• Non ci sono risultati.

Effects of sub-seabed CO2 leakage: short- and medium-term responses of benthic macrofaunal assemblages

N/A
N/A
Protected

Academic year: 2021

Condividi "Effects of sub-seabed CO2 leakage: short- and medium-term responses of benthic macrofaunal assemblages"

Copied!
18
0
0

Testo completo

(1)

1 2 3 4 5 6

Effects of sub-seabed CO2 leakage: short- and medium-term responses of benthic macrofaunal 7 assemblages 8 9 10 11 12 13 14

Amaro, T.1,2,3*, Queiros, A.M.4,Rastelli, E.3,5, Borgersen, G.2, Brkljacic, M.2, Nunes, J.4, Bertocci, I. 3, , Sorensen, 15 K.2, Danovaro, R.3,5, Widdicombe, S.4 16 17 18 19 20 21 22 23 24 25 26 27 28

*Corresponding author: teresa.amaro@szn.it

29

1Hellenic Center for Marine Research (HCMR), 710 03 Heraklion, Crete, Greece. 30

2 Norwegian Institute for Water Research, Oslo, Norway. 31

3 Stazione Zoologica Anton Dohrn, Villa Comunale, Naples, Italy 32

4 Plymouth Marine Laboratory, Prospect Place, West Hoe, PL1 3DH, Plymouth, UK

33

5 Department of Life and Environmental Sciences, Polytechnic University of Marche, Ancona, 34

Italy

35

Formattato: Italiano (Italia) Formattato: Italiano (Italia)

Formattato: Inglese (Regno Unito) Formattato: Inglese (Regno Unito)

(2)

ABSTRACT 36

The continued rise in atmospheric carbon dioxide (CO2) levels is driving climate change and 37

temperature shifts at a global scale. CO2 Capture and Storage (CCS) technologies have been 38

suggested as a feasible option for reducing CO2 emissions and mitigating their effects. 39

However, before CCS can be employed at an industrial scale, any environmental risks

40

associated with this activity should be identified and quantified. Significant leakage of CO2 41

from CCS reservoirs and pipelines is considered to be unlikely, however direct and/or indirect

42

effects of CO2 leakage on marine life and ecosystem functioning must be assessed, with 43

particular consideration given to spatial (e.g. distance from the source) and temporal (e.g.

44

duration) scales at which leakage impacts could occur. In the current mesocosm experiment we

45

tested the potential effects of CO2 leakage on macrobenthic assemblages by exposing infaunal 46

sediment communities to different levels of CO2 concentration (400, 1000, 2000, 10000 and 47

20000 ppm CO2), simulating a gradient of distance from a hypothetic leakage, over short-term 48

(a few weeks) and medium-term (several months). A significant impact on community

49

structure, abundance and species richness of macrofauna was observed in the short-term

50

exposure. Individual taxa showed idiosyncratic responses to acidification. We conclude that the

51

main impact of CO2 leakage on macrofaunal assemblages occurs almost exclusively at the 52

higher CO2 concentration and over short time periods, tending to fade and disappear at 53

increasing distance and exposure time. Although under the cautious perspective required by the 54

possible context-dependency of the present findings, thisThe results of the present study

55

contributes to the cost-benefit analysis (environmental risk versus the achievement of the

56

intended objectives) of CCS strategies.

(3)

58

INTRODUCTION 59

The accelerating rise in atmospheric carbon dioxide (CO2) levels (IPCC, 2013) is 60

causing ocean warming and acidification at unprecedented rates, posing critical threats to

61

single species, habitats, oceanic regions and overall global ecosystem functioning (Caldeira

62

and Wickett, 2003; Feely et al., 2004; Hale et al., 2011; Mora et al., 2013, Cerrano et al., 2013;

63

Meadows et al., 2015; Gattuso et al., 2015). As a direct consequence, it is urgently needed to

64

identify suitable options for reducing/mitigating CO2 emissions (McCormack et al., 2016). One 65

particularly promising technology involves capturing CO2 from point source effluents (mostly, 66

energy generation plants), then transporting it as a supercritical liquid to be stored in deep

67

porous geological rock formations, such as saline aquifers or existing hydrocarbon reservoirs

68

(Gibbins et al. 2006; Holloway 2007). This process is defined as CO2 Capture and Storage 69

(CCS). In Europe and North America the technical feasibility of CCS approaches has been

70

already demonstrated. For example, at the Sleipner West gas field in the Norwegian sector of

71

the North Sea, CCS has been operational since 2000 with approximately 1 million tons of CO2 72

pumped into the storage reservoir every year (Paulley et al., 2012, Jones et al., 2015).

73

However, as with almost any other human activity, this technology is not risk-free in terms of

74

posing potential environmental hazards (reviewed by Damen et al. 2006). Before industrial

75

scale CCS activities become widely accepted and implemented these risks need to be identified

76

and quantified. Perhaps the greatest environmental risk associated with CCS is that of CO2 77

leaking into the marine environment either during transport, sequestration or from the

78

geological storage reservoir itself. Whilst current evidence suggests that leakage from CCS

79

reservoirs would be extremely unlikely it is not impossible (Blackford et al., 2009; 2014).

80

Given that any major increase in seawater CO2 concentrations , and the associated changes in 81

carbonate chemistry, has the potential to considerably impact marine life and ecosystem

82

functions, assessing the biological and ecological effects of CO2 leakage is essential to support 83

environmental risks assessments required by CCS operations (Widdicombe et al., 2013; Jones

84

et al., 2015). This is especially relevant for benthic assemblages living in the immediate

85

proximity of any potential leak, since they would be exposed to relatively large and rapid

86

changes in carbonate chemistry, in both the sediment pore waters and the overlying seawater

87

(Lichtschlag et al., 2014; Queiros et al., 2014).

88

The exposure to a range of CO2 concentrations has been tested on a variety of marine 89

organisms, as well as on some biogeochemical processes and ecosystem functions

90

(Widdicombe et al., 2013, 2015; Laverock et al., 2013; Tait et al., 2014; Rastelli et al., 2015). It

(4)

has also been demonstrated that the impact of elevated CO2 on marine organisms depends on 92

both the severity and the duration of the exposure (Blackford et al., 2013). In general, it is

93

hypothesized that a CCS leakage is immediately associated with a localized acute exposure to

94

harmful high CO2 conditions whose effects are likely to get attenuated at increasing distance 95

from the source. Moreover, more prolonged leakage or persisting influences of temporary

96

seepage of any level could represent chronic stressed conditions to the surrounding abiotic and

97

biological environment (Jones et al., 2015).

