+ All Categories
Home > Documents > Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic...

Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic...

Date post: 16-Aug-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
51
Carbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen 1,* , Melanie Bergmann 2 , Karline Soetaert 1 , Eduard Bauerfeind 2 , Christiane Hasemann 2 , Michael Klages 2 , Ingo Schewe 2 , Thomas Soltwedel 2 , Nataliya E. Budaeva 3 1 Centre for Estuarine and Marine Ecology, Netherlands Institute of Ecology (NIOO-KNAW), PO Box 140, 4400 AC Yerseke, The Netherlands 2 Alfred Wegener Institute for Polar and Marine Research, Am Handelshafen 12, D-27570 Bremerhaven, Germany 3 P.P. Shirshov Institute of Oceanology, Russian Academy of Sciences, Nakhimovsky Pr., 36, 117997 Moscow, Russia * Corresponding author: [email protected] 1 2 3 4 5 6 9 10 11 12 13 14 15 16
Transcript
Page 1: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

Carbon flows in the benthic food web at the deep-sea

observatory HAUSGARTEN (Fram Strait)

Dick van Oevelen1,*, Melanie Bergmann2, Karline Soetaert1, Eduard Bauerfeind2, Christiane

Hasemann2, Michael Klages2, Ingo Schewe2, Thomas Soltwedel2, Nataliya E. Budaeva3

1 Centre for Estuarine and Marine Ecology, Netherlands Institute of Ecology (NIOO-KNAW), PO Box

140, 4400 AC Yerseke, The Netherlands

2 Alfred Wegener Institute for Polar and Marine Research, Am Handelshafen 12, D-27570

Bremerhaven, Germany

3 P.P. Shirshov Institute of Oceanology, Russian Academy of Sciences, Nakhimovsky Pr., 36, 117997

Moscow, Russia

* Corresponding author: [email protected]

1

2

3

4

5

8

9

10

11

12

13

14

15

Page 2: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

ABSTRACT

The HAUSGARTEN observatory is located in the eastern Fram Strait (Arctic Ocean) and used as

long-term monitoring site to follow changes in the Arctic benthic ecosystem. Linear inverse modelling

was applied to decipher carbon flows among the compartments of the benthic food web at the central

HAUSGARTEN station (2500 m) based on an empirical data set consisting of data on biomass,

prokaryote production, total carbon deposition and community respiration. The model resolved 99

carbon flows among 4 abiotic and 10 biotic compartments, ranging from prokaryotes up to megafauna.

Total carbon input was 3.78±0.31 mmol C m-2 d-1, which is a comparatively small fraction of total

primary production in the area. The community respiration of 3.26±0.20 mmol C m-2 d-1 is dominated

by prokaryotes (93%) and has lower contributions from surface-deposit feeding macro- (1.7%) and

suspension feeding megafauna (1.9%), whereas contributions from nematode and other macro- and

megabenthic compartments were limited to <1%. The high prokaryotic contribution to carbon

processing suggests that functioning of the benthic food web at the central HAUSGARTEN station is

comparable to those of abyssal plain sediments that are characterised by strong energy limitation.

Faunal diet compositions suggest that labile detritus is important for deposit-feeding nematodes (24%

of their diet) and surface-deposit feeding macrofauna (~44%), but that semi-labile detritus is more

important in the diets of deposit-feeding macro- and megafauna. Dependency indices on these food

sources were also calculated as these integrate direct (i.e. direct grazing and predator – prey

interactions) and indirect (i.e. longer loops in the food web) pathways in the food web. Projected sea-

ice retreats for the Arctic Ocean typically anticipate a decrease in the labile detritus flux to the already

food-limited benthic food web. The dependency indices indicate that faunal compartments depend

similarly on labile and semi-labile detritus, which suggests that the benthic biota may be more

sensitive to changes in labile detritus inputs than when assessed from diet composition alone.

Species-specific responses to different types of labile detritus inputs, e.g. pelagic algae versus

sympagic algae, however, are presently unknown and are needed to assess the vulnerability of

individual components of the benthic food web.

Keywords: Food web – Modelling – Sediment – Benthos – Arctic Ocean – Carbon processing

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

Page 3: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

1. Introduction

The Earth is warming rapidly due to anthropogenic inputs of CO2 into the atmosphere (IPCC,

2007). While research is mainly directed at the terrestrial consequences of global warming the

changes in the deep oceans, especially those in the vulnerable Polar regions receive less attention.

Climate change is expected to affect Arctic marine ecosystems in various direct and indirect ways.

One direct effect is that seawater temperatures will rise and this will directly affect organisms

physiology (Pörtner et al., 2001). However, observed temperature changes in the deep Arctic ocean

are still limited to <0.01°C y-1 (Glover et al., 2010). A more profound and faster impact is to be

expected through an indirect mechanism: the retreat of the ice-edge and the continuous loss of multi-

year ice will lead to a decreased flux of fast-sinking sympagic algae and fauna (Hop et al., 2006). The

dominant primary producers in the upper water column may therefore shift from sympagic algae to

pelagic phytoplankton, which may be retained in the twilight zone (Buesseler et al., 2007). This change

could shift an ecosystem characterized by strong benthic-pelagic coupling to one characterized by a

water column – dominated food web (Grebmeier et al., 2006; Hop et al., 2006).

It may not be easy to detect changes in the quantity and composition of primary producers in

the upper water column directly because algal blooms and ice cover are erratic and difficult to sample

at appropriate temporal resolution (Bauerfeind et al., 2009; Forest et al., 2010). The benthic

ecosystem, which depends directly on phytodetritus produced in the euphotic zone and which

integrates patterns in the overlying productivity over longer time periods, may yield a more consistent

signal.

In this context, the Alfred Wegner Institute for Polar and Marine Research (Germany)

established the deep-sea observatory HAUSGARTEN west of Svalbard (Soltwedel et al., 2005) to

provide a long-term monitor of changes in the Arctic benthic ecosystem (Fig. 1). The observatory

comprises nine sampling stations along a bathymetric transect (1000 – 5500 m). A latitudinal transect

crosses the bathymetric transect at the central HAUSGARTEN station (2500 m), which serves as an

experimental area for biological long-term experiments (Gallucci et al., 2008; Kanzog et al., 2009).

Repeated sampling and deployments of moorings and long-term landers has been conducted on an

annual basis since 1999 and has yielded a unique time-series dataset on mega-, macro- and

meiobenthic, prokaryotic, biogeochemical and geological properties as well as on hydrography and

sedimentation patterns (Bauerfeind et al., 2009; Bergmann et al., 2009; Hoste et al., 2007). This time-

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

Page 4: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

series has revealed decreases in the proportions of fresh phytodetrital matter at the seafloor and in the

concentration of sediment-bound organic matter in the period 2001 – 2005 (Soltwedel et al., 2005).

Changes in the quality and quantity of detrital input can affect the structure of the benthic food

web profoundly (Billett et al., 2010; Ruhl et al., 2008; Smith et al., 2009). Indeed, Hoste et al. (2007)

showed a decline in the microbial biomass of sediments and changes in nematode community

structure at HAUSGARTEN. These changes, however, operate in a food web context, in which biota

are linked through consumption and predation processes. Data sets on benthic food webs are typically

restricted to biomass estimates of large functional groups and occasional rate measurement (Soetaert

and Van Oevelen, 2009), rendering knowledge based on field measurements alone insufficient to

derive a coherent picture of carbon flows in these systems. Recent advances in the use of so-called

‘inverse modelling’ techniques, however, enable us not only to quantify food web flows based on

limited data sets, but also to assess the uncertainty associated with this quantification (Van Oevelen et

al., 2010). These techniques allow us to analyse even complex deep-sea food webs quantitatively

(Van Oevelen et al., 2009). The basic advantage is that site-specific field data on carbon processing

and carbon biomass are combined with more uncertain data from the literature to collectively constrain

the magnitudes of the food web flows.

In this paper, we combine the comprehensive set of available empirical data to quantify the

carbon flows in the benthic food web of the central HAUSGARTEN station (2500 m). Detrital input to

the food web is divided into three classes of lability to do justice to the heterogeneity of natural detritus

and assess differences in diet contributions of these different detritus classes. Moreover, we determine

partitioning of respiration and secondary production to identify which food web compartments are

important pathways in the benthic food web. Trophic levels of the faunal compartments are calculated

and compared with trophic level position based on δ15N isotope data (Bergmann et al., 2009), to verify

the resulting food web structure. Finally, dependency indices of biotic compartments on the basal

detritus and prokaryotic resources are calculated. Dependency indices quantify the dependence of a

biotic compartment on other compartments via direct (i.e. consumption) and indirect (transfer via

longer pathways) interactions (Ulanowicz, 2004). The model results will be used to speculate on

changes that can be anticipated in the benthic food web under a scenario of receding sea ice.

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

Page 5: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

2. Material and methods

2.1 Data collection

An overview of the field data and references that were used in the food web model is given in

Table 1 and 2, with a brief summary of the sampling methodology given here. Most samples were

taken during expedition ARK XIX/3c (July–August 2003) with the German research ice breaker

Polarstern at the central HAUSGARTEN station (2500 m water depth).

The deposition of particulate organic carbon (POC) represents an important input parameter of

the inverse model, since it determines the total carbon processing by the benthic food web. Long-term

deployments of deep sediment traps provide important constraints on the POC input. The sedimenting

particles were sampled by modified automatic Kiel sediment traps (see Bauerfeind et al., 2009 for

details). The sediment traps were installed in bottom-tethered moorings at different depths, but here

only the data from the deepest sediment trap (170 mab) are considered. The traps were programmed

to collect at 15 day intervals. POC input data from the productive spring-summer season are used in

the model, since this depositional flux is most relevant for the benthic food web compartments that

were sampled in July 2003. The collector cups were filled with sterile water, adjusted to a salinity of 40

and poisoned with mercury chloride (0.14% final solution) and kept refrigerated till further processing

after recovery. Sub-samples were analyzed for, amongst other parameters, particulate organic carbon

(see Bauerfeind et al., 2009 for details). POC deposition showed little variation during March – May

(range 0.77 to 1.44 mmol C m-2 d-1), but rose substantially to 3.99 mmol C m-2 d-1 in June (Fig. 2).

Because of this range in deposition rates, it was decided to include the full range as input for the

model (Table 2).

