; Chaudhary, P. P.; Rulik, M.,

Hope, D.; Palmer, S. M.; Billett, M. F.; Dawson, J. J. C., Carbon dioxide and methane

DelSontro, T.; Kunz, M. J.; Kempter, T.; Wueest, A.; Wehrli, B.; Senn, D. B., Spatial

Borges, A. V.; Darchambeau, F.; Teodoru, C. R.; Marwick, T. R.; Tamooh, F.; Geeraert,

Downing, J. A.; Cole, J. J.; Duarte, C. M.; Middelburg, J. J.; Melack, J. M.; Prairie, Y. T.;

Richey, J. E.; Melack, J. M.; Aufdenkampe, A. K.; Ballester, V. M.; Hess, L. L.,

24 ACS Paragon Plus Environment

Page 24 of 29

Page 25 of 29

Environmental Science & Technology

475

28.

Marohn, L.; Prigge, E.; Hanel, R., Escapement success of silver eels from a German river

476

system is low compared to management-based estimates. Freshwat. Biol. 2014, 59, (1), 64-72.

477

29.

478

Plan. Provisional Management Plan Pursuant to the EU Water Framework Directive; Network

479

on the implementation of EU Water Framework Directive in the Baltic Sea Catchment: March

480

2006, 2006; p 144.

481

30.

482

and mobile platforms - Merging trends in the field of platform and sensor development. In

483

OCEANS 2011, 2011; pp 1-9.

484

31.

485

to dissolved organic matter quality across contrasting stream ecosystems. Sci. Total Environ.

486

2016, 553, 574-586.

487

32.

488

P.; Grossart, H.-P., Enhancing Surface Methane Fluxes from an Oligotrophic Lake: Exploring

489

the Microbubble Hypothesis. Environ. Sci. Technol. 2015, 49, (2), 873-880.

490

33.

491

Bastviken, D.; Flury, S.; McGinnis, D. F.; Maeck, A.; Müller, D.; Premke, K., Technical note:

492

drifting versus anchored flux chambers for measuring greenhouse gas emissions from running

493

waters. Biogeosciences 2015, 12, (23), 7013-7024.

494

34.

495

methane bubbles in stratified waters: How much methane reaches the atmosphere? Journal of

496

Geophysical Research-Oceans 2006, 111, (C9), 15.

Anonymous BERNET CATCH Regional Report: Schwentine River, Water Management

Fietzek, P.; Kramer, S.; Esser, D., Deployments of the HydroC (CO2/CH4) on stationary

Bodmer, P.; Heinz, M.; Pusch, M.; Singer, G.; Premke, K., Carbon dynamics and their link

McGinnis, D. F.; Kirillin, G.; Tang, K. W.; Flury, S.; Bodmer, P.; Engelhardt, C.; Casper,

Lorke, A.; Bodmer, P.; Noss, C.; Alshboul, Z.; Koschorreck, M.; Somlai-Haase, C.;

McGinnis, D. F.; Greinert, J.; Artemov, Y.; Beaubien, S. E.; Wuest, A., Fate of rising

25 ACS Paragon Plus Environment

Environmental Science & Technology

497

35.

Wilkinson, J.; Maeck, A.; Alshboul, Z.; Lorke, A., Continuous Seasonal River Ebullition

498

Measurements Linked to Sediment Methane Formation. Environ. Sci. Technol. 2015, 49, (22),

499

13121-13129.

500

36.

501

aqueous solutions. AICHE J. 1974, 20, (3), 611-615.

502

37.

503

oxide fluxes across sediment-water interface in a eutrophic lake. Chemosphere 2003, 52, (8),

504

1287-93.

505

38.

506

Environmental & Engineering Geoscience 2003, 9, (2), 167-178.

507

39.

508

Novo, E. M. L. M., Regionalization of methane emissions in the Amazon Basin with microwave

509

remote sensing. Global Change Biol. 2004, 10, (5), 530-544.

510

40.

511

from coastal freshwater wetlands in Veracruz Mexico: Effect of plant community and seasonal

512

dynamics. Atmos. Environ. 2015, 107, 107-117.

513

41.

514

Meziane, T.; Kim, J.-H.; Bernardes, M. C.; Savoye, N.; Deborde, J.; Souza, E. L.; Alberic, P.;

515

Landim de Souza, M. F.; Roland, F., Amazon River carbon dioxide outgassing fuelled by

516

wetlands. Nature 2014, 505, (7483), 395-398.