98

Whilst previous studies have started to provide a better understanding of the potential

99

impacts of CCS leakage on specific benthic organisms (e.g. Widdicombe & Needham 2007),

100

our knowledge of the possible effects at the community level remains limited (Widdicombe et

101

al. 2015). In addition, the mechanisms underlying such changes are still largely unknown, as

102

well as the difference between direct and indirect effects of increasing CO2 leakages on the 103

macrofaunal community. It has been reported, however, that low-pH levels predicted by

104

realistic scenarios of CCS leakage might severely reduce the prokaryotic-mediated processes

105

(Rastelli et al. 2015), while acidified conditions could favor blooms of benthic microbial

106

primary producers including cyanobacteria and diatoms (Tait et al., 2015). Notably, the

107

exposure to high CO2 levels can alter microbial-mediated processes able to affect the quality 108

and quantity of the sedimentary organic matter (OM) (Rastelli et al., 2015). Since the

109

availability of OM is a key driver of the abundance, distribution and biodiversity of benthic

110

fauna (Fabiano and Pusceddu, 1998, Pusceddu et al., 2009), the effects of changes in this

111

variable due to CCS leakage might indirectly propagate to associated macrofaunal

112

assemblages.

113

Full community level effects of CO2 leakages can only be unequivocally assessed 114

using simulated leakage experiments conducted in the field (e.g. Blackford et al. 2014) or from

115

studying actual leakage events or accidents in areas where data on the response variables of

116

interest are available before and after the event. Both options are normally unavailable either

117

due to the lack of data or to logistic, financial and/or ethical constraints. Performing

118

manipulative experiments in mesocosms can be a feasible alternative especially when the

119

results are used to inform ecosystem level models. A strength of an experimental approach is

120

that by exposing initially comparable assemblages to different levels of CO2 concentration 121

(such as ‘naturally’ occurring along a gradient of distance from a supposed leakage) under

122

controlled conditions allows testing for their relative effects in an unconfounded way. 123

In this study, we performed a mesocosm experiment to test the potential impact of

124

CO2-enriched (from 400 ppm to 20000 ppm) seawater plumes on the abundance and diversity

(5)

of soft-bottom macrofauna. Specifically, we tested the null hypotheses that (i) the whole

126

structure (taxon composition and relative abundance), richness, total abundance of the

127

macrofaunal assemblages and the abundance of individual taxa, did not differed depending on

128

the CO2 concentration; (ii) such a lack of differences was consistentchanged between a few 129

weeks (short-term) and some months (medium. term) of continued exposure.

130 131

MATERIAL AND METHODS 132

Collection of sediment samples and associated fauna and mesocosm setup 133

Using a KC Denmark boxcorer, intact sediment samples containing natural infaunal

134

assemblages were collected during the 3rd week of August 2012 from randomly selected points 135

located some meters apart from each other at the outer Oslofjord (59°49.4788’ N, 10°58.8595’

136

E), Norway, at 100 m water depth. Each boxcorer was equipped with an inner liner, which

137

allowed the sediments and the overlying water to be retrieved with minimal disturbance.

138

A total of 46 independent liners (0.09 m2 each, with average sediment penetration of 139

~40 cm) were collected and transferred immediately to the benthic mesocosm systems at the

140

Marine Research Station, Norwegian Institute of Water Research, Solbergstrand, Norway.

141

During transportation, all liners were shaded and continuously covered with seawater to

142

prevent desiccation and minimise temperature changes.

143

The experimental system was set up according to Widdicombe et al. (2009), as

144

described in detail elsewhere (Queiros et al., 2015, Rastelli et al., 2015). Briefly, all liners were

145

placed in an aquarium in a flow-through holding basin filled with seawater to a depth of 1 m

146

(mesocosm) and supplied continuously with unfiltered natural seawater at a flow rate of 120

147

ml/min from a pipeline situated at 60 m depth in the adjacent fjord. All liners were

148

maintained in these conditions for two weeks prior to the beginning of the experiment to allow

149

the fauna, microbes and geochemical processes to acclimatize to mesocosm conditions.

150 151

Preliminary survey 152

To guarantee that the randomly assigned experimental levels of CO2 were not 153

confounded by initial differences between replicate cores in terms of hosted macrofaunal

154

assemblages would have required us to compare macrofauna among all (allocated) treatments

155

before manipulation. Unfortunately, the needed destructive sampling made such an option

156

impossible. Alternatively, a total of 6 liners were chosen at random from the 46 liners initially

157

collected and these 6 were randomly allocated to one of two groups of three. These were then

158

compared (by means of one-way PERMANOVA, see Supplement S1) for the structure of

(6)

macrofauna, under the hypothesis that the lack of significant difference between one group and

160

the other could provide information (not exhaustive, but relevant) to assume that significant

161

differences were not likely to exist also among the sets of replicate liners allocated at random to

162

the experimental levels. In addition, data on the sediment grain size, estimated by laser analysis 163

at the beginning of the experiment, were available for was also estimated by laser analysis one 164

liner per each of the total five experimental conditions (e.g. McCave, 2013).

165

For each liner, macrofaunal assemblages were sampled, after the acclimation period, by

166

sieving all the sediment over a 500 μm mesh, with the residue from each sample being fixed in

167

10% buffered formalin until further processing. In the laboratory, the fauna was extracted from

168

the residue under a binocular microscope and all specimens were sorted into major taxa and

169

then identified to species level whenever possible. Species (or higher taxa) abundance was

170

determined in each replicate and expressed as the total number of individuals per m2 of 171

sampled area.

172 173

Experimental setup and sampling 174

The 40 liners remaining after the preliminary survey were randomly allocated in equal

175

numbers (4) to each of five CO2 treatments: 400 (control), 1000, 2000, 5000, and 20000 ppm, 176

with two sampling times (2 weeks, 20 weeks). These levels were consistent with those

177

specifically tested by Rastelli et al. (2015) and Queiros et al. (2015). Seawater acidification

178

was achieved as described by Widdicombe et al. (2009). Briefly, CO2 gas passed through a 179

450 L reservoir tanks filled with natural seawater. Using an automated feedback relay system

180

(Walchem), the CO2 flux into the reservoir tanks was regulated in order to maintain the 181

required pH level. The reservoir tanks were continuously supplied with natural seawater (pH∼

182

8.1).

183

To assess short-term (a few weeks scale) and medium-term (several months) effects of

184

CO2 exposure, sampling took place after 2 weeks exposure (T1) and again, on a different set of 185

replicate liners not previously sampled, after 20 weeks exposure (T2). At each time,

186

macrofaunal assemblages were sampled as described for the preliminary survey.