Sediment samples were taken by a multiple corer and the top 5 cm of the sediment analyzed

at 1-cm intervals for organic carbon content, pigment concentration, prokaryotic biomass, hydrolytic

activity, meio- and macrofaunal biomass. Sediment porosity was estimated by measuring the weight

loss of wet sediment samples dried at 60 °C (average of 0.60 for the top 5 cm). Total organic carbon

content was determined as the ash-free dry weight after combustion and converted to total organic

carbon in the sediment using sediment porosity and assuming a density of 2.5 g cm-3 for the sediment

fraction. Particulate proteins were analyzed photometrically following 0.5 N NaOH extraction (Greiser

and Faubel, 1988). Chloroplastic pigments were extracted and the chlorophyll a content was

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

131

Page 6: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

determined with a fluorometer. Prokaryotic cell volume was determined with the Porton grid

(Grossmann and Reichardt, 1991) after staining with acridine orange and converted to prokaryotic

biomass using a conversion factor of 3.0×10−13 g C μm−3 (Borsheim et al., 1990). Prokaryotic

enzymatic turnover rates were measured as an indicator of the potential hydrolytic activity of

prokaryotes using fluorescein-di-acetate as fluorogenic substrate (Köster et al., 1991), hydrolysis rates

were converted to carbon units assuming that one mole fluorescein is equivalent to four moles of

carbon (i.e. 2 acetate molecules). Sediment samples for nematode enumeration were sieved through

a 1 mm sieve and nematodes that were retained on a 32 μm were extracted by Ludox centrifugation

(Hoste et al., 2007). Macrofaunal density estimates were based on box-core samples (Budaeva et al.,

2008). Megafaunal density estimates were acquired by analysis of still images of the seafloor taken by

a towed camera system during RV Polarstern expedition ARK XVIII/1 in 2002 (Soltwedel et al., 2009).

Total oxygen uptake by the benthic community was determined from the decrease in oxygen

concentration in sediment cores that were incubated in situ during the RV Polarstern cruise ARK XVI/2

(2002) and RV Maria S. Merian expedition 2-4 (2006) (Winkler titration were done on-board)

(Soltwedel, unpublished).

2.2 Food web model

The food web model was set up as a linear inverse model (LIM). The term linear refers to the

food web model being described as a linear function of the flows, inverse means that the food web

flows are derived from observed data. The model itself is the topology of the food web, which is

determined a priori by delineating the compartments and connecting them with flows.

Several reviews on linear inverse modelling have been recently published and contain simple

models to exemplify the setup and solution of benthic food web LIMs (Soetaert and Van Oevelen,

2009; Van Oevelen et al., 2010). Here, we therefore limit our methodological discussion on linear

inverse models. A LIM contains a mass balance for each food web compartment and a set of

quantitative data constraints. A LIM is captured by two matrix equations:

Equality equation: Ax=b (1)

Inequality equation: Gx≥h (2)

in which vector x contains the unknown flows. Each row in the equality equation (1) imposes a strict

constraint: a linear combination of the flows must match the corresponding value in vector b. The

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

Page 7: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

inequality equation (2) imposes lower and upper bounds on flows or on linear combinations of flows. A

default set of inequalities is the condition x≥0, which ensures that flows have directions that are

consistent with the imposed food web topology.

For the HAUSGARTEN station, the compartments of the benthic food web were defined as: labile

detritus (lDet), semi-labile detritus (sDet), refractory detritus (rDet), dissolved organic carbon (DOC),

prokaryotes (Pro), deposit-feeding nematodes (NemDF), predatory+omnivore nematodes (NemPO),

surface deposit-feeding macrofauna (MacSDF), deposit-feeding macrofauna (MacDF), suspension-

feeding macrofauna (MacSF), predatory+scavenging macrofauna (MacPS), deposit-feeding

megafauna (MegDF), suspension-feeding megafauna (MegSF) and predatory+scavenging megafauna

(MegPS).

Carbon stocks were available for all compartments, except DOC (Table 1). Labile detritus was defined

as all carbon associated with chlorophyll a. Chlorophyll a concentrations were summed in the top 5 cm

and were converted to carbon units by assuming a carbon to chlorophyll a ratio of 40 that is typical for

diatoms (Allen et al., 2005). Semi-labile detritus was defined as the carbon equivalents of particulate

proteins (converted to carbon equivalents by the conversion factor 0.49, Pusceddu et al., 2010) in the

top 5 cm (Hoste et al., 2007) minus the labile detritus stock. Refractory detritus was defined as the

total OC stock in the top 5 cm of the sediment minus the labile and semi-labile detritus stocks.

Prokaryotic carbon stocks were inferred from cell volumes (see above). The biomass of nematodes (>

85% of the meiobenthic community, Hoste et al., 2007) was partitioned among feeding modes based

on the following nematode feeding types (Wieser, 1953): deposit-feeding nematodes (Wieser type 1A,

1B and 2A) and predatory+omnivore nematodes (Wieser type 2B). Macrobenthic and megabenthic

species were divided into feeding types using specialized literature, natural abundance stable isotope

values (Bergmann et al., 2009) and expert judgement.

Carbon inputs into the food web are deposition and/or feeding on suspended labile (lDet_w),

semi-labile (sDet_w) and refractory detritus (rDet_w). Carbon outputs from the food web are

respiration to dissolved inorganic carbon (DIC), burial of rDet, DOC efflux to the water column and

export by the macro- and megafaunal compartments (e.g. consumption by fish).

Within the food web, the labile and semi-labile detritus pools in the sediment can be

hydrolysed to DOC, or are grazed upon by nematodes (NemDF and NemPS) and MacSDF, MacDF,

MacPS, MegSDF, MegDF and MegPS. Refractory detritus is only hydrolysed to DOC. The DOC is

161

162

163

164

165

166

167

168

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

Page 8: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

taken up by prokaryotes or effluxes to the water column. Predatory feeding links are primarily defined

based on size class; prokaryotes are consumed by the nematode and non-suspension-feeding macro-

and megafaunal compartments, deposit-feeding nematodes are consumed by predatory nematodes,

both nematode compartments are consumed by non-suspension-feeding macro- and megafaunal

compartments, the macrofaunal compartments MacSDF, MacDF and MacSF are preyed upon by

predatory macro- and megafauna and predatory macrofauna is predated upon by predatory

megafauna.

Part of the sources ingested by the faunal compartments is not assimilated but instead

expelled as faeces. The non-assimilated labile (e.g. labile detritus, prokaryotic and faunal

compartments) and semi-labile (semi-labile detritus) carbon enter the semi-labile and refractory

detritus, respectively. Respiration by faunal compartments is defined as the sum of maintenance

respiration (biomass-specific respiration) and growth respiration (overhead on new biomass

production). Prokaryotic mortality is defined as a flux to DOC and faunal mortality is defined as a flux

to labile detritus.

2.3 Data constraints

The range in POC fluxes, as measured with deep sediment traps, was included in the

inequality equation (Table 2). There were two measurements of sediment oxygen consumption rates

and these were quite variable, and were therefore also included in the inequality equation (Table 2).

Esterase activity reflects potential hydrolysis rates rather than in situ hydrolysis rates (Gumprecht et

al., 1995) and the measured hydrolysis rate was therefore imposed as upper bound on total hydrolysis

(Table 2).

In addition to the site-specific data, a set of general constraints from the literature were

included in the inequality equation. These constraints were used to set bounds on degradation rates of

the labile, semi-labile and refractory detritus pools, burial efficiency, prokaryotic growth efficiency, viral-

induced prokaryotic lysis, release of DOC from the sediment, grazing of prokaryotes by nematodes,

assimilation efficiency of all faunal compartments, net growth efficiency of all faunal compartments,

production and mortality rates of all faunal compartments (Table 2). The biomass-specific production

and mortality rates in combination with the biomass values of the faunal stocks constrain the total

carbon demand by the faunal compartments. Since measurements of assimilation and growth

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

216

217

218

219

Page 9: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

efficiencies of deep-sea benthos are very rare, an extensive literature review (Van Oevelen et al.,

2006) of temperate benthos was used as basis for these constraints. Assimilation efficiencies for semi-

labile carbon were set to half the values of the assimilation efficiencies of labile carbon for the macro-

and megafaunal compartments. Faunal maintenance respiration was defined as 0.01 d-1 at 20°C (see

references in Van Oevelen et al., 2006) and is corrected with a temperature-correction factor (Tlim)

based on the Q10 formulation with a doubling of rates for every 10°C increase (Table 2). The bottom

water temperatures at HAUSGARTEN were ca. -0.8°C.

Both surface-deposit and deposit-feeding holothurians and other echinoderms ingest organic

matter with higher than ambient chlorophyll a and total hydrolysable amino acid concentrations

(Ginger et al., 2001; Witbaard et al., 2001), although selectivity differs between feeding modes with

surface-deposit feeders typically exhibiting stronger selectivity than deposit feeders (Wigham et al.,

2003). Selectivity between labile detritus and semi-labile detritus for these organisms was defined as

the ratio of chlorophyll a concentrations in the gut with respect to the ambient surface sediment. The

level of selectivity varies from 1 to 10 for deposit feeding holothurians at the Porcupine Abyssal Plain

to >500 for the surface-deposit-feeding holothurian Amperima rosea (Wigham et al., 2003). Selectivity

at the Antarctic peninsula was less evident (selectivity of 2 to 7), possibly because of the existence of

a food bank, but there was a clear separation between deposit and surface-deposit feeders (Wigham

et al., 2008). Therefore, zero to moderate (1 to 10) selectivity for deposit feeders and strong selectivity

(50 to 100) for surface deposit feeders was assumed in the model (Table 2). Since no comparable

data are available for macrofauna, similar selectivity ranges were defined for these communities

(Table 2). Finally, the predatory nematodes and macro- and megafaunal compartments were assumed

to ingest a minimum of 75% through predatory feeding (Table 2).

2.4 Model solution

The complete food web model consists of 99 flows, 16 compartments and mass balances, 99

inequalities of ≥0 and 123 data inequalities. It is clear that the total number of flows in a food web

greatly outnumbers the equations in the LIM (99≫16). As a result, a food web LIM is mathematically

under-determined, which implies that an infinitely large set of solutions fits the matrix equations. Since

no unique solution can be found for an under-determined model, a recently developed likelihood

approach was followed (Van den Meersche et al., 2009; Van Oevelen et al., 2010). In short, a large

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

235

236

237

238

239

240

241

242

243

244

245

246

247

248

Page 10: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

set of 50,000 solutions is sampled from the infinitely large set of solutions. Each solution represents a

different food web configuration and is consistent with the matrix equations Ax=b and Gx≥h. The

mean and standard deviation for each food web flow is calculated from this set of sampled solutions

and represents a central estimate (i.e. the mean) of the flow value and its associated uncertainty (i.e.

standard deviation) (Van Oevelen et al., 2010). This will be noted as mean ± standard deviation.

Trophic levels of the biotic compartments and dependency indices were calculated for each solution in

the set of 50,000 solutions using the R-package NetIndices (Kones et al., 2009). By running the model

50,000 times, the uncertainty in the empirical data (indicated by the flow ranges in Table 2) is

propagated onto an uncertainty estimate of the carbon flows as indicated by its standard deviation.