517

42.

518

Lambert, T.; Bouillon, S., Divergent biophysical controls of aquatic CO2 and CH4 in the World's

519

two largest rivers. Scientific Reports 2015, 5, 10.1038/srep15614.

Hayduk, W.; Laudie, H., Prediction of diffusion coefficients for nonelectrolytes in dilute

Liikanen, A.; Martikainen, P. J., Effect of ammonium and oxygen on methane and nitrous

Joyce, J.; Jewell, P. W., Physical controls on methane ebullition from reservoirs and lakes.

Melack, J. M.; Hess, L. L.; Gastil, M.; Forsberg, B. R.; Hamilton, S. K.; Lima, I. B. T.;

Marin-Muniz, J. L.; Hernandez, M. E.; Moreno-Casasola, P., Greenhouse gas emissions

Abril, G.; Martinez, J.-M.; Artigas, L. F.; Moreira-Turcq, P.; Benedetti, M. F.; Vidal, L.;

Borges, A. V.; Abril, G.; Darchambeau, F.; Teodoru, C. R.; Deborde, J.; Vidal, L. O.;

26 ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29

Environmental Science & Technology

520

43.

Teodoru, C. R.; Nyoni, F. C.; Borges, A. V.; Darchambeau, F.; Nyambe, I.; Bouillon, S.,

521

Dynamics of greenhouse gases (CO2, CH4, N2O) along the Zambezi River and major tributaries,

522

and their importance in the riverine carbon budget. Biogeosciences 2015, 12, (8), 2431-2453.

523

44.

524

Turnover in a Freshwater Marsh. Appl. Environ. Microbiol. 2006, 72, (1), 596-605.

525

45.

526

retention in the Iron Gate I reservoir on the Danube River: The role of side bays as nutrient.

527

River Res. Appl. 2006, 22, (4), 441-456.

528

46.

529

ebullition and diffusion from northern ponds and lakes regulated by the interaction between

530

temperature and system productivity. Limnol. Oceanogr. 2016, 10.1002/lno.10335.

531

47.

532

input is primary controller of methane bubbling in subarctic lakes. Geophys. Res. Lett. 2014, 41,

533

(2), 555-560.

534

48.

535

surfaces as sources of greenhouse gases to the atmosphere: A global estimate. Bioscience 2000,

536

50, (9), 766-775.

537

49.

538

rivers: Major drivers and implications for fluvial greenhouse emissions under climate change

539

scenarios. Global Change Biol. 2014, 20, (4), 1075-1088.

540

50.

541

the Environment, Nature Conservation and Nuclear Safety (BMU): Berlin, Germany, 2009.

Buesing, N.; Gessner, M. O., Benthic Bacterial and Fungal Productivity and Carbon

McGinnis, D. F.; Bocaniov, S.; Teodoru, C.; Friedl, G.; Lorke, A.; Wuest, A., Silica

DelSontro, T.; Boutet, L.; St-Pierre, A.; del Giorgio, P. A.; Prairie, Y. T., Methane

Wik, M.; Thornton, B. F.; Bastviken, D.; MacIntyre, S.; Varner, R. K.; Crill, P. M., Energy

St Louis, V. L.; Kelly, C. A.; Duchemin, E.; Rudd, J. W. M.; Rosenberg, D. M., Reservoir

Campeau, A.; Del Giorgio, P. A., Patterns in CH4 and CO2 concentrations across boreal

BMU Combating climate change. The German adaptation strategy.; Federal Ministry for

27 ACS Paragon Plus Environment

Environmental Science & Technology

542

51.

Feldmann, H.; Schädler, G.; Panitz, H.-J.; Kottmeier, C., Near future changes of extreme

543

precipitation over complex terrain in Central Europe derived from high resolution RCM

544

ensemble simulations. International Journal of Climatology 2013, 33, (8), 1964-1977.

545

52.

546

Krusche, A. V.; Snidvongs, A., Physical controls on carbon dioxide transfer velocity and flux in

547

low-gradient river systems and implications for regional carbon budgets. Journal of Geophysical

548

Research: Biogeosciences 2011, 116, (G1), 10.1029/2010JG001398.

Alin, S. R.; de Fátima F. L. Rasera, M.; Salimon, C. I.; Richey, J. E.; Holtgrieve, G. W.;

549 550

28 ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29

Environmental Science & Technology

TOC Art 64x47mm (300 x 300 DPI)

ACS Paragon Plus Environment