187

A procedural control involving the flux of air only with no CO2 enrichment could not 188

be established due to logistic constraints. Therefore, the present experimental setup cannot 189

separate the actual intended effects of CO2 treatments from the possible influence of the 190

physical disturbance by the manipulated flux of gas per se. However, the fact that the used 191

experimental device was analogous for all experimental units and conditions allowed to test for 192

the relative effects of the treatments in an unconfounded way. 193

Formattato: Pedice

Formattato: Tipo di carattere:

(7)

In each of the liners and header tanks, seawater temperature, salinity, oxygen

194

concentration and pH were monitored a total of 31 times (at 2 to 6 days intervals) during the

195

course of the experiment (October 2012 to February 2013) using macroprobes.

196 197

Statistical analyses 198

Permutational multivariate analysis of variance (PERMANOVA) was used to test for

199

the null hypothesis of no differences in the macrofaunal community structure among

200

experimental CO2 treatments and for their consistency independently ofpossible variation 201

depending on the exposure time (Anderson, 2001). The analysis was based on Bray-Curtis

202

square-root transformed dissimilarities, calculated from the whole matrix of square root-203

transformed (to reduce the weight of the most abundant taxa) abundance data, and a two-way

204

model including the crossed factors ‘Time’ (random, two levels: short- vs. medium term

205

exposure; note that treating this factor as random was driven by the fact that we did not intend

206

to examine differences precisely between the 2 and the 20 weeks exposure, but only to test for

207

the consistency of the effects of acidification treatments between ‘a few weeks’ and a ‘some

208

months’ exposure, denominated as short- and medium-term, respectively) and ‘Treatment’

209

(fixed, five levels: 400, 1000, 2000, 10000 and 20000 ppm CO2). The four liners allocated to 210

each combination of factors provided the replicates for the analysis. Since the Bray-Curtis

211

measure combines differences in both the identity and the relative abundance of taxa between

212

samples, the same analysis was repeated twice using, as the original input data matrix, either

213

square root-transformedraw abundances, or presence/absence data (Clarke & Green, 1988).

214

The PERMDISP test was used to assess whether multivariate differences among groups were 215

due to differences in the dispersion rather than in the location of centroids (Anderson, 2006). 216

Multivariate patterns were illustrated by non-metric multidimensional scaling (nMDS)

217

ordination based on Bray–Curtis dissimilarities calculated on both square-root and

presence-218

absence data.

219

The same model of analysis, but based on Euclidean distances between samples, was

220

used to test for responses to experimental treatments of the total abundance, total richness of

221

taxa and the abundance of individual conspicuous (the most common in all treatments)

222

macrofaunal taxa.

223

When relevant, post-hoc comparisons between levels of the CO2 treatment were 224

performed with paired t-tests. All analyses were carried out using the PRIMER 6.0 &

225

PERMANOVA+ β 3 package (Anderson et al., 2008).

226 227

(8)

RESULTS 228

Preliminary survey and effectiveness of experimental treatments 229

The PERMANOVA performed on six liners before the start of the experiment did not

230

detect any significant differences in the structure of macrofaunal assemblages between

231

replicates belonging to each of two randomly established groups (MS = 2717.6, pseudo-F1,4 = 232

2.3, p>0.1, full details are reported in Appendix S1).

233

The chosen levels of CO2 concentration were capable of producing clear differences in 234

pH between treatments (Appendix S2; see also Queiros et al. 2015, Rastelli et al., 2015). On

235

the contrary, temperature, salinity and O2 values were maintained considerably constant and 236

comparable across all liners independently of the treatment (Appendix S2; see also Queiros et

237

al. 2015, Rastelli et al., 2015). Analogously, the sediment grain size was very similar (mean +/-

238

SE = 10.37 +/- 0.49 μm) among the five (one per experimental condition)all sampled liners

239

examined before the start of the experiment.

240 241

Macrofaunal abundance, diversity and community structure responses to increasing CO2 242

A total of 180 173 macrofaunal species or higher taxa (1026 Annelida, 34 27 Mollusca,

243

followed by 15 23 Arthropoda, 7 Sipuncula, 67 Echinodermata, 6 and Cnidaria, 1 Nemertea 244

and 1 Hemichordata - Appendix S3) were identified in the experiment and used for

245

PERMANOVA on the whole assemblage structure.

246

Macrofaunal assemblages changed between experimental conditions depending on

247

time, irrespectively of the square root or the presence/absence transformation (Table 1). At 2

248

weeks of exposure, pairwise tests indicated a significant difference between the control and all

249

treatments, but the 20000 ppm CO2 treatment. At 20 weeks, the only significant difference was 250

between the control and the highest CO2 treatment (Table 1 and Fig. 1 A, B). However, both 251

MDS ordination plots based on square root- and presence/absence-transformed data did not

252

show a clear separation between centroids corresponding to each treatment and exposure time, 253

while the dispersion of the points representing assemblages exposed to the 20 weeks exposure 254

was clearly larger than that of points corresponding to assemblages exposed to the shorter 255

exposure (Fig. 1 A and PERMDISP: F = 62.1, p = 0.001; Fig. 1 B and PERMDISP: F = 55.4, p 256

= 0.001).

257

Both total richness of taxa and total abundance of individuals differed among

258

treatments depending on the exposure time (Table 2). Specifically, 2 weeks after the start of the

259

experiment, the control hosted a larger number of taxa than all treatments, but the 20000 ppm,

260

Formattato: Pedice

Formattato: Non Evidenziato Formattato: Non Evidenziato Formattato: Non Evidenziato Formattato: Non Evidenziato Formattato: Non Evidenziato Formattato: Non Evidenziato

(9)

while no significant differences were displayed at 20 weeks of exposure (Table 2 and Fig. 2

261

A). The total number of individuals per m2 at 2 weeks was also higher in the control than in 262

any treatment, but significant differences were detected only relative to the 2000 and 5000 ppm

263

CO2 concentrations. Analogously to richness, all significant differences disappeared after 20 264

weeks exposure (Table 2 and Fig. 2 B).