Convergence of the mean and standard deviation of the flows was checked visually to confirm that the

set of 50,000 model solutions was sufficiently large. Generally, model convergence (within 10% of the

final mean and standard deviation for each flow value) was achieved after <5,000 solutions. In the

calculation of trophic levels, the three detritus and dissolved organic carbon compartments were fixed

to a trophic level of one. The model code is made available in the R-package LIM (Soetaert and Van

Oevelen, 2008).

2.5 Sensitivity analysis

The data set included in the inverse model is inherently uncertain. The uncertainty of the flux data and

rate parameters is included in the model by incorporating them as lower and upper bounds on their

values (Table 2). In this way, this uncertainty propagates onto the final model solution as standard

deviation for each flow value (see description of sampling methodology above). The stock data,

however, are also uncertain, but this uncertainty cannot be directly included by lower and upper

bounds using this sampling methodology. This is because, with a perturbation of the stock inputs, the

core equations Ax=b and Gx≥h are not guaranteed to be valid when all solutions are averaged to

obtain the final model solution. Henceforth, a sensitivity analysis was performed in which the stock

values were perturbed one-by-one by increasing or decreasing a stocks value with 15% of its default

value. With the perturbed stock value a new set of 500 solutions was sampled. The number of 500

was chosen to save computing time, while at the same time it was large enough to reasonably

approach the final model solution. The set of solutions was subsequently averaged to obtain a

249

250

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

Page 11: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

perturbed model solution. These perturbed model solutions were compared with the default model

solution to assess the sensitivity of our model results for changes in the stock values.

3. Results

A complete overview of the mean and standard deviation for each food web flow is given in

the Appendix.

3.1 Carbon flows inferred by the inverse model

Total carbon input to the food web was 3.78±0.31 mmol C m-2 d-1 and is partitioned among

labile detritus deposition (30%), semi-labile detritus deposition (31%), refractory detritus deposition

(32%) and suspension feeding (8%). Total respiration is 3.26±0.20, burial is 0.32±0.08 and export from

the food web is 0.02±0.006 mmol C m-2 d-1. Total respiration is dominated by prokaryotes (93%) with

contributions that are <2% for each of the faunal compartments (Table 3). The contributions to total

respiration by the individual compartments are well-constrained, given the small standard deviations

(Table 3).

Largest carbon flows in the food web at the central HAUSGARTEN station is the deposition of

the three classes of detritus, which subsequently dissolve into DOC that is taken up by prokaryotes

and then respired by prokaryotes (Fig. 3A). All carbon flows in this pathway are >1 mmol C m-2 d-1.

Prokaryotic production is 1.84±0.12 mmol C m-2 d-1 and the prokaryotic growth efficiency is 0.38±0.03.

Much of the prokaryotic production (92±4%) undergoes cell lysis after viral infection (Danovaro et al.,

2008), and this carbon cycles back to DOC (Appendix). Other important flows are carbon burial

(0.31±0.08 mmol C m-2 d-1) and efflux of DOC from the sediment (0.19±0.10 mmol C m-2 d-1) (Fig. 3B).

Important faunal flows (>0.1 mmol C m-2 d-1) are uptake by surface-deposit feeding macrofauna and

suspension-feeding macro- and megafauna (Fig. 3B). Most carbon flows related to faunal

compartments, however, are between 0.005 and 0.05 mmol C m -2 d-1 (Fig. 3C). Finally, export flows

and carbon flows associated with the predatory+omnivore nematodes, predatory macrofauna and

megafauna are typically <3·10-3 mmol C m-2 d-1 (Fig. 3D).

Faunal secondary production is highest for macrofauna (0.10±0.004 mmol C m-2 d-1), followed

by megafauna (0.07±0.003 mmol C m-2 d-1) and nematodes (0.04±0.004 mmol C m-2 d-1) (Fig. 4). The

fate of the secondary production by the non-predatory faunal compartments shows that 83% of the

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

Page 12: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

deposit-feeding nematode production is grazed, but only to a small extent (6%) by predatory

nematodes, most production is predated upon by macro- (40%) and megafauna (38%) (Fig. 4B). The

maintenance costs are relatively higher for the macro- (22%) and megafauna (77%) compared to the

nematodes, because maintenance costs are a fixed fraction of the biomass per day, whereas

biomass-specific production rates decrease with faunal size (Table 2). For macrofauna, a total of 56%

is grazed by predatory macro- (18%) and megafauna (38%) (Fig. 4C). Finally, for non-predatory

megafauna, a similar proportion (6-10%) of the secondary production is grazed by predatory

megafauna, lost through mortality and exported from the food web (Fig. 4D).

The model results suggest that faunal diets are typically dominated by labile and semi-labile

detritus, with variable contributions among the compartments (Fig. 5). Despite the fact that deposit-

feeding nematodes form the principle carbon source of predatory nematodes (>80%, Fig. 5), this

represents only 6% of the fate of secondary production by deposit-feeding nematodes (Fig. 4A). The

surface-deposit feeding macrofauna and deposit-feeding macro- and megafauna derive carbon mainly

from three principle sources: labile detritus, semi-labile detritus and prokaryotes. Semi-labile detritus

dominates the diets of the deposit-feeding compartments (52±9% for MacDF and 52±12% for MegDF),

whereas prokaryotes (50±28%) are of similar importance for surface-deposit feeding macrofauna as

labile detritus (44±28%) with a lower contribution from semi-labile detritus (4±1%). Diets of suspension

feeding macro- and megafauna diets are dominated by semi-labile detritus with a lower contribution of

labile detritus (67±23% and 33±23% for MacSF, respectively and 59±17% and 41±17% for MegSF,

respectively) (Fig. 5). The diets of predatory macro- and megafauna are diverse and seem to be

similar among the two predatory compartments. Important contributions (>50%) are from the

macrofaunal compartments, most notably surface-deposit feeding macrofauna (>24%), nematodes

(>11%), with labile (<5%) and semi-labile (<13%) detritus representing a much lower contribution.

3.2 Trophic levels and dependencies on primary resources

The trophic level (TL) of suspension feeding macro- and megafauna is fixed at two (Fig. 6),

because the two suspended detritus sources are presumed to have a fixed trophic position of one (see

Material and Methods). The TL of deposit feeding nematodes is fairly well-determined and is slightly

higher than 2 because of a small contribution of prokaryotes in their diet. The TL of deposit-feeding

macro- and megafauna is similar and fairly well-determined with lower and upper quartiles of 2.2 and

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

328

329

330

331

332

333

Page 13: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

2.4 (Fig. 6), corresponding with the similarity in their diet compositions (Fig. 5). The TL of surface-

deposit feeding macrofauna, however, is much more uncertain and has lower and upper quartiles of

2.3 and 2.8 with a median of 2.5 (Fig. 6). This uncertainty is due to the uncertain diet contributions

described above of labile detritus (TL of 1) and prokaryotes (TL of 2). The TL of predatory+omnivore

nematodes is well constrained between 2.8 and 2.9, because of their predominant feeding on deposit-

feeding nematodes. The predatory macro- and megafauna have similar and highest TL with lower and

upper quartiles between 2.8 – 3.0, respectively, but with large excursions to lower and higher values

for their trophic level (Fig. 6). The high uncertainty is a result of the uncertainty in the diet composition

and the diets of its preys, but overall the higher TLs are expected for these predatory compartments.

The direct and indirect dependence on refractory detritus is lowest for all biotic compartments

in the food web (Fig. 7), with lower and upper quartiles between 0.15 and 0.77. Dependence of

prokaryotes is highest on semi-labile detritus (median of 1.5) and prokaryotes (median 2.5) (Fig. 7A).

Dependence on labile and semi-labile detritus is comparable for most biotic compartments with lower

quartiles between 0.98 and 1.15 and upper quartiles between 2.18 and 2.28 (Fig. 7B-H). The level of

uncertainty is high for the dependency values, particularly with respect to the upper levels of

dependency that can be >8, which is substantially higher than the median values (Fig. 7).

Overall it is clear that the standard deviations are fairly limited for respiration rates (Table 3)

and secondary production (see above) by the various compartments, which are measures of total

carbon processing. There is, however, a much higher variability in trophic levels (Fig. 6) and

dependencies (Fig. 7), which indicates that the uncertainty on the flows between the compartments is

substantially higher than the uncertainty on the total carbon processing by the compartments.

3.3 Sensitivity analysis

The perturbations of the stock values with ±15% of their default value in the sensitivity analysis

gave following results: for 46% of the flow values, the deviation in the perturbed solution was between

0 and 10% of the default flow value, for 37% this deviation was between 10 and 25% of the default

flow values, for 13% the deviation was between 25 and 50%, for 4% it was between 50 and 100% and

for <1% of the flow values a deviation of more than 100% of its default flow value was found. The

maximum deviation for a flow was 169%, which involved the export flow of surface-deposit feeding

macrofauna and occurred under a reduced stock value of the predatory megafauna. The reduced

334

335

336

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

352

353

354

355

356

357

358

359

360

361

362

Page 14: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

stock of predatory megafauna involved a reduction in its predation pressure on surface-deposit

feeding macrofauna due to which the export flow increased. Overall, however, the sensitivity analysis

revealed that in the model results were generally insensitive to perturbations in the stock values.

4. Discussion

Ecosystem dynamics in the Arctic Ocean are regulated by the strong seasonality in the light

and temperature regime and the cover of sea-ice (Grebmeier and Barry, 1991; Honjo et al., 2010;

Wassmann et al., 2006). The Arctic Ocean is surrounded by landmasses with extensive shallow

continental shelves with strong benthic-pelagic coupling (Grebmeier and Barry, 1991), particularly in

the ice-edge and ice-free regions such as the Bering Sea (Grebmeier et al., 2006), Barents Sea (De

Laender et al., 2010; Wassmann et al., 2006) and Chukchi Sea (Moran et al., 2005). This benthic-

pelagic coupling on the shallow shelves results in a comparatively high fraction (>15%) of the primary

production being processed by the benthos, sustaining high levels of macrofaunal biomass

(Grebmeier et al., 1988; Renaud et al., 2007). Compared to the shallow continental shelves, the deep

sediments of the Arctic Ocean are much less studied (e.g. Bergmann et al., 2009; Clough et al., 1997;

Iken et al., 2005; Kröncke, 1994; Vanreusel et al., 2000; Wlodarska-Kowalczuk and Pearson, 2004),

especially at the integration level of the whole food web.

