• Nenhum resultado encontrado

Inoculum pretreatment differentially affects the active microbial community performing mesophilic and thermophilic dark fermentation of xylose

N/A
N/A
Protected

Academic year: 2023

Share "Inoculum pretreatment differentially affects the active microbial community performing mesophilic and thermophilic dark fermentation of xylose"

Copied!
35
0
0

Texto

(1)

1

Inoculum pretreatment differentially affects the

1

active microbial community performing

2

mesophilic and thermophilic dark fermentation

3

of xylose

4

Paolo Dessì a,*, Estefania Porca b, Luigi Frunzoc, Aino–Maija Lakaniemi a, Gavin Collins b, 5

Giovanni Espositod, Piet N. L. Lens a,e 6

7

aTampere University of Technology, Faculty of Natural Sciences, P.O. Box 541, FI-33101 8

Tampere, Finland 9

bMicrobial Communities Laboratory, School of Natural Sciences, National University of Ireland 10

Galway, University Road, Galway, H91 TK33, Ireland 11

cDepartment of Mathematics and Applications “Renato Caccioppoli”, University of Naples 12

Federico II, Via Cintia, Monte S. Angelo, Naples, Italy 13

dDepartment of Civil and Mechanical Engineering, University of Cassino and Southern Lazio, 14

Via di Biasio, 43, Cassino, FR, Italy 15

eUNESCO–IHE, Institute for Water Education, Westvest 7, 2611AX Delft, The Netherlands 16

17

Manuscript submitted to: Bioresource Technology 18

19

*Corresponding author:

20

Phone: +358 417239696, e-mail: paolo.dessi@tut.fi, mail: Tampere University of Technology, 21

P.O. Box 541, FI-33101 Tampere, Finland 22

(2)

2 Abstract

23

The influence of different inoculum pretreatments (pH and temperature shocks) on 24

mesophilic (37 °C) and thermophilic (55 °C) dark fermentative H2 production from 25

xylose (50 mM) and, for the first time, on the composition of the active microbial 26

community was evaluated. At 37 °C, an acidic shock (pH 3, 24 hours) resulted in the 27

highest yield of 0.8 mol H2 mol-1 xylose. The H2 and butyrate yield correlated with the 28

relative abundance of Clostridiaceae in the mesophilic active microbial community, 29

whereas Lactobacillaceae were the most abundant non-hydrogenic competitors 30

according to RNA-based analysis. At 55 °C, Clostridium and Thermoanaerobacter were 31

linked to H2 production, but only an alkaline shock (pH 10, 24 hours) repressed lactate 32

production, resulting in the highest yield of 1.2 mol H2 mol-1 xylose. This study showed 33

that pretreatments differentially affect the structure and productivity of the active 34

mesophilic and thermophilic microbial community developed from the same inoculum.

35 36

Keywords 37

Biohydrogen; pH shock; Temperature shock; Clostridium; Lactobacillus; MiSeq.

38 39 40 41 42 43 44 45 46

(3)

3 1. Introduction

47

The increasing energy demand results in the depletion of fossil fuel reserves and in the 48

emission of enormous quantities of greenhouse gases to the atmosphere. Due to its high 49

energy content and carbon neutrality, hydrogen (H2) is a promising green alternative to 50

fossil fuels [1]. Among H2 production technologies (for a review, see Nikolaidis and 51

Poullikkas, 2017), biological methods have the advantage of coupling H2 generation 52

with the treatment of organic carbon-containing wastes and wastewaters. Biological H2

53

production technologies include photo and dark fermentation, direct and indirect 54

photolysis, and the water gas-shift reaction [3]. The high H2 production rate, the simple 55

reactor technology (similar to the well-established anaerobic digestion), and the 56

abundance of microorganisms capable to produce H2 from a wide range of substrates 57

may promote the establishment of dark fermentation in industrial applications [4].