265

The analysis performed on the abundance of the 8 most common macrofaunal taxa

266

were tested between the different experimental conditions over time. Specifically, after two

267

weeks of exposure, the abundance of the polychaete Heteromastus filiformis was higher, 268

although not significantly, higher in the control than in all CO2 treatments. At 20 weeks, this 269

species was less abundant in the control than in the two highest concentrations (Table 3 and

270

Fig. 3 A). Another polychaete, Prionospio cirrifera, was, on the other hand, significantly

271

moreless abundant in the control than in either the 20000 ppm treatment or all treatments after

272

two and twenty weeks of exposure, while no significant differences in the abundance of this 273

species occurred at twenty weeks of exposure respectively (Table 3 and Fig. 3 B).

274

Three taxa, namely the Nemertea, and the polychaetes Paradoneis eliasoni/lyra and

275

Paramphinome jeffreysii differed significantly between times of exposure, irrespectively of the 276

CO2 concentration (Table 3), with the first two taxaon being, on average, lessmore abundant 277

after 2 than after 20 weeks of exposure, and the other two species displaying the opposite

278

pattern (Fig. 3 C, D and E).

279

Of tThe remaining three taxa,(the polychaete Chaetozone sp. and the bivalves Thyasira 280

equalis and Adontorhina similis and the polychaete Chaetozone sp. were not significantly 281

affected by any CO2 treatment applied over any time (Table 3 and Fig. 3 F and. G and H), 282

while the bivalve Thyasira equalis was comparably abundant in each CO2 treatment at 2 283

weeks exposure and completely absent at 20 weeks exposure (Table 3 and Fig. 3 H).

284 285

DISCUSSION 286

The present study was designed to investigate the impact on the abundance and

287

diversity of benthic macrofaunal assemblages of exposure to a plume of CO2-enriched 288

seawater that could result from CO2 leakages from sub-seabed CCS. Results indicated that over 289

a short-term period (2 weeks), the macrofaunal assemblage structure was significantly affected

290

by all experimental levels of increased CO2, with the only exception of the highest 291

concentration. Conversely, after 20 weeks of exposure, the only significant difference was

292

between the control assemblages and those subject to the highest CO2 concentration. 293

(10)

Rapid impacts on macrofauna community structure, diversity and abundance following

294

short-term exposure to elevated CO2, similar to that seen in the current experiment, has been 295

reported from a number of previous mesocosm studies (e.g. Widdicombe et al., 2009,

296

Meadows et al., 2015). In this study, after 2 weeks of exposure, the control treatments hosted a

297

larger number of taxa than all treatments, but the 20000 ppm. Similarly, the total abundance

298

was larger in the control than in any treatment, but significant differences occurred only

299

relative to the 2000 and 5000 ppm CO2 concentrations. The apparent lack of impact in the 300

20,000 ppm treatments is perhaps surprising. However, one explanation could be that many of

301

the organisms in this treatment had actually died as a result of this extremely high CO2 302

exposure but their bodies had not had time to decay, especially if the microbial decomposition

303

was also inhibited by the low pH, and these organisms were then falsely counted as living in

304

the subsequent analysis. Another possibility is that under very extreme CO2 exposure 305

organisms go into a severe state of metabolic depression that maintains them for a limited

306

period of time before they inevitably die. The rapid response of all CO2 enriched treatments, 307

except the 20,000 ppm treatment discussed above, indicates that the most sensitive species

308

were likely affected negatively by even relatively low CO2 treatments injected just for 2 weeks. 309

Such conditions may have led to behavioural and/or metabolic changes, ultimately leading to

310

mortality and, consequently, to changes in the whole benthic community composition.

311

Notably, previous studies have also reported a rapid, negative impact on macrofaunal

312

diversity and structure from a controlled experimental release of CO2 from below the seafloor . 313

Even though this response only became evident five weeks after the start of the release it took

314

several weeks for the within sediment porewater pH to drop significantly. This was due to

315

natural chemical buffering processes which affected the carbonate dynamics (Lichtschlag et al.,

316

2014; Taylor et al., 2015; Widdicombe et al., 2015). In the present study, it seems that there

317

was less potential for the sediment buffer the changes in seawater chemistry and the impacts on

318

infauna occurred rapidly. This highlights the importance of understanding how the different

319

chemical and biological characteristics of different sediments will affect the speed of impact

320

following a CO2 leak. 321

The results observed after medium-term exposure (20 weeks) in the current study

322

would suggest that the only impacts of prolonged CO2 exposure were observed in the highest 323

treatment level (20,000 ppm). This is in contradiction to previous mesocosm studies that have

324

shown significant impacts of CO2 exposure to persist over many weeks (e.g. Widdicombe et al 325

2009). However, it should be noted that in the current study the similarity observed between

326

the control treatments and the majority of CO2 exposure treatments after 20 weeks was not due 327

(11)

to any recovery in the fauna of the CO2 treatments but due to a decrease in the abundance and 328

diversity of the fauna in the control treatments. So it could be hypothesized that the similarity

329

of assemblages exposed to almost all treatments after the twenty weeks exposure was due, at

330

least in part, to the negative effect of holding this particular fauna under mesocosm conditions.

331

Such negative impacts could result from limiting food availability (e.g. Guppy and Withers,

332

1999), which, once maintained over or occurred after a relatively long period, could have

333

exerted a negative influence on macrofaunal assemblages able to mask any concomitant effect

334

of CO2. Unfortunately, we do not have empirical data suitable to unambiguously support or 335

discard this hypothesis. Other stressful environmental variables, such as temperature,

336

desiccation, anoxia and hypersalinity, which are capable of inducing drastic reductions in

337

metabolic rates of almost all animal taxa (Guppy and Withers, 1999), could also have, in

338

principle, occurred in mesocosms and played a role in the present findings. The continuous

339

supply of mesocosms with new water from the adjacent fjord, however, suggests that reaching

340

drastically limiting conditions of such variables was also unlikely during the experiment. In

341

addition, the previous mesocosm experiments of Widdicombe et al (2009) used similar

342

conditions as used in the current study and saw no evidence of detrimental mesocosm impacts

343

during a 20 week experiment. It is most likely therefore that the sediment or community

344

selected for this experiment was less suitable for mesocosm experimentation than that which

345

was used in the previous studies. In the current experiment the sediment was collected from an

346

area twice as deep than the area used for collection of materials in Widdicombe et al (2009);

347

100m compared with 50m. Given that many potential CCS sites are located in deep water, it

348

may be that the value of mesocosm experiments may be limited to assessing short term

349

exposures and that there is a greater need for developing in-situ experimental procedures in

350

these areas.