Fram Strait is located between Spitsbergen and Greenland and forms a deep (>2000 m) and

narrow connection between the Arctic Ocean and the Atlantic Ocean (Fig. 1). In this region, the

amount of organic matter processed in the sediment decreases as allometric functions of water depth

and primary production (Schluter et al., 2000). To detect and track the impact of large-scale

environmental changes in the transition zone between the northern North Atlantic and the central

Arctic Ocean, the long-term observatory HAUSGARTEN was established (Budaeva et al., 2008; Hoste

et al., 2007; Soltwedel et al., 2009). Export fluxes from the euphotic zone at the HAUSGARTEN

observatory are restricted to <10% of the primary production suggesting an efficient processing by the

pelagic food web (Bauerfeind et al., 2009). Although information on biomass of various compartments

(Budaeva et al., 2008; Hoste et al., 2007; Soltwedel et al., 2009) and organic carbon deposition

(Bauerfeind et al., 2009) is well-documented, no inferences have been made on how the organic

carbon that arrives at the seafloor is processed within the benthic community. To address this

deficiency, the available data were merged to a food web model using linear inverse methodology

363

364

365

366

367

368

369

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

Page 15: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

(Soetaert and Van Oevelen, 2009). Since the resulting food web structure depends heavily on the

model assumptions and data quality, it is essential to start with a critical appraisal of these.

4.1 Model assumptions

The inherent heterogeneity of sedimentary detritus implies that various detritus fractions have

degradation rates differing over orders of magnitude (Middelburg, 1989; Moore et al., 2004; Westrich

and Berner, 1984). On the one hand, it is impossible to do justice to this continuum of degradation

rates within a food web context, firstly because too many detritus compartments would need to be

defined and secondly because no data are available to constrain their dynamics. On the other hand,

mass-balance models of sediment food webs typically merge all dead organic matter into one

homogeneous “detritus” compartment (e.g. Rowe et al., 2008), which may be a too crude

simplification. Here, three detritus compartments were defined based on empirical data using a similar

approach as Van Oevelen et al. (2011). Labile detritus was defined as all carbon associated with

chlorophyll a. Chlorophyll a deposition is typically linked to the input of fresh phytodetritus, both in

coastal (Sun et al., 1991), canyon (Van Oevelen et al., 2011) and abyssal plain (Stephens et al., 1997;

Witbaard et al., 2000) sediments. Henceforth, the particulate organic carbon that is associated with

chlorophyll a, using a carbon : chlorophyll a ratio for living phytoplankton (following Stephens et al.,

1997), represents a natural choice to define the labile detritus compartment. In addition to a labile

detritus fraction, it is readily established that there is a refractory detritus pool in marine sediments that

is only degradable by prokaryotes (Benner et al., 1986; Deming and Baross, 1993; Pfannkuche, 2005).

Therefore, total particulate organic carbon in the sediment minus the labile and semi-labile detritus

was defined refractory and hence not degradable by benthic fauna. The most ambiguous detritus pool

to define was semi-labile detritus, which was here defined as the sum of extractable particulate

proteins in the top 5 cm. Amino acids are frequently used as indicator for the lability of detritus (Dauwe

et al., 1999; Kiriakoulakis et al., 2001; Mayer et al., 1995) and termed are semi-labile (Fabiano et al.,

2001), because they do not degrade in short time scales and proteins have intermediate degradation

rates in experimental decays studies (Harvey et al., 1995). Since the extraction and hydrolyzation

methods used for the characterization of proteins are harsher (e.g. low pH, high temperature) than

those present in digestive tracts (Mayer et al., 1995), it is likely that this definition represents an upper

limit on the semi-labile detritus stock. Van Oevelen et al. (2011) followed a similar approach, although

392

393

394

395

396

397

398

399

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

Page 16: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

they defined semi-labile detritus as the sum of lipids, carbohydrates and proteins. However, proteins

comprise ~50% of this total pool. To account for the uncertainty in classifying the detritus classes, the

lower and upper bounds on the degradation rates of semi-labile detritus encompasses two orders of

magnitude (Table 2). Also, our sensitivity analysis showed that the model results presented are

insensitive to the stock values ± 15% (see Results). Admittedly, our separation into various detritus

classes is an operational definition. However, it does better justice to the natural detritus heterogeneity

than considering simply one pool and it is linked to measurable quantities. Moreover, there is roughly

an order of magnitude increase in the different detritus stocks with decreasing lability (Table 2),

suggesting a reasonable coverage using these metrics.

Some biotic compartments are missing from the food web topology (Fig. 3). Microfauna or

nanobenthos (i.e. flagellates and ciliates) are not included because of a lack of biomass data. This is a

problem in most deep-sea studies such that the role of the nanofauna in carbon cycling of deep-sea

food webs remains an open question. However, their limited biomass compared to, for example

Foraminifera (Alongi, 1992), may suggest a limited role in carbon processing. Meiofauna were

represented by two nematode compartments (Fig. 3). Hoste et al. (2007) showed that nematodes

strongly dominated the metazoan meiofauna (85-99%) at HAUSGARTEN, such that the omission of

other metazoan meiofauna is probably not significant, at least when it comes down to carbon

processing rates. Foraminifera, i.e. protozoan meiofauna, had to be omitted from the Hausgarten food

web model because no biomass data were available. Therefore their role in carbon processing could

not be assessed and this represents a shortcoming of the present model. Foraminifera can have

comparable or higher biomass levels compared to the metazoan meiofauna in deep sediments

(Bernhard et al., 2008; Gooday, 1986; Witte et al., 2003) and have been shown to be important

contributors to short-term processing of fresh phytodetritus in deep sediments (Moodley et al., 2002).

A specific study on the Foraminifera showed that their contribution to total respiration was limited to

0.5 – 2.5% only (Geslin et al., 2010) and the uptake of 13C-phytodetritus by the larger (>300 μm)

Foraminifera is generally limited (Woulds et al., 2007). Moreover, if biomass of Foraminifera were of

comparable magnitude as the nematodes at the Hausgarten, then their role will be limited considering

the limited role of nematodes in carbon processing (Fig. 3 and 4 and Table 3). The absence of

nanobenthos and Foraminifera as specific compartments in essence implies that their role in carbon

processing is included in the prokaryotes. This latter compartment acts as a closure term on

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

448

449

450

Page 17: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

respiration, because specific information of the prokaryotic respiration or production was unavailable.

Overall, however, very detailed benthic biomass data were available ranging from prokaryotes to

megafauna, that could be split among feeding types using taxonomic information and stable isotope

studies (Bergmann et al., 2009). Carbon flows could therefore be inferred at a high resolution,

especially considering the fact that we are dealing with a deep-sea food web.

The resulting standard deviations on the carbon flows are limited for prokaryotic and faunal

respiration rates and secondary production (see above), which are all measures of total carbon

processing. These carbon flows are predominantly constrained by biomass data and literature

constraints (Table 2). There is, however, higher variability in diet compositions (see ‘Results’), trophic

levels (Fig. 6) and food dependencies (Fig. 7), indicating that the uncertainty on the flows between the

compartments is substantially higher than the uncertainty on the total carbon processing by the

compartments (see also Appendix). This higher uncertainty on flows between compartments is

undoubtedly the result of limited data that are available to constrain these flows, a situation that is

typical for benthic food web reconstructions. This may be improved by more detailed information on

diet composition using for example fatty acid composition data (Iverson et al., 2004; Meziane et al.,

1997) or trophic level indicatios using δ15N values of specific amino acids (Chikaraishi et al., 2009).

4.2 Carbon budget

Respiration by the total community was estimated at 3.26±0.20 mmol C m-2 d-1. This respiration

rate is 3 to 10 times lower than the range of 10.3 – 35.6 mmol C m-2 d-1 that can be inferred from an

empirical relation based on in-situ sediment oxygen consumption rates from open slope sediments

(Andersson et al., 2004), although substantially higher than the 0.12 – 0.25 mmol C m-2 d-1 measured

at similar depth at an Arctic continental slope of the Laptev Sea (Boetius and Damm, 1998).

Bauerfeind et al. (2009) report low export from the pelagic food web (<10% of primary production) at

the central HAUSGARTEN station and explain this by effective recycling within the pelagic community.

However, short-term sedimentation events like ice-edge blooms or detached sympagic algae may

cause a temporary decoupling from the pelagic food web. The food web model resolves carbon fluxes

in spring/summer 2003, during which carbon export was comparatively large compared to other years

and was composed mostly of diatomaceous material (Bauerfeind et al., 2009). This decoupling in the

451

452

453

454

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

473

474

475

476

477

478

Page 18: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

pelagic food web renders total respiration rates of the sediment comparatively low compared to other

open slopes.

Respiration was clearly dominated by prokaryotes (93±0.6%) with contributions of less than 2% by

different faunal compartments (Table 3). Piepenburg et al. (1995) conducted an extensive study in the

north-eastern Barents Sea, in which sediment community oxygen consumption (SCOC) rates,

including micro-, meio and macrofaunal respiration, were amended with respiration rates of

megafauna and fish. These authors also report a microbial dominance (57%) for slope sediments (200

– 550 m), with more limited contributions of meio- (7%), macro- (21%) and megafauna (16%).

Ambrose et al. (2001) found contributions of up to 25% at shallow (<50 m) stations, with respiration

rates of >1 mmol C m-2 d-1 for epibenthic echinoderms. These faunal contributions and rates are

consistently higher than estimated here for the deeper HAUSGARTEN station (2500 m). This pattern

is also seen in other regions were faunal contributions can be as high as 50% at the shelf and shelf

break sediments compared to continental slope and abyssal plain environments (Heip et al., 2001;

Woulds et al., 2009). The shift towards an increased microbial contribution to carbon processing

possibly relates to energy limitation at greater depths, such that population densities of large

organisms simply become too low to remain reproductively viable (Rex et al., 2006). In all, the

respiration partitioning at the central HAUSGARTEN station is more comparable to abyssal plain food

webs that are under strong energy limitation, compared to shallower sediments where the faunal

contribution is typically larger.

4.3 Faunal carbon flows and position in food web

The nematode community contributed surprisingly little to total respiration (<1%, Table 3),

especially when compared to a global estimate of around 7.5% for nematodes (Soetaert et al., 2009).

This limited contribution of nematodes is also found in two isotope pulse-chase experiments

conducted in Arctic sediments. Ingels et al. (2010) inferred that <1% of the added organic matter

sources (bacteria and diatoms) at the central HAUSGARTEN station were processed by nematodes.