58 59

Promising H2 yields have been obtained through mesophilic (30–40 °C), thermophilic 60

(50–60 °C), and hyperthermophilic (> 65 °C) dark fermentation of sugars by pure 61

cultures or synthetic co-cultures of H2 producing microorganisms (for reviews, see Lee 62

et al., 2011 and Elsharnouby et al., 2013). However, a diverse microbial consortium is 63

often required for dark fermentation of more complex substrates, such as waste-derived 64

carbohydrate mixtures. Furthermore, the easier and more economic process control as 65

well as the lack of requirements for sterilisation make mixed cultures more suitable for 66

industrial waste treatment applications than pure cultures [7]. The main drawback of 67

mixed cultures and unsterile conditions is the possible development of microorganisms 68

competing with H2 producers for the substrates or even H2-consuming microorganisms 69

in the dark fermentation reactor, which can drastically reduce the overall H2 yield.

70

(4)

4 Homoacetogenic bacteria and hydrogenotrophic methane producing archaea are the two 71

most common H2 consuming microorganisms, but also propionate producers as well as 72

sulfate and nitrate reducers use H2 during their metabolism [3]. Also lactate producing 73

bacteria are commonly found in dark fermentation systems [8]. They also compete for 74

the substrate, and can even inhibit H2 producers by either causing acidification or 75

excreting bacteriocins [9].

76 77

Inoculum pretreatment technologies have been widely studied on mixed cultures to 78

select H2 producing microorganisms at the expense of H2 consumers. When exposed to 79

harsh conditions, some H2 producers, such as Clostridium, Bacillus and 80

Thermoanaerobacter, produce endospores as a defence mechanism [10,11]. This 81

increases their survival chances compared to non-spore-forming H2 consumers as they 82

are able to regerminate once the environmental conditions become favourable.

83

Temperature and pH shocks are the most widely applied pretreatment methods to enrich 84

the microbial community prior to dark fermentation [12]. High temperature disrupts the 85

cell wall of microorganisms, causing cell lysis and protein denaturation [13]. Low 86

temperature can cause protein gelling, intracellular formation of ice crystals and 87

membrane lipid stiffening [14]. Acidic and alkaline shocks affect the electric charge of 88

the cell membrane, may inactivate key enzymes, and may cause a change of the 89

intracellular pH, which could lead to cell wall disruption [15]. Other pretreatment 90

methods include substrate loading shock, chemical treatment (e.g. using 2- 91

bromoethansulphonate acid or chloroform, which inhibit methanogenic archaea), 92

aeration, electric shocks, ionising irradiation, microwaves, or ultrasonication (for a 93

review, see Wang and Yin 2017).

94

(5)

5 95

Many studies have compared the effect of the various pretreatments, and their 96

combination, on mesophilic [16–23] and thermophilic [24,25] dark fermentative H2

97

production. However, due to the different inoculum, substrate, and operating conditions, 98

many results appear controversial. Most of the studies compared the various 99

pretreatments in terms of H2 production, rather than focusing on the microbial 100

community. In a few cases, DNA-based analyses have been applied to reveal the 101

microbial community composition [16–18,22–24]. Microbial community analyses were 102

mainly conducted using low sensitive techniques, such as denaturing gradient gel 103

electrophoresis (DGGE), allowing detection of only the dominant species of the 104

community. Recent microbiological techniques, such as next-generation high- 105

throughput sequencing, likely allow more detailed information on how the microbial 106

community is affected by various pretreatments, in order to develop new strategies to 107

increase the H2 yields.

108 109

RNA-based approaches provide even more useful information than DNA-based studies 110

[26]. Indeed, they enable to depict the composition of the microbial community that 111

remains active after the various pretreatments, and thus to find out which species are 112

involved in H2 production and possible competitive pathways. Therefore, this study 113

pursues the double aim of (i) finding out how different pretreatments shape the active 114

microbial communities in dark fermentation reactors and (ii) how such microbial 115

communities evolve and produce H2 during mesophilic (37 °C) and thermophilic (55 116

°C) dark fermentation.