351

Despite the issues associated with mesocosm effects, it was clear that over a longer

352

exposure, the communities in all the CO2 enriched treatments, except 20,000 ppm, converged 353

as the hardiest and most resistant species persisted (the number of taxa that were absent in all 354

treatments almost doubled from the 2 weeks to the 20 weeks time). What actually constitutes a

355

resistant species to elevated levels of CO2 will depend on the specific metabolic and 356

physiological adaptations of macrofaunal organisms (see Widdicombe and Spicer, 2008). This

357

resistance, therefore, is largely variable among taxa, both in terms of overall extent and

358

underlying mechanisms (Lessin et al., 2016). Echinoderms show very little compensation for

359

hypercapnia-related disturbance (Spicer et al., 1988, Spicer 1995, Kroeker et al., 2013).

360

Calcifying organisms need to increase pH (by active removal of H+ ions) in order to maintain 361

(12)

the formation of biogenic structures where needed (Widdicombe et al., 2015). Other organisms

362

prefer to suppress metabolism by shutting down various cellular processes (Guppy and

363

Withers, 1999, Widdicombe et al., 2009). In this study, the exposure to the most extreme CO2 -364

leakage scenario, in a medium-term (20 weeks), was tolerated by highly resistant taxa, such as

365

borrowing polychaete worms from the family Capitellidae (to which H. filiformis belongs). In

366

fact, this taxon is described as opportunistic able to dominate macrofaunal invertebrate

367

communities under perturbed conditions, likely due to its short generation time and direct

368

development which can allow their efficient use of the habitat (Pearson and Rosenberg, 1978;

369

Berge, 1990; Preckler, 2015) and increases in biomass after the elimination of more sensitive

370

species (Lessin et al., 2016). Analogously, the spionid polychaete P. cirrifera, which was here 371

also more abundant in the most acidified treatment compared to the control, especially after 372

short-term exposure, was described among the dominant species in polluted areas (Shen et al., 373

2010). This ability has been also explored for restoring polluted sediments by adding

374

bioturbating species of capitellid polychaetes (Chareonpanich et al. 1994, Ueda et al. 1994).

375

Therefore, the fact that we did not observe a reduced abundance of H. filiformis in the most

376

acidified treatments compared with the control is consistent with its known

disturbance-377

resistant trait as a capitellid polychaete. At the same time, this species was the most abundant

378

in the examined macrofaunal samples, hence its response to experimental treatments could

379

have driven, at least for a considerable part, that of the whole structure of assemblages. It is

380

worth noting, however, that the convergence between more acidified and control assemblages

381

could have been also modulated by the temporal variability (obviously due to processes other

382

than changes in CO2 inputs) of the latter ones, which can be as large as that driven by the 383

increased CO2 inputs (Widdicombe et al., 2015). In principle, temporal fluctuations in patterns 384

of abundance and diversity of meiofaunal macrofaunal assemblages in the control could have

385

made them similar, even just by chance, to the treated ones during the experiment. A large and

386

significant temporal variability, irrespectively of CO2 treatments, was confirmed by several 387

conspicuous taxa here examined.

388

Although the results of the current mesocosm study can provide crucial information on

389

actual cause-effect relationships between CO2 enrichment and macrofaunal responses, any 390

attempt to extrapolate them to predicting the ecological and biological consequences of

391

possible field leaks should be made with caution. The main reasons for this include: a) the

392

mesocosms being a confined system, which does not allow an organism to escape or relocate to

393

avoid unfavourable conditions, potentially leading to overestimate mortality rates over scales

394

larger than the experimental one; b) the likelihood that the response is specific for the

(13)

examined assemblage, which, in spite of even analogous main traits, would not be necessarily

396

the same as another one from a different location and/or time; c) possible different buffering

397

effects due to the different mineralogy of the sediments, with special focus on carbonate

398

content; d) the potential change between mesocosms and field biological responses due to the

399

drastic difference in the depth (hydrostatic pressure) to which the system is exposed; e) the

400

lower resilience or recovery of communities due to the lack of immigration of

401

specimens/species from surrounding non-impacted systems (Danovaro 2010; Widdicombe et

402

al., 2015). Moreover, present findings cannot obviously provide any unambiguous information

403

to derive expectations on possible responses, even of the same assemblages, to longer-term (>

404

20 weeks) exposure, or on responses to any exposure of assemblages dominated by other

405

groups of organisms, such as echinoderms (Spicer et al. 1988, Spicer 1995, Kroeker et al.,

406

2013) and more calcified taxa. For example, it is reasonable to assume that calcifying

407

organisms would be particularly sensitive to increases of CO2 and consequent reductions of pH, 408

which could eventually lead to critical loss of their fitness and survival rates through the

409

allocation of more energy to ion removal processes to detriment of other important

410

physiological processes (Pörtner, 2008; Wood et al., 2008). In this context, other studies

411

carried out in natural acidic shallow vents (e.g. Rodolfo-Metalpa et al. 2011; Gambi et al.,

412

2016; Kamenos et al., 2016) indicated that even CO2 increases considerably less than in the 413

present experiment can determine profound changes in exposed benthic assemblages. None of

414

such studies, however, are fully comparable to the present one. Specifically, Rodolfo-Metalpa

415

et al. (2011) focused on calcifying organisms, which, instead, were almost not represented in

416

present meiofaunal macrofaunal assemblages. Both Gambi et al. (2016) and Kamenos et al.

417

(2016) examined the distribution and diversity of benthic organisms (coralline algae and

418

polychaetes) at increasing distance and increasing pH from natural vents, thus along natural

419

gradients of CO2 concentrations to which such organisms were adapted for a much longer time 420

compared to the temporal scale of our experiment.

421 422

CONCLUSION 423

In spite of all the above listed factors and processes which are likely to jeopardise the

424

accuracy of extrapolations of mesocosm findings to real circumstances, the structure of the

425

present experiment was suitable to examine the relative responses of macrofaunal assemblages

426

and individual taxa to increased CO2 inputs in an unconfounded way. As such, present findings, 427

although not guaranteeing that field responses to possible future leakages associated to CCS

428

strategies will be exactly the same, reasonably suggest that the main significant impact of such

(14)

events on macrofaunal assemblages would occur close to the hypothetical source of CO2 and 430

would occur rapidly over short time periods (<2 weeks).. Additional experiments are needed to

431

understand the mechanisms responsible for the present findings and their possible consistency

432

under field conditions, but this controlled study provides a relevant contribution to the debate

433

on the cost-benefit balance (environmental risk vs. intended goals) of CCS technologies.