Urban-Malinga and Moens (2006) conducted 13C-phytodetritus tracer experiments in two Arctic beach

sediments and reported meiofaunal processing of <5%. The diet composition of deposit-feeding

nematodes indicates that a substantial 24% of their carbon requirements is derived from labile detritus

(Fig. 5). This seems to contradict the findings of Ingels et al. (2010), who report a higher uptake of 13C-

479

480

481

482

483

484

485

486

487

488

489

490

491

492

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

Page 19: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

labelled bacteria compared to13C-labelled diatoms. However, the total uptake of labelled bacteria was

limited to <6·10-5 mmol C m-2 d-1, which is less than 0.1% of the carbon assimilation of 0.06±0.007

mmol C m-2 d-1 inferred here. This indicates that although uptake rates of labelled bacteria were higher

than labelled diatoms, these labelled carbon sources were insignificant in the total carbon

requirements of the nematodes. Whether this is related to the reduced food availability because of

reduced mixing of the carbon sources into the sediment or methodological bias because of freeze-

drying of the carbon sources, as Ingels et al. (2010) note, is presently unclear. The limited uptake of

bacterial carbon, however, is in agreement with experimental results from Guilini et al. (2010), who

injected different 13C enriched dissolved organic compounds into the top 5 cm of the sediment of a

shallower HAUSGARTEN station and traced the 13C into prokaryotes and subsequently nematodes

(henceforth the constraints in Table 2). We lack sufficient evidence to more precisely quantify labile

and semi-labile diet contributions for the nematode community (Fig. 5 and Fig. 7B-C), but modelling

results indicate a dominance of semi-labile detritus. Hoste et al. (2007) found inter-annual changes in

nematode biomass in the years 2001 – 2004 at the central HAUSGARTEN site, but these could not be

related directly to export fluxes from the water column. The authors indicated that the absence of this

relation may have been related to uncertainties in exact timing of export flux and the amount actually

arriving at the seafloor. Present results, however, indicate that this may also be due to a dependence

on a semi-labile pool that are much less dynamic given the degradation rate of 0.0018 d-1

corresponding to a half-life of 385 days. Ruhl et al. (2008) report response times of the benthic

community to a POC deposition event of 4 – 6 months for abundance and up to 10 months for

biomass. It is not clear, however, to what extent the benthos were directly acquiring their carbon from

the freshly deposited carbon compared to the more stable semi-labile detritus pools. Their figures also

show that benthic biomass does not decrease monotonically with POC fluxes approaching zero,

indicating at least partial reliance on a less fraction detritus pool.

Only a limited fraction of secondary production (~6%) of the deposit-feeding nematodes is

transferred to predatory nematodes, despite the fact that deposit-feeding nematodes represent 80% of

the diet of predatory nematodes. This indicates that predation may not control their biomass at this

station. Gallucci et al. (2008) used cage experiments at HAUSGARTEN to investigate the impact of

presence/absence of megafauna on the community structure of nematodes. Total nematode densities

were higher within the cages, probably related to higher food abundance, although the percentage of

508

509

510

511

512

513

514

515

516

517

518

519

520

521

522

523

524

525

526

527

528

529

530

531

532

533

534

535

536

537

Page 20: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

predatory nematodes was low (~3%) and similar within and outside the cages. Similarly low

abundances of predatory nematodes were reported in sediments of the Laptev Sea (Vanaverbeke et

al., 1997), suggesting that the ‘incomplete’ utilisation of secondary production within the nematode

sub-food web may be a more general phenomenon in Arctic sediments.

The non-predatory macrofaunal compartments account for half of the faunal secondary

production, which is efficiently transferred up the food web to the macro- (18%) and megabenthic

(38%) predators. The transfers also result in high contributions of non-predatory macrofauna in diets of

predatory macro- and megafauna. Only few studies are available for comparison. Rowe et al. (2008)

determined predation rates by megafauna in the Gulf of Mexico, but a priori assumed that organisms

preferentially feed on larger prey items. Predation rates ranged from 2·10-4 – 0.03 mmol C m-2 d-1,

depending on station depth. Total predation by the megafauna at the HAUSGARTEN station amounts

to 0.055±0.007 mmol C m-2 d-1 and is comparable to the upper slope stations (500 – 1000 m) of the

Gulf of Mexico. Total megafaunal biomass at HAUSGARTEN (6 mmol C m-2) is somewhat higher

compared to these shallower stations (2.2 – 2.6 mmol C m-2) despite the lower organic matter input at

HAUSGARTEN (3.78 vs. 4.2 – 9.5 mmol C m-2 d-1). This implies that the Arctic HAUSGARTEN

megafauna appear to take more advantage of the organic matter flux than megafaunal assemblages

from the tropical Gulf of Mexico.

The diet contribution of surface-deposit feeding macrofauna seems to be dominated by

prokaryotes (50±28%), although there is substantial uncertainty in these diet reconstructions as

evidenced by the large standard deviation (see also ‘Results’). There is no experimental evidence on

the importance of prokaryotes in the diets of deposit-feeding macro- and megafauna, but the model

results suggest 15 to 20%. Constraints on deep-sea feeding strategies based on optimal foraging

theory imply that deposit feeders may use a strategy in which oxygen and ammonium is supplied to

prokaryotes to facilitate (pre-)degradation of detritus and the prokaryotes are subsequently grazed

(Jumars et al., 1990). Although this theory has not been rigorously tested for benthic food webs, our

results here show that prokaryotes may indeed contribute to carbon demands, but further experimental

work is required to substantiate these findings.

The trophic levels (TL) that are inferred from the model are not directly comparable with those

estimated at species-level with δ15N measurements (e.g. Bergmann et al., 2009; Iken et al., 2005; Iken

et al., 2001). Here, the trophic level is calculated for benthic compartments from a large set of model

538

539

540

541

542

543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

566

567

Page 21: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

solutions that are feasible within the current data set. As such, these results can be interpreted as the

range of TLs that are feasible within the different biotic compartments, with two restrictions. (1) The

ranges in the model results are based on a compartment in which species are lumped into functional

groups and will therefore be more limited than those based on species-specific δ15N, in which more

extreme values can be found. (2) The δ15N of detritus may increase during progressive degradation

resulting in a δ15N difference between labile and semi-labile detritus (Altabet, 1996). This fractionation

effect is not included in the TL calculation, because it does not influences an organism’s TL. Moreover,

the estimates of TL based on δ15N may not be accurate if there is a large difference in δ15N of the

primary food sources.

Overall, the model results are consistent with the δ15N results that Bergmann et al. (2009) obtained

at the HAUSGARTEN stations, in that deposit feeders mostly occupy the second and most predators

the third TL. In addition, the model results show that the largest range of trophic levels is found for the

predatory compartments (TL range of 2.5 – 3.3, respectively), which is qualitatively consistent with the

results from Bergmann et al. (2009) but their δ15N data indicate a larger range in TLs of up to 3.5. The

largest discrepancy is found for suspension feeders that are at TL of 2 in the model, because in the

model setup they are assumed to feed exclusively on suspended detritus (with fixed TL of 1).

Bergmann et al. (2009), however, find a much larger variation in the suspension feeders (range of 3

TLs), which they attribute to starvation effects, feeding selectivity and the uncertain classification of

sponges and anthozoans (that are potentially carnivorous or rely on microbial farming).

The spread in TLs also agrees with a study of δ15N signatures of benthic fauna in the Canada

Basin (Iken et al., 2005), where benthic fauna are found mainly at a TL of 2 – 3. Moreover, they infer

that deposit feeders rely to a large extent on less labile detritus because of the large difference in δ15N

between benthic deposit feeders and fresh detritus (i.e. sympagic algae and pelagic algae). Our model

results also indicate that semi-labile detritus is an important (>50%) component of the diets of deposit-

feeding macro- and megafauna.

Trophic dependency quantifies the dependence of a consumer on a resource through direct (i.e.

grazing) and indirect (i.e. longer loops in the food webs) pathways in the food web, thereby giving a

more complete view on trophic interactions than when looking at direct interactions only (Ulanowicz,

2004). Overall, dependence on refractory detritus is low for all biotic compartments. The dependencies

that are inferred for the other basal resources show an interesting feature: although labile detritus is

568

569

570

571

572

573

574

575

576

577

578

579

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

596

597

Page 22: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

generally substantially less important in the diet compositions than semi-labile detritus, the

dependence on labile detritus is comparable or even slightly higher than dependence on semi-labile

detritus. This is due to the effect of combining all pathways, and not only the direct interactions, into

the dependency indices. As such, this may indicate that the benthic food web is more sensitive to

changes in labile detritus input as may be inferred from their diet compositions alone.

4.4 Speculations on future conditions

Based on the results of this food web reconstruction, we conclude that carbon mineralisation at

the central HAUSGARTEN station (2500 m) is strongly dominated by prokaryotes with limited

contributions of the faunal compartments. The limited vertical export of particulate organic matter

imposes energy limitation on the benthic food web such that carbon processing resembles that of an

abyssal plain food web. The Arctic Ocean is a region where major shifts in the ecosystem are

expected due to climate change. Grebmeier et al. (2006) describe a shift in the Bering Sea, where

high production of sympagic algae favoured a high export to the benthos and resulted in a high benthic

production. This situation changed with a receding ice-edge to a pelagic dominated food web with

limited export fluxes and decreasing benthic production.

How the benthic food web at HAUSGARTEN will change under projected climate change will

mainly depend on the changes in the pelagic food web structure and export of organic matter. It could

be argued that ice-free conditions promote phytoplankton growth, as projected for some Arctic regions

due to temperature and light penetration increases as a result of shrinking sea ice (Arrigo et al., 2008;

Slagstad et al., 2010). However, primary production may rise only slightly if increased thermal or

haline stratification limits mixing and upward nutrient transport (Carmack et al., 2006; Slagstad et al.,

2010). In addition, mesozooplankton abundance may increase in the Fram Strait, since Atlantic

species extend their range as more Atlantic water masses prevail and sea surface temperatures rise

(Hirche and Kosobokova, 2007). This would amplify the grazing pressure and lead to increased

retention in the water column (Carroll and Carroll, 2003). The retreat of the ice edge and the

continuous loss of multi-year ice will lead to a lower flux of fast-sinking sympagic algae and ice-related

POM (Forest et al., 2010; Hop et al., 2006), which may affect megafaunal deposit feeders such as

holothurians (Bergmann et al., 2011). In sum, all this would lead to a decreased carbon deposition at

the deep seafloor, which is already characterised by food limitation.

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

619

620

621

622

623

624

625

626

Page 23: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

Based on the results from our model, we do not expect that shifts in the overall functioning of the

benthic food web will occur rapidly, because semi-labile detritus plays an significant role in the benthic

food web. The semi-labile detritus is a stock with a comparatively low degradation rate (0.0018±0.0007

d-1 and corresponding half-life of 382 days) such that changes will be slower to observe than expected

from sole differences in detritus deposition rates. The dependency indices of the benthic fauna on

labile and semi-labile detritus were, however, of comparable magnitude such that that the benthic food

web may be more sensitive to changes in labile detritus input as may be inferred from their diet

compositions alone. Unfortunately, there were not enough data to allow better discrimination between

pelagic and sympagic plankton inputs; if some species specifically select sympagic phytodetritus, as

seen for some pelagic fauna (Hop et al., 2006), these species will be especially vulnerable. It will be

important to study the species-specific feeding preferences in detail to assess the vulnerability of

individual components of the benthic food web.