117 118

(6)

6 2. Materials and methods

119

2.1 Source and pretreatment of inoculum 120

Activated sludge was selected as the parent inoculum, as it contains a huge variety of 121

microbial species and enabled higher H2 production than digester sludge in a previous 122

study [27]. It was collected from a secondary settler of a municipal wastewater 123

treatment plant (Mutton Island, Galway, Ireland). It was dewatered by filtration through 124

a 0.1 mm diameter mesh. After filtration, total solids, volatile solids, and pH were 28.1 125

(± 2.4) g L-1, 22.7 (± 2.0) g L-1, and 6.9, respectively. The activated sludge was stored at 126

4 °C for approximately one week prior to being used in the experiment.

127 128

Temperature (heat treatment and freezing and thawing) and pH (acidic and alkaline) 129

shocks, widely used to select H2 producing organisms from several inocula [12], were 130

chosen among the possible pretreatment methods. The heat shock was conducted by 131

exposing the inoculum, collected in thin 15 mL tubes, to 90 °C for 15 minutes using a 132

pre-heated water bath (Clifton, UK). Freezing and thawing was done by incubating 133

sludge samples at -20 °C for 24 hours in thin 15 mL tubes and then defrosting them at 134

30 °C in a water bath. The acidic shock was given by adjusting the pH of the sludge to 135

3.0 with HCl, incubating at room temperature (about 20 °C) for 24 hours, and then 136

increasing the pH back to 7.0 with NaOH. The alkaline shock was done by adjusting the 137

pH to 10.0 with NaOH, incubating at room temperature for 24 hours, and then 138

decreasing the pH back to 7.0 with HCl. Both HCl and NaOH were used at a 139

concentration of 1 or 3 M. The sludge was continuously stirred by using a magnetic 140

stirrer while adjusting the pH.

141 142

(7)

7 2.2 Incubation and sampling

143

Glass serum bottles (120 mL volume) were filled with 45 mL of a modified DSMZ 144

medium n.144 (Dessì et al., 2017), and inoculated with 5 mL of pretreated activated 145

sludge. Anaerobic incubations with untreated sludge were also prepared as negative 146

controls. The initial pH was adjusted to 7.0 with 1 M NaOH, and bottles were flushed 147

with N2 for 5 min prior to incubation. Thermostatic conditions (37 or 55 °C) and 148

shaking at 60 rpm were provided by shaker incubators (Thermo Scientific, USA). The 149

batch cultures lasted until cumulative H2 production was not observed consecutively for 150

48 hours in any of the triplicate bottles (96-144 hours in total). All batch cultures were 151

conducted as triplicates.

152 153

Gas samples (5 mL) were collected daily, and stored into 5.9 mL EXETAINER®

154

hydrogen-tight gas sampling tubes (Labco, UK). Liquid samples were also collected 155

daily and stored at -20 °C in 1.5 mL Eppendorf tubes for analysis. Sludge samples were 156

collected before being inoculated into the serum bottles (0 hours samples) and from the 157

bottom of serum bottles at the end of the batch culture at 37 °C or 55 °C. They were 158

then stored at -80 °C in RNA-free 1.5 mL Eppendorf tubes untill microbiological 159

analysis.

160 161

2.3 Microbial community analyses based on 16S cDNA 162

Microbiological samples were pelleted by centrifugation (10,000xg, 10 min). Nucleic 163

acid co-extraction was done using a method modified from Griffiths et al. (2000). DNA 164

inhibition, cDNA synthesis, high-throughput sequencing (Illumina MiSeq) and 165

bioinformatics analysis of partial 16S rRNA gene sequences using primers 515f and 166

(8)

8 805r were performed as reported previously (Dessì et al., manuscript under revision), 167

but using a more recent version of Mothur (1.39.5) and the Silva database (v128). The 168

total number of sequences was 5,494,444 and they were reduced to 3,294,730 169

(2,276,495 unique sequences) after a quality check. The hierarchical clustering 170

dendrogram was done using the ”Pvclust” package of ”R” software [29].