434 435

ACKNOWLEDGEMENTS 436

437

This research was conducted as part of the European Community’s Seventh Framework

438

Programme FP7/2007-2013 for the project Sub-seabed CO2 storage: impact on marine

439

ecosystems (ECO2), grant agreements N. 265847, DEVelopment Of innovative Tools for 440

understanding marine biodiversity and assessing good Environmental Status (DEVOTES),

441

grant agreement no. 308392 and FME Success. TA was partially supported by Marie Curie

442

Actions through the project CEFMED (project number 327488). We are grateful to Oddbjorn

443

Petersen, Per Ivar Johannessen, Morten Schaanning personnel at the Marine Research Station

444

(Solbergstrand, Norway) of the Norwegian Institute of Water Research (NIVA, Oslo, Norway)

445

and at Plymouth Marine Laboratory for support and advice during the ECO2 mesocosm 446

experiments. Dr Mats Walday at NIVA, Dr Andrew Sweetman from Heriot Watt University of

447

Edinburgh and Dr Sarah Dashfield and Dr Carolyn Harris from Plymouth Marine Laboratory

448

are also thanked for support during the organization and analyses for this work.

449

450

REFERENCES 451

1. Anderson, M. J. A new method for non-parametric multivariate analysis of variance.

452

Austral. Ecol. 26, 32–46 (2001). 453

454

2. Anderson M.J. 2006. Distance-based tests for homogeneity of multivariate dispersions. 455

Biometrics 62: 245-253. 456

457

32. Anderson, M. J., Gorley, R. N. & Clarke, K. R. PERMANOVA + for PRIMER: Guide to

458

Software and Statistical Methods. PRIMER-E (Plymouth, 2008). 459

460

43. Berge, J.A. Macrofauna recolonization of subtidal sediments. Experimental studies on

461

defaunated sediment contaminated with crude oil in two Norwegian fjords with unequal

462

eutrophication status. I. Community responses. Mar. Ecol. Prog. Ser. 66, 103-115 (1990).

463 464

54. Blackford, J. C., Jones, N., Proctor, R., Holt, J., Widdicombe, S., Lowe, D. & Rees, A. An

465

initial assessment of the potential environmental impact of CO2 escape from marine carbon

466

capture and storage systems. Proc. Inst. Mech. Eng. Part A – J. Power Energy 223, 269–280

467

(2009).

468 469

Formattato: Tipo di carattere:

Times New Roman, 12 pt

Formattato: Tipo di carattere:

(15)

65. Blackford, J. C., Hattam, C., Widdicombe, S., Burnside, N., Naylor, M., Kirk, K., Maul, P.

470

& Wright, I. CO2 leakage from geological storage facilities: environmental, societal and 471

economic impacts, monitoring and research strategies. In Geological storage of carbon dioxide

472

(CO2): Geoscience, technologies, environmental aspects and legal frameworks (ed. Gluyas, J.

473

& Mathias, S.) 149-178 (Swaston, Cambridge, GB, Woodhead publishing, 2013).

474 475

76. Blackford, J. C., Stahl, H., Bull, J. M. , Bergès, B. J. P., Cevatoglu, M., Lichtschlag, A.,

476

Connelly, D., James, R. H. , Kita, J., Long, D., Naylor, M., Shitashima, K., Smith, D., Taylor,

477

P., Wright, I., Akhurst, M., Chen, B., Gernon, T. M., Hauton, C., Hayashi, M., Kaieda, H.,

478

Leighton, T. G., Sato, T., Sayer, M. D. J., Suzumura, M., Tait, K.,Vardy, M. E., White, P. R. &

479

Widdicombe, S. Detection and impacts of leakage from sub-seafloor deep geological carbon

480

dioxide storage. Nat. Clim. Change 4, 1011–1016 (2014).

481 482

87. Caldeira, K. & Wickett., M. Anthropogenic carbon and ocean pH. Nature 425:365 (2003).

483 484

98. Chareonpanich, C., Tsutsumi, H. & Montani, S. Efficiency of the decomposition of organic

485

matter, loaded on the sediment, as a result of the biological activity of Capitella sp. I.. Mar.

486

Pollut. Bull. 28, 314–318 (1994).

487 488

109. Cerrano, C., Cardini, U., Bianchelli, S., Corinaldesi, C., Pusceddu, R Danovaro. A. Red

489

coral extinction risk enhanced by ocean acidification. Scientific reports 3 (2013).

490 491

101. Clarke, K. R. & Green, R. H. Statistical design and analysis for a ‘ biological effects ’

492

study. Mar. Ecol. Prog. Ser. 46, 213 – 226 (1988).

493 494

112. Damen , K., Faaij, A. & Turkenburg, W. Health, Safety and Environmental Risks of

495

Underground Co2 Storage – Overview of Mechanisms and Current Knowledge. Climatic

496

Change 74(1), 289-318 (2006).

497 498

123. Danovaro, R. Methods for the study of deep-sea sediments, their functioning and

499

biodiversity. (ed. Taylor & Francis) 456 pp (CRC Press, London, 2010). 500

501

134. Fabiano, M. & Pusceddu, A. Total and hydrolizable particulate organic matter

502

(carbohydrates, proteins and lipids) at a coastal station in Terra Nova Bay (Ross Sea,

503

Antarctica). Polar Biol. 19, 125–132 (1998).

504 505

145. Feely, R. A., Sabine, C. L., Lee, K., Berelson ,W., Kleypas, J., Fabry V.J. & Millero, F. J.

506

Impact of anthropogenic CO2 on the CaCO3 system in the oceans. Science 305, 362–366.

507

(2004).

508 509

156. Gambi, M. C., Musco, L., Giangrande, A., Badalamenti, F., Micheli, F. & Kroeker, K. J.

510

Distribution and functional traits of polychaetes in a CO2 vent system: winners and losers

511

among closely related species. Mar. Ecol. Prog. Ser. 550, 121-134 (2016).

512 513

167. Gattuso, J. P., Magnan, A., Billé, R., Cheung, W.W., Howes, E. L., Joos, F., Allemand,

514

D., Bopp, L., Cooley, S. R., Eakin, C. M., Hoegh-Guldberg, O., Kelly, R. P., Pörtner, H. O.,

515

Rogers, A. D., Baxter, J. M., Laffoley, D., Osborn, D., Rankovic, A., Rochette, J., Sumaila, U.

516

R., Treyer, S., & Turley, C. Contrasting futures for ocean and society from different

517

anthropogenic CO2 emissions scenarios. Science 349, aac4722 (2015). 518

519

178. Gibbins, J., Haszeldine, S., Holloway, S., Pearce, J., Oakey, J., Shackley, S. & Turley, C.,

(16)

Scope for future CO2 emission reductions from electricity generation through the deployment 521

of carbon capture and storage technologies in Avoiding Dangerous Climate Change (ed.