5. Acknowledgements

We thank the officers and crews of RVs Polarstern and Maria S. Merian and the team of the

remotely operated vehicle ‘‘Victor 6000’’ for their support. We also acknowledge the work of our

technicians and student workers in the laboratory and at sea. Three anonymous reviewers and Andy

Gooday are thanked for constructive comments that considerably improved an earlier version of this

manuscript. This research was supported by the HERMES project (contract GOCE-CT-2005-511234),

funded by the European Commission’s Sixth Framework Programme under the priority “Sustainable

Development, Global Change and Ecosystems”, and HERMIONE project (grant agreement n°

226354") funded by the European Community's Seventh Framework Programme (FP7/2007-2013).

This is publication **** from the Netherlands Institute of Ecology (NIOO-KNAW), Yerseke and

publication awi-n**** of the Alfred Wegener Institute for Polar and Marine Research.

627

628

629

630

631

632

633

634

635

636

637

638

639

640

641

642

643

644

645

646

647

648

649

Page 24: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

6. References

Allen, J.T., Brown, L., Sanders, R., Moore, C.M., Mustard, A., Fielding, S., Lucas, M., Rixen, M.,

Savidge, G., Henson, S., Mayor, D., 2005. Diatom carbon export enhanced by silicate upwelling in

the northeast Atlantic. Nature 437 (7059), 728-732.

Alongi, D.M., 1992. Bathymetric patterns of deep-sea benthic communities from bathyal to abyssal

depths in the western south Pacific (Solomon and Coral Seas). Deep-Sea Research Part a-

Oceanographic Research Papers 39 (3-4A), 549-565.

Altabet, M.A., 1996. Nitrogen and carbon isotopic tracers of the source and transformation of particles

in the deep sea. In: Ittekkot, V., Schäffer, P., Honjo, S., Depetris, P.J. (Eds.), Particle flux in the

ocean. Wiley, Chichester, pp. 155-184.

Ambrose, W.G., Clough, L.M., Tilney, P.R., Beer, L., 2001. Role of echinoderms in benthic

remineralization in the Chukchi Sea. Marine Biology 139 (5), 937-949.

Andersson, J.H., Wijsman, J.W.M., Herman, P.M.J., Middelburg, J.J., Soetaert, K., Heip, C., 2004.

Respiration patterns in the deep ocean. Geophysical Research Letters 31 (3), L03304, doi

03310.01029/02003gl018756.

Arrigo, K.R., van Dijken, G., Pabi, S., 2008. Impact of a shrinking Arctic ice cover on marine primary

production. Geophys. Res. Lett. 35 (19), L19603.

Bauerfeind, E., Nothig, E.M., Beszczynska, A., Fahl, K., Kaleschke, L., Kreker, K., Klages, M.,

Soltwedel, T., Lorenzen, C., Wegner, J., 2009. Particle sedimentation patterns in the eastern Fram

Strait during 2000-2005: Results from the Arctic long-term observatory HAUSGARTEN. Deep-Sea

Research Part I-Oceanographic Research Papers 56 (9), 1471-1487.

Benner, R., Moran, M.A., Hodson, R.E., 1986. Biogeochemical cycling of lignocellulosic carbon in

marine and freshwater ecosystems: Relative contributions of procaryotes and eucaryotes.

Limnology and Oceanography 31 (1), 89-100.

Bergmann, M., Dannheim, J., Bauerfeind, E., Klages, M., 2009. Trophic relationships along a

bathymetric gradient at the deep-sea observatory HAUSGARTEN. Deep-Sea Research Part I-

Oceanographic Research Papers 56 (3), 408-424.

Bergmann, M., Soltwedel, T., Klages, M., 2011. The interannual variability of megafaunal

assemblages in the Arctic deep sea: Preliminary results from the HAUSGARTEN observatory

650

651

652

653

654

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

673

674

675

676

677

678

Page 25: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

(79°N). Deep Sea Research Part I: Oceanographic Research Papers In Press, Accepted

Manuscript.

Bernhard, J.M., Sen Gupta, B.K., Baguley, J.G., 2008. Benthic foraminifera living in Gulf of Mexico

bathyal and abyssal sediments: Community analysis and comparison to metazoan meiofaunal

biomass and density. Deep Sea Research Part II: Topical Studies in Oceanography 55 (24-26),

2617-2626.

Billett, D.S.M., Bett, B.J., Reid, W.D.K., Boorman, B., Priede, I.G., 2010. Long-term change in the

abyssal NE Atlantic: The 'Amperima Event' revisited. Deep-Sea Research Part II-Topical Studies in

Oceanography 57 (15), 1406-1417.

Boetius, A., Damm, E., 1998. Benthic oxygen uptake, hydrolytic potentials and microbial biomass at

the Arctic continental slope. Deep-Sea Research Part I-Oceanographic Research Papers 45 (2-3),

239-275.

Borsheim, K.Y., Bratbak, G., Heldal, M., 1990. Enumeration and biomass estimation of planktonic

bacteria and viruses by transmission electron-microscopy. Applied and Environmental Microbiology

56 (2), 352-356.

Budaeva, N.E., Mokievsky, V.O., Soltwedel, T., Gebruk, A.V., 2008. Horizontal distribution patterns in

Arctic deep-sea macrobenthic communities. Deep-Sea Research Part I-Oceanographic Research

Papers 55 (9), 1167-1178.

Buesseler, K.O., Lamborg, C.H., Boyd, P.W., Lam, P.J., Trull, T.W., Bidigare, R.R., Bishop, J.K.B.,

Casciotti, K.L., Dehairs, F., Elskens, M., Honda, M., Karl, D.M., Siegel, D.A., Silver, M.W.,

Steinberg, D.K., Valdes, J., Van Mooy, B., Wilson, S., 2007. Revisiting carbon flux through the

ocean's twilight zone. Science 316 (5824), 567-570.

Carmack, E., Barber, D., Christensen, J., MacDonald, R., Rudels, B., Sakshaug, E., 2006. Climate

variability and physical forcing of the food webs and the carbon budget on panarctic shelves.

Progress In Oceanography 71 (2-4), 145-181.

Carroll, M., Carroll, J., 2003. The Arctic Seas. In: Black, K., Shimmield, G. (Eds.), Biogeochemistry of

marine systems. Blackwell, Oxford, pp. 127-147.

Chikaraishi, Y., Ogawa, N.O., Kashiyama, Y., Takano, Y., Suga, H., Tomitani, A., Miyashita, H.,

Kitazato, H., Ohkouchi, N., 2009. Determination of aquatic food-web structure based on compound-

679

680

681

682

683

684

685

686

687

688

689

690

691

692

693

694

695

696

697

698

699

700

701

702

703

704

705

706

707

Page 26: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

specific nitrogen isotopic composition of amino acids. Limnology and Oceanography-Methods 7,

740-750.

Clough, L.M., Ambrose, W.G., Cochran, J.K., Barnes, C., Renaud, P.E., Aller, R.C., 1997. Infaunal

density, biomass and bioturbation in the sediments of the Arctic Ocean. Deep-Sea Research Part

II-Topical Studies in Oceanography 44 (8), 1683-1704.

Danovaro, R., Dell'Anno, A., Corinaldesi, C., Magagnini, M., Noble, R., Tamburini, C., Weinbauer, M.,

2008. Major viral impact on the functioning of benthic deep-sea ecosystems. Nature 454 (7208),

1084-U1027.

Dauwe, B., Middelburg, J.J., Herman, P.M.J., Heip, C.H.R., 1999. Linking diagenetic alteration of

amino acids and bulk organic matter reactivity. Limnology and Oceanography 44 (7), 1809-1814.

De Laender, F., Van Oevelen, D., Soetaert, K., Middelburg, J.J., 2010. Carbon transfer in herbivore-

and microbial loop-dominated pelagic food webs in the southern Barents Sea during spring and

summer. Marine Ecology-Progress Series 398, 93-107.

Deming, J.W., Baross, J.A., 1993. The early diagenesis of organic matter: Bacterial activity. In: Engel,

M.H., Macko, S.A. (Eds.), Organic Geochemistry. Plenum Press, New York.

Fabiano, M., Pusceddu, A., Dell'Anno, A., Armeni, M., Vanucci, S., Lampitt, R.S., Wolff, G.A.,

Danovaro, R., 2001. Fluxes of phytopigments and labile organic matter to the deep ocean in the

NE Atlantic Ocean. Progress in Oceanography 50 (1-4), 89-104.

Forest, A., Wassmann, P., Slagstad, D., Bauerfeind, E., Nöthig, E.-M., Klages, M., 2010. Relationships

between primary production and vertical particle export at the Atlantic-Arctic boundary (Fram Strait,

HAUSGARTEN). Polar Biology http://dx.doi.org/10.1007/s00300-010-0855-3.

Gallucci, F., Fonseca, G., Soltwedel, T., 2008. Effects of megafauna exclusion on nematode

assemblages at a deep-sea site. Deep-Sea Research Part I-Oceanographic Research Papers 55

(3), 332-349.

Geslin, E., Risgaard-Petersen, N., Lombard, F., Metzger, E., Langlet, D., Jorissen, F., 2010. Oxygen

respiration rates of benthic foraminifera as measured with oxygen microsensors. Journal of

Experimental Marine Biology and Ecology doi:10.1016/j.jembe.2010.10.011.

Ginger, M.L., Billett, D.S.M., Mackenzie, K.L., Kiriakoulakis, K., Neto, R.R., Boardman, D.K., Santos,

V., Horsfall, I.M., Wolff, G.A., 2001. Organic matter assimilation and selective feeding by

708

709

710

711

712

713

714

715

716

717

718

719

720

721

722

723

724

725

726

727

728

729

730

731

732

733

734

735

736

Page 27: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

holothurians in the deep sea: some observations and comments. Progress in Oceanography 50 (1-

4), 407-421.

Glover, A.G., Gooday, A.J., Baily, D.M., Billett, D.S.M., Chevaldonné, P., Colaco, A., Copley, J.,

Cuvelier, D., Desbruyères, D., Kalogeropoulou, V., Klages, M., Lampadariou, N., Lejeusne, C.,

Mestre, N.C., Paterson, G.L.J., Perez, T., Ruhl, H., Sarrazin, J., Soltwedel, T., Soto, E.H., Thatje,

S., Tselepides, A., Van Gaever, S., Vanreusel, A., 2010. Temporal Change in Deep-Sea Benthic

Ecosystems: A Review of the Evidence From Recent Time-Series Studies. Advances in Marine

Biology 85 (1-95).

Gooday, A.J., 1986. Meiofaunal foraminiferans from the bathyal Porcupine Seabight (northeast

Atlantic) - size structure, standing stock, taxonomic composition, species-diversity and vertical-

distribution in the sediment. Deep-Sea Research 33 (10), 1345-1373.

Grebmeier, J.M., Barry, J.P., 1991. The influence of oceanographic processes on pelagic-benthic

coupling in polar regions: A benthic perspective. Journal of Marine Systems 2, 495-518.

Grebmeier, J.M., McRoy, C.P., Feder, H.M., 1988. Pelagic-benthic coupling on the shelf of the

northern Bering and Chukchi Seas. 1. Food-supply source and benthic biomass. Marine Ecology-

Progress Series 48 (1), 57-67.

Grebmeier, J.M., Overland, J.E., Moore, S.E., Farley, E.V., Carmack, E.C., Cooper, L.W., Frey, K.E.,

Helle, J.H., McLaughlin, F.A., McNutt, S.L., 2006. A major ecosystem shift in the northern Bering

Sea. Science 311 (5766), 1461-1464.

Greiser, N., Faubel, A., 1988. Biotic factors. In: Higgens, R.P., Thiel, H. (Eds.), Introduction to the

study of meiofauna. Smithsonian Institution Press, Washington D.C., London, pp. 79-114.

Grossmann, S., Reichardt, W., 1991. Impact of Arenicola marina on bacteria in intertidal sediments.

Marine Ecology Progress Series 77 (1), 85-93.

Guilini, K., Van Oevelen, D., Soetaert, K., Middelburg, J.J., Vanreusel, A., 2010. Nutritional importance

of benthic bacteria for deep-sea nematodes from the Arctic ice margin: Results of an isotope tracer

experiment. Limnology and Oceanography 55 (5), 1977-1989.

Gumprecht, R., Gerlach, H., Nehrkorn, A., 1995. FDA hydrolysis and resazurin reduction as a

measure of microbial activity in sediments from the south-east Atlantic. Helgolander

Meeresuntersuchungen 49 (1-4), 189-199.

737

738

739

740

741

742

743

744

745

746

747

748

749

750

751

752

753

754

755

756

757

758

759

760

761

762

763

764

765

Page 28: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

Harvey, H.R., Tuttle, J.H., Bell, J.T., 1995. Kinetics of Phytoplankton Decay During Simulated

Sedimentation - Changes in Biochemical-Composition and Microbial Activity under Oxic and Anoxic

Conditions. Geochimica Et Cosmochimica Acta 59 (16), 3367-3377.

Heip, C.H.R., Duineveld, G., Flach, E., Graf, G., Helder, W., Herman, P.M.J., Lavaleye, M.,

Middelburg, J.J., Pfannkuche, O., Soetaert, K., Soltwedel, T., de Stigter, H., Thomsen, L.,

Vanaverbeke, J., de Wilde, P., 2001. The role of the benthic biota in sedimentary metabolism and

sediment-water exchange processes in the Goban Spur area (NE Atlantic). Deep-Sea Research

Part II-Topical Studies in Oceanography 48 (14-15), 3223-3243.

Hirche, H.-J., Kosobokova, K., 2007. Distribution of Calanus finmarchicus in the northern North

Atlantic and Arctic Ocean--Expatriation and potential colonization. Deep-Sea Research II 54 (23-

26), 2729-2747.

Honjo, S., Krishfield, R.A., Eglinton, T.I., Manganini, S.J., Kemp, J.N., Doherty, K., Hwang, J., McKee,

T.K., Takizawa, T., 2010. Biological pump processes in the cryopelagic and hemipelagic Arctic

Ocean: Canada Basin and Chukchi Rise. Progress in Oceanography 85 (3-4), 137-170.

Hop, H., Falk-Petersen, S., Svendsen, H., Kwasniewski, S., Pavlov, V., Pavlova, O., Soreide, J.E.,

2006. Physical and biological characteristics of the pelagic system across Fram Strait to

Kongsfjorden. Progress in Oceanography 71 (2-4), 182-231.

Hoste, E., Vanhove, S., Schewe, I., Soltwedel, T., Vanreusel, A., 2007. Spatial and temporal variations

in deep-sea meiofauna assemblages in the Marginal Ice Zone of the Arctic Ocean. Deep-Sea

Research Part I-Oceanographic Research Papers 54 (1), 109-129.

Iken, K., Bluhm, B.A., Gradinger, R., 2005. Food web structure in the high Arctic Canada Basin:

evidence from delta C-13 and delta N-15 analysis. Polar Biology 28 (3), 238-249.

Iken, K., Brey, T., Wand, U., Voigt, J., Junghans, P., 2001. Food web structure of the benthic

community at the Porcupine Abyssal Plain (NE Atlantic): a stable isotope analysis. Progress in

Oceanography 50 (1-4), 383-405.

Ingels, J., Van den Driessche, P., De Mesel, I., Vanhove, S., Moens, T., Vanreusel, A., 2010.

Preferred use of bacteria over phytoplankton by deep-sea nematodes in polar regions. Marine

Ecology-Progress Series 406, 121-133.

IPCC, 2007. Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to

the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. In: Solomon,

766

767

768

769

770

771

772

773

774

775

776

777

778

779

780

781

782

783

784

785

786

787

788

789

790

791

792

793

794

795

Page 29: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M., Miller, H.L. (Eds.),

Cambridge, United Kingdom and New York, NY, USA.

Iverson, S.J., Field, C., Bowen, W.D., Blanchard, W., 2004. Quantitative fatty acid signature analysis:

A new method of estimating predator diets. Ecological Monographs 74 (2), 211-235.

Jumars, P.A., Mayer, L.M., Deming, J.W., Baross, J.A., Wheatcroft, R.A., 1990. Deep-sea deposit-

feeding strategies suggested by environmental and feeding constraints. Philosophical Transactions

of the Royal Society of London Series A-Mathematical Physical and Engineering Sciences 331

(1616), 85-101.

Kanzog, C., Ramette, A., Queric, N.V., Klages, M., 2009. Response of benthic microbial communities

to chitin enrichment: an in situ study in the deep Arctic Ocean. Polar Biology 32 (1), 105-112.

Kiriakoulakis, K., Stutt, E., Rowland, S.J., Vangriesheim, A., Lampitt, R.S., Wolff, G.A., 2001. Controls

on the organic chemical composition of settling particles in the Northeast Atlantic Ocean. Progress

in Oceanography 50 (1-4), 65-87.

Kones, J.K., Soetaert, K., van Oevelen, D., Owino, J.O., 2009. Are network indices robust indicators of

food web functioning? a Monte Carlo approach. Ecological Modelling 220, 370–382.

Köster, M., Jensen, P., Meyer-Reil, L.A., 1991. Hydrolytic activity associated with organisms and

biogenic structures in deep-sea sediments. In: Chrost, R. (Ed.), Microbial Enzymes in Aquatic

Environments. Springer Verlag, Berlin, pp. 298-310.

Kröncke, I., 1994. Macrobenthos composition, abundance and biomass in the Arctic Ocean along a

transect between Svalbard and the Makarov Basis. Polar Biology 14 (8), 519-529.

Mayer, L.M., Schick, L.L., Sawyer, T., Plante, C.J., Jumars, P.A., Self, R.L., 1995. Bioavailable amino-

acids in sediments - A biomimetic, kinetics-based approach. Limnology and Oceanography 40 (3),

511-520.

Meziane, T., Bodineau, L., Retiere, C., Thoumelin, G., 1997. The use of lipid markers to define

sources of organic matter in sediment and food web of the intertidal salt-marsh-flat ecosystem of

Mont-Saint-Michel Bay, France. Journal of Sea Research 38 (1-2), 47-58.

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

Geochimica Et Cosmochimica Acta 53 (7), 1577-1581.

796

797

798

799

800

801

802

803

804

805

806

807

808

809

810

811

812

813

814

815

816

817

818

819

820

821

822

823

Page 30: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

Moodley, L., Middelburg, J.J., Boschker, H.T.S., Duineveld, G.C.A., Pel, R., Herman, P.M.J., Heip,

C.H.R., 2002. Bacteria and Foraminifera: Key players in a short-term deep-sea benthic response to

phytodetritus. Marine Ecology Progress Series 236, 23-29.

Moore, J.C., Berlow, E.L., Coleman, F.C., De Ruiter, P.C., Dong, Q., Hastings, A., Collins, N.,

McCann, K.S., Melville, K., Morin, P.J., Nadelhoffer, K., Rosemond, A.D., Post, D.M., Sabo, J.L.,

Scow, K.M., Vanni, M.J., Wall, D.H., 2004. Detritus, trophic dynamics and biodiversity. Ecology

Letters 7, 584-600.

Moran, S.B., Kelly, R.P., Hagstrom, K., Smith, J.N., Grebmeier, J.M., Cooper, L.W., Cota, G.F., Walsh,

J.J., Bates, N.R., Hansell, D.A., Maslowski, W., Nelson, R.P., Mulsow, S., 2005. Seasonal changes

in POC export flux in the Chukchi Sea and implications for water column-benthic coupling in Arctic

shelves. Deep-Sea Research Part II-Topical Studies in Oceanography 52 (24-26), 3427-3451.

Pfannkuche, O., 2005. Allochtonous deep-sea benthic communities: functioning and forcing. In:

Kristensen, E., Haese, R.R., Kostka, J.E. (Eds.), Interactions between macro- and microorganisms

in marine sediments. American Geophysical Union, Washington DC, pp. 251-266.

Piepenburg, D., Blackburn, T.H., Vondorrien, C.F., Gutt, J., Hall, P.O.J., Hulth, S., Kendall, M.A.,

Opalinski, K.W., Rachor, E., Schmid, M.K., 1995. Partitioning of benthic community respiration in

the Arctic (northwestern Barents Sea). Marine Ecology-Progress Series 118 (1-3), 199-213.

Pörtner, H.O., Berdal, B., Blust, R., Brix, O., Colosimo, A., De Wachter, B., Giuliani, A., Johansen, T.,

Fischer, T., Knust, R., Lannig, G., Naevdal, G., Nedenes, A., Nyhammer, G., Sartoris, F.J.,

Serendero, I., Sirabella, P., Thorkildsen, S., Zakhartsev, M., 2001. Climate induced temperature

effects on growth performance, fecundity and recruitment in marine fish: developing a hypothesis

for cause and effect relationships in Atlantic cod (Gadus morhua) and common eelpout (Zoarces

viviparus). Continental Shelf Research 21 (18-19), 1975-1997.

Pusceddu, A., Bianchelli, S., Canals, M., Sanchez-Vidal, A., Durrieu De Madron, X., Heussner, S.,

Lykousis, V., de Stigter, H., Trincardi, F., Danovaro, R., 2010. Organic matter in sediments of

canyons and open slopes of the Portuguese, Catalan, Southern Adriatic and Cretan Sea margins.