522

Schnellhuber, H.J.) 379–384 (Cambridge University Press, 2006). 523

524

189. Guppy, M. & Withers, P. Metabolic depression in animals: physiological perspectives and

525

biochemical generalizations. Biol. Rev. Camb. Philos. Soc. 74, 1–40 (1999).

526 527

1920. Hale, R., Calosi, P., McNeill, L., Mieszkowska, N. & Widdicombe, S. Predicted levels

528

of future ocean acidification and temperature rise could alter community structure and

529

biodiversity in marine benthic communities. Oikos 120, 661–674 (2011).

530 531

201. Holloway, S. Carbon dioxide capture and geological storage. Philos. Trans. R. Soc. A

532

Math. Phys. Eng. Sci. 365 (1853), 1095–1107 (2007). 533

534

212. IPCC, 2013: Climate Change 2013: The Physical Science Basis. Contribution of Working

535

Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change 536

(eds. Stocker, T.F., et al. ). 1535 pp (Cambridge University Press, Cambridge, United Kingdom

537

and New York, NY, USA, 2013).

538 539

223. Jones, D. G., Beaubien, S. E., Blackford, J. C., Foekema, E. M., Lions, J., De Vittor, C.,

540

West, J.M., Widdicombe, S., Hauton, C & Queiros, A. Developments since 2005 in

541

understanding potential environmental impacts of CO2 leakage from geological storage. 542

International Journal of Greenhouse Gas Control 40, 350-377 (2015). 543

234. Kamenos, N. A., Perna, G., Gambi, M. C., Micheli, F. & Kroeker, K. J. Coralline algae in

544

a naturally acidified ecosystem persist by maintaining control of skeletal mineralogy and size.

545

Proc. Royal Soc. B 283, 20161159 (2016). 546

245. Kroeker, K., Kordas Rebecca Crim, R., Hendriks, I., Ramajo, L., Singh, G., Duarte, C. &

547

Gattuso, J.-P. Impacts of ocean acidification on marine organisms: quantifying sensitivities and

548

interaction with warming. Glob. Change Biol. 19, 1884–1896 (2013).

549 550

256. Laverock, B., Kitidis, V; Tait, K; Gilbert, JA; Osborn, M. & Widdicombe, S. Bioturbation

551

activity determines the response of coastal benthic nitrogen-cycling microorganisms to ocean

552

acidification. Phil. Trans. of the Royal Soc. of London Series B.-Biol. Sci. 368 (1627), 1-13

553

(2013).

554 555

267. Lessin, G., Artioli, Y., Queiros, A., Widdicombe, S. & Blackford, J.C. Modelling impacts

556

and recovery in benthic communities exposed to localized high CO2. Mar. Poll. Bull. (in press)

557

(2016).

558 559

278. Lichtschlag, A., James, R. H., Stahl, H., & Connelly, D. Effect of a controlled subseabed

560

release of CO2 on the biogeochemistry of shallow marine sediments, their pore waters, and the

561

overlying water. Int. J. Greenh. G. Con. 38, 80-92 (2014).

562 563

289. McCave, I. N. Evaluation of a laser-diffraction-size analyzer for use with natural

564

sediments. J. Sedim. Res. 56, 561-564 (2013).

565 566

2930. McCormack, C.G. Key impacts of climate engineering on biodiversity and ecosystems,

567

with priorities for future research. Journal of Int. Env. Sci. (in press). (2016).

568 569

Formattato: Inglese (Regno Unito) Formattato: Inglese (Regno Unito)

(17)

301. Meadows, A.S., Ingels, J., Widdicombe, S., Hale, R. & Rundle, S. D. Effects of elevated

570

CO2 and temperature on an intertidal meiobenthic community. Journal of Exp. Mar. Biol. and 571

Ecol. 469, 44-56 (2015). 572

573

313. Mora, C., Wei, C.-L., Rollo, A., Amaro, T., Baco, A.R. Billett, D. Bopp, L., Chen, Q.,

574

Collier, M., Danovaro, R., Gooday, A.J., Grupe, B.M., Halloran, P.R., Ingels, J., Jones,

575

D.O.B., Levin, L.A., Nakano, H., Norling, K., Ramirez-Llodra, E., Rex, M., Ruhl, H.A., Smith,

576

C.R., Sweetman, A.K., Thurber, A.R., Tjiputra, J.F., Usseglio, P., Watling, L., Wu, T. &

577

Yasuhara, M. Biotic and Human Vulnerability to Projected Changes in Ocean Biogeochemistry

578

over the 21st Century. PLoS Biol 11(10), e1001682 (2013).

579 580

323. Paulley, A., Maul, P.R. & Metcalfe, R. Research into impacts and safety in CO2 storage:

581

Scenarios for Potential Impacts from Hypothetical Leakage from Geological Storage Facilities

582

for Carbon Dioxide in RISCS Deliverable D1.3 (ed. Quintessa Limited) 60 pp (2012).

583 584

343. Pearson, T. H. & Rosenberg, R. Macrobenthic succession in relation to organic

585

enrichment and pollution of the marine environment. Oceanogr. Mar. Biol. Ann. Rev. 16, 229–

586

311 (1978).

587 588

345. Pörtner, H-O. Ecosystem effects of ocean acidification in times of ocean warming: a

589

physiologist’s view. Mar. Ecol. Prog. Ser. 373, 203–217 (2008).

590 591

356. Preckler, C.A. Biodiversity and chemical interactions in Antartic benthic communities of

592

Deception Island (South Shetland Islands). PhD thesis. 314 pp (Barcelona university, 2015). 593

594

367. Pusceddu, A., Dell’Anno, A., Fabiano, M. & Danovaro, R. Quantity and bioavailability of

595

sediment organic matter as signatures of benthic trophic status. Mar. Ecol. Progr. Ser. 375,

596

41-52 (2009).

597 598

378. Queiros, A. M.,Taylor P., Cowles A., Reynolds A., Widdicombe S. & Stahl H. Optical

599

assessment of impact and recovery of sedimentary pH profiles in ocean acidification and

600

carbon capture and storage research. Int. J. Greenh. G. Con. 38, 110-120 (2014).