Deep Sea Research Part I: Oceanographic Research Papers 27, 441-457.

Renaud, P.E., Morata, N., Ambrose, W.G., Bowie, J.J., Chiuchiolo, A., 2007. Carbon cycling by

seafloor communities on the eastern Beaufort Sea shelf. Journal of Experimental Marine Biology

and Ecology 349 (2), 248-260.

824

825

826

827

828

829

830

831

832

833

834

835

836

837

838

839

840

841

842

843

844

845

846

847

848

849

850

851

852

853

Page 31: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

Rex, M.A., Etter, R.J., Morris, J.S., Crouse, J., McClain, C.R., Johnson, N.A., Stuart, C.T., Deming,

J.W., Thies, R., Avery, R., 2006. Global bathymetric patterns of standing stock and body size in the

deep-sea benthos. Marine Ecology-Progress Series 317, 1-8.

Rowe, G.T., Wei, C.L., Nunnally, C., Haedrich, R., Montagna, P., Baguley, J.G., Bernhard, J.M.,

Wicksten, M., Ammons, A., Briones, E.E., Soliman, Y., Deming, J.W., 2008. Comparative biomass

structure and estimated carbon flow in food webs in the deep Gulf of Mexico. Deep-Sea Research

Part II-Topical Studies in Oceanography 55 (24-26), 2699-2711.

Ruhl, H.A., Ellena, J.A., Smith, K.L., 2008. Connections between climate, food limitation, and carbon

cycling in abyssal sediment communities. Proceedings of the National Academy of Sciences of the

United States of America 105 (44), 17006-17011.

Schluter, M., Sauter, E.J., Schafer, A., Ritzrau, W., 2000. Spatial budget of organic carbon flux to the

seafloor of the northern North Atlantic (60 degrees N-80 degrees N). Global Biogeochemical

Cycles 14 (1), 329-340.

Slagstad, D., Ellingsen, I., Wassmann, P., 2010. Primary and secondary production in a future Arctic

Ocean without summer sea ice. Progress In Oceanography.

Smith, K.L., Ruhl, H.A., Bett, B.J., Billett, D.S.M., Lampitt, R.S., Kaufmann, R.S., 2009. Climate,

carbon cycling, and deep-ocean ecosystems. Proceedings of the National Academy of Sciences of

the United States of America 106 (46), 19211-19218.

Soetaert, K., Franco, M., Lampadariou, N., Muthumbi, A., Steyaert, M., Vandepitte, L., vanden Berghe,

E., Vanaverbeke, J., 2009. Factors affecting nematode biomass, length and width from the shelf to

the deep sea. Marine Ecology-Progress Series 392, 123-132.

Soetaert, K., Van Oevelen, D., 2008. LIM: Linear Inverse Model examples and solution methods. R

package version 1.2.

Soetaert, K., Van Oevelen, D., 2009. Modeling food web interactions in benthic deep-sea ecosystems:

a practical guide. Oceanography 22 (1), 130-145.

Soltwedel, T., Bauerfeind, E., Bergmann, M., Budaeva, N., Hoste, E., Jaeckisch, N., von Juterzenka,

K., Matthiessen, J., Mokievsky, V., Nöthig, E.M., Quéric, N.V., Sablotny, B., Sauter, E., Schewe, I.,

Urban-Malinga, B., Wegner, J., Wlodarska-Kowalczuk, M., Klages, M., 2005. HAUSGARTEN:

Multidisciplinary investigations at a deep-sea, long-term observatory in the Arctic Ocean.

Oceanography 18 (3), 46-61.

854

855

856

857

858

859

860

861

862

863

864

865

866

867

868

869

870

871

872

873

874

875

876

877

878

879

880

881

882

883

Page 32: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

Soltwedel, T., Jaeckisch, N., Ritter, N., Hasemann, C., Bergmann, M., Klages, M., 2009. Bathymetric

patterns of megafaunal assemblages from the arctic deep-sea observatory HAUSGARTEN. Deep-

Sea Research Part I-Oceanographic Research Papers 56 (10), 1856-1872.

Stephens, M.P., Kadko, D.C., Smith, C.R., Latasa, M., 1997. Chlorophyll-a and pheopigments as

tracers of labile organic carbon at the central equatorial Pacific seafloor. Geochimica Et

Cosmochimica Acta 61 (21), 4605-4619.

Sun, M.Y., Aller, R.C., Lee, C., 1991. Early diagenesis of chlorophyll-a in Long Island Sound

sediments: A measure of carbon flux and particle reworking. Journal of Marine Research, pp. 379-

401.

Ulanowicz, R.E., 2004. Quantitative methods for ecological network analysis. Computational Biology

and Chemistry 28 (5-6), 321-339.

Urban-Malinga, B., Moens, T., 2006. Fate of organic matter in Arctic intertidal sediments: Is utilisation

by meiofauna important? Journal of Sea Research 56 (3), 239-248.

Van den Meersche, K., Soetaert, K., Van Oevelen, D., 2009. xsample(): an R function for sampling

linear inverse problems. Journal of Statistical Software 30 (1), 1-15.

Van Oevelen, D., Duineveld, G.C.A., Lavaleye, M.S.S., Mienis, F., Soetaert, K., Heip, C.H.R., 2009.

The cold-water coral community as hotspot of carbon cycling on continental margins: a food web

analysis from Rockall Bank (northeast Atlantic). Limnology and Oceanography 54 (6), 1829–1844.

Van Oevelen, D., Soetaert, K., García, R., De Stigter, H., Cunha, M.R., Pusceddu, A., Danovaro, R.,

2011. Canyon conditions impact carbon flows in food webs of three sections of the Nazaré canyon.

Deep Sea Research Part II: Topical Studies in Oceanography.

Van Oevelen, D., Soetaert, K., Middelburg, J.J., Herman, P.M.J., Moodley, L., Hamels, I., Moens, T.,

Heip, C.H.R., 2006. Carbon flows through a benthic food web: Integrating biomass, isotope and

tracer data. Journal of Marine Research 64 (3), 1-30.

Van Oevelen, D., Van den Meersche, K., Meysman, F., Soetaert, K., Middelburg, J.J., Vézina, A.F.,

2010. Quantitative reconstruction of food webs using linear inverse models. Ecosystems 13, 32–

45.

Vanaverbeke, J., Arbizu, P.M., Dahms, H.U., Schminke, H.K., 1997. The Metazoan meiobenthos

along a depth gradient in the Arctic Laptev Sea with special attention to nematode communities.

Polar Biology 18 (6), 391-401.

884

885

886

887

888

889

890

891

892

893

894

895

896

897

898

899

900

901

902

903

904

905

906

907

908

909

910

911

912

913

Page 33: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

Vanreusel, A., Clough, L., Jacobsen, K., Ambrose, W., Jivaluk, J., Ryheul, V., Herman, R., Vincx, M.,

2000. Meiobenthos of the central Arctic Ocean with special emphasis on the nematode community

structure. Deep-Sea Research Part I-Oceanographic Research Papers 47 (10), 1855-1879.

Wassmann, P., Slagstad, D., Riser, C.W., Reigstad, M., 2006. Modelling the ecosystem dynamics of

the Barents Sea including the marginal ice zone II. Carbon flux and interannual variability. Journal

of Marine Systems 59 (1-2), 1-24.

Westrich, J.T., Berner, R.A., 1984. The role of sedimentary organic matter in bacterial sulfate

reduction: The G model tested. Limnology and Oceanography 29 (2), 236-249.

Wieser, W., 1953. Die Beziehung zwischen Mundhöhlengestalt, Ernährungsweise und Vorkommen bei

freilebenden marinen Nematoden. Eine skologisen-morphologische studie. Arkiv für Zoologie 4,

439-484.

Wigham, B.D., Galley, E.A., Smith, C.R., Tyler, P.A., 2008. Inter-annual variability and potential for

selectivity in the diets of deep-water Antarctic echinoderms. Deep-Sea Research Part II-Topical

Studies in Oceanography 55 (22-23), 2478-2490.

Wigham, B.D., Hudson, I.R., Billett, D.S.M., Wolff, G.A., 2003. Is long-term change in the abyssal

Northeast Atlantic driven by qualitative changes in export flux? Evidence from selective feeding in

deep-sea holothurians. Progress in Oceanography 59 (4), 409-441.

Witbaard, R., Duineveld, G.C.A., Kok, A., van der Weele, J., Berghuis, E.M., 2001. The response of

Oneirophanta mutabilis (Holothuroidea) to the seasonal deposition of phytopigments at the

porcupine Abyssal Plain in the Northeast Atlantic. Progress in Oceanography 50 (1-4), 423-441.

Witbaard, R., Duineveld, G.C.A., Van der Weele, J.A., Berghuis, E.M., Reyss, J.P., 2000. The benthic

response to the seasonal deposition of phytopigments at the Porcupine Abyssal Plain in the North

East Atlantic. Journal of Sea Research 43 (1), 15-31.

Witte, U., Wenzhofer, F., Sommer, S., Boetius, A., Heinz, P., Aberle, N., Sand, M., Cremer, A.,

Abraham, W.R., Jorgensen, B.B., Pfannkuche, O., 2003. In situ experimental evidence of the fate

of a phytodetritus pulse at the abyssal sea floor. Nature 424 (6950), 763-766.

Wlodarska-Kowalczuk, M., Pearson, T.H., 2004. Soft-bottom macrobenthic faunal associations and

factors affecting species distributions in an Arctic glacial fjord (Kongsfjord, Spitsbergen). Polar

Biology 27 (3), 155-167.

914

915

916

917

918

919

920

921

922

923

924

925

926

927

928

929

930

931

932

933

934

935

936

937

938

939

940

941

942

Page 34: Structure and stability of the benthic food web at … … · Web viewCarbon flows in the benthic food web at the deep-sea observatory HAUSGARTEN (Fram Strait) Dick van Oevelen1,*,

Woulds, C., Andersson, J.H., Cowie, G.L., Middelburg, J.J., Levin, L.A., 2009. The short-term fate of

organic carbon in marine sediments: Comparing the Pakistan margin to other regions. Deep-Sea

Research Part II-Topical Studies in Oceanography 56 (6-7), 393-402.

Woulds, C., Cowie, G.L., Levin, L.A., Andersson, J.H., Middelburg, J.J., Vandewiele, S., Lamont, P.A.,

Larkin, K.E., Gooday, A.J., Schumacher, S., Whitcraft, C., Jeffreys, R.M., Schwartz, M., 2007.

Oxygen as a control on seafloor biological communities and their roles in sedimentary carbon

cycling. Limnology and Oceanography 52 (4), 1698-1709.

943

944

945

946

947

948

949

950

951


Recommended