601 602

389. Queiros, A.M. Norling, K., Amaro, T., Nunes, J., Cummings, D., Yakushev, E., Sorensen,

603

K., Harris, C., Woodward, M., Danovaro, R., Rastelli, E., Alve, E., DeVittor, C., Karuza, A.,

604

Cibic, T., Monti, M., Ingrosso, G., Fornasaro, D., Beaubien, S. E., Guilini, K., Vanreusel, A.,

605

Molari, M., Boetius, A., Ramette, A., Wenzhöfer, F., deBeer, D.,Weber, M., Grünke, S.,

606

Bigalke, N. & Widdicombe, S. Potential Impact of CCS Leakage on Marine Communities. (Ed.

607

Plymouth Marine Laboratory) 101 pp. (Plymouth, UK, 2015).

608 609

3940. Rastelli, E., Corinaldesi, C., Dell’Anno, A., Amaro, T., Queiros, A. M., Widdicombe, S.

610

& Danovaro, R. Impact of CO2 leakage from sub-seabed carbon dioxide capture and storage 611

(CCS) reservoirs on benthic virus-prokaryote interactions and functions. Fronteirs in

612

Microbiology 6, 935. doi:10.3389/fmicb.2015.00935 (2015). 613

614

401. Rodolfo-Metalpa, R., Houlbrèque, F., Tambutté, É., Boisson, F., Baggini, C., Patti, F. P.,

615

Jeffree, R., Fine, M., Foggo, A., Gattuso, J.-P & Hall-Spencer, J. M. Coral and mollusc

616

resistance to ocean acidification adversely affected by warming. Nat. Clim. Change 1, 308-312

617

(2011).

618 619

41. Shen, P.-P, Zhou, H. & Gu, J.-D. Patterns of polychaete communities in relation to 620

(18)

environmental perturbations in a subtropical wetland of Hong Kong. Journal of the 621

Marine Biological Association of the United Kingdom 90, 923-932 (2010). 622

623

42. Spicer, J. I, Taylor, A. C. & Hill, A. D. Acid–base status in the sea urchins Psammechinus

624

miliaris and Echinus esculentus (Echinodermata: Echinoidea) during emersion. Mar. Biol. 99,

625

527–534 (1988).

626 627

43. Spicer, J. I. Oxygen and acid–base status of the sea urchin Psammechinus miliaris during

628

environmental hypoxia. Mar. Biol. 124, 71–76 (1995).

629 630

44. Tait, K., Stahl, H., Taylor, P., & Widdicombe, S. Rapid response of the active microbial

631

community to CO2 exposure from a controlled sub-seabed CO2 leak in Ardmucknish Bay

632

(Oban, Scotland). Int. J. Greenh. G. Con. 38, 171-181 (2014).

633 634

45. Tait, K., Beesley, A., Findlay, H.S., McNeill, C.L. & Widdicombe, S. Elevated CO2 635

induces a bloom of microphytobenthos within a shell gravel mesocosm. FEMS

636

10.1093/femsec/fiv092 (2015).

637 638

46. Taylor, P., Stahl, H., Vardy, M. E., Bull, J. M., Akhurst, M., Hauton, C., James, R.H.,

639

Lichtschlag, A., Long, D., Aleynik, D., Toberman, M., Naylor, M., Connelly, D., Smith, D.,

640

Sayer, M.D.J., Widdicombe, S., Wright, I.C. & Blackford, J. A novel sub-seabed CO2 release 641

experiment in forming monitoring and impact assessment for geological carbon storage. Int. J.

642

Greenh. G. Con. 38, 3–17. (2015). 643

644

47. Ueda, N., Tsutsumi, H., Yamada, M., Takeuchi, R. & Kido, K. Recovery of the marine

645

bottom environment of a Japanese bay. Mar. Pollut. Bull. 28, 676–682 (1994).

646 647

48. Widdicombe, S. & Needham, H. R. Impact of CO2 induced seawater acidification on the 648

burrowing activity of Nereis virens and sediment nutrient flux. Mar. Ecol. Prog. Ser. 341,

111-649

122 (2007).

650 651

49. Widdicombe, S. Dashfield, S. L., McNeill, C. L., Needham, H. R., Beesley, A., McEvoy,

652

A., Øxnevad, S., Clarke, K. R. & Berge, J. A. Effects of CO2 induced seawater acidification on 653

in faunal diversity and sediment nutriente fluxes. Mar. Ecol. Prog. Ser. 379, 59–75 (2009).

654 655

50. Widdicombe, S., Blackford, J. C. & Spicer, J. I. Assessing the environmental consequences

656

of CO2 leakage from geological CCS: generating evidence to support environmental risk

657

assessment. Mar. Pollut. Bull.73, 399–401 (2013).

658 659

51. Widdicombe, S., McNeill, C. L., Stahl, H., Taylor, P., Queiros, A. M., Nunes, J., & Tait, K.

660

Impact of sub-seabed CO2 leakage on macrobenthic community structure and diversity. Int. J. 661

Greenh. G. Con. 38, 182-192 (2015). 662

663

52. Widdicombe, S., Spicer, J. I. Predicting the impact of ocean acidification on benthic

664

biodiversity: What can animal physiology tell us? J Exp Mar Biol Ecol 366(1-2),

187-665

197(2008).

666 667

53. Wood, H. L., Spicer, J. I. & Widdicombe, S. Ocean acidification may increase calcification

668

rates—but at a cost. Proc. R. Soc. B. 275, 1767–1773 (2008).

Riferimenti

Documenti correlati

En este sentido, en su encuentro con la literatura, la filosofía se ve forzada a pensar sobre toda una serie de parámetros de lo humano que normalmente pone entre paréntesis, y

The defect of a spherical system is defined as the non-negative integer given by the difference between the number of colors and the number of spherical roots. In particular,

Moreover, the curves of the measured temperature within the air-lube oil separator show similar trends but higher initial plateaus and lower slopes at lower ambient

occurrence (PLT versus DN) and volume of disease (LV versus HV) of mHSPC patients in a hospital-based registry are significantly independent prognostic factors and that a

In the previous sections, we have given the conditions under which a solution of the Stieltjes and Hamburger (reduced) moment problems may be found in the genuine Jaynes’ spirit

Di particolare interesse, a questo proposito, è la distinzione hus- serliana (Husserl, ) tra Körper, il corpo reso oggetto di una conoscenza fattuale, e Leib, il

OEA significantly increased pCREB in the hippocampus of HDC þ/þ mice compared with vehicle treated animals (Bonferroni's post hoc test p &lt;

International Archives of the Photogrammetry, Remote Sensing and Spatial Information Sciences, Volume XXXVIII-5/W16, 2011 ISPRS Trento 2011 Workshop, 2-4 March 2011, Trento, Italy...