171 172

2.4 Analytical methods 173

Gas produced was quantified by a syringe method (Owen et al., 1979) and its 174

composition was determined by gas chromatography as reported previously (Dessì et al.

175

2017). Cumulative H2 and CO2 production was calculated using a H2 mass balance 176

equation [30], and corrected to standard temperature (0 °C). Total solids and volatile 177

solids were determined by the standard methods (APHA, 1995). pH was measured 178

using a pH meter (WTW inoLab) equipped with a Slimtrode electrode (Hamilton). The 179

composition of the liquid phase (xylose, volatile fatty acids and alcohols) was 180

determined by high-performance liquid chromatograph (Shimadzu) equipped with a 181

Rezex TM ROA-Organic Acid H+ (8%) column (Phenomenex, USA), held at 70 °C, and 182

a refractive index detector (Shimadzu, Japan). The mobile phase and flow rate were 183

0.013 N H2SO4 and 0.6 mL min-1, respectively.

184 185

2.5 Statistical analysis 186

To assess significant differences in the effect of the tested pretreatments on H2 yield, 187

one-way analysis of variance (ANOVA) and the Tukey test [31] at p = 0.05 were 188

conducted using the IBM SPSS Statistics package. The output of the statistical analysis 189

is provided in the supporting material (File S1).

190

(9)

9 191

3. Results and discussion 192

3.1 Effect of the inoculum pretreatments on the active microbial community 193

No dominant microorganisms were detected in the untreated activated sludge, which 194

was characterised by a presence of various families of microorganisms with relative 195

abundance below 20% (Figure 1), mainly belonging to the phylum of Proteobacteria 196

and Bacteroidetes (Figure S1).

197 198

Despite the aerobic origin of the sludge, heat shock, freezing and thawing, and acidic 199

pretreatment favoured the establishment of families of spore forming anaerobic 200

microorganisms such as Clostridiaceae (Figure 1). Some non-spore-forming 201

microorganisms, such as Comamonadaceae sp. [32], were able to resist the heat shock 202

and the freezing and thawing treatment, with a relative abundance of 12-22% and 17- 203

19%, respectively, but their relative abundance was < 5% after the acidic or alkaline 204

shock (Figure 1).

205 206

After the alkaline shock, the active microbial community was very different than the 207

one obtained after the other pretreatments, as shown by the hierarchical clustering 208

dendrogram (Figure 2); it was dominated by Proteobacteria belonging to the families of 209

Moraxellaceae (relative abundance of 48-54%) and Aeromonadaceae (21-28%), and 210

more specifically by microorganisms closely related to Acinetobacter sp. and 211

Aeromonas sp. (Figure 1; Table 1). Although being non-spore forming [33,34], both 212

Acinetobacter and Aeromonas survived the alkaline pretreatment, but their relative 213

abundance dropped to < 1% after applying the other pretreatments. This suggests that 214

(10)

10 microorganisms are differently affected by the different pretreatments, and that also 215

some non-spore forming microorganisms could survive the thermal and pH stress 216

applied. The alkaline shock pretreatment repressed more microbial families than the 217

other pretreatments, as confirmed by the low diversity of the microbial community, 218

according to the diversity indexes (Table 2).

219 220

Figure 1, Figure 2, Table 1, Table 2 221

222

3.2 Effect of pretreatments on mesophilic (37 °C) dark fermentation of xylose 223

3.2.1 Effect of pretreatments on H2 yield at 37 °C 224

Based on H2 yields, the acidic shock was the most effective pretreatment for mesophilic 225

(37 °C) dark fermentation, resulting in a yield of 0.8 mol H2 mol-1 xylose (Figure 3). No 226

statistically significant differences were found between the alkaline and heat shock, 227

which resulted in a H2 yield of 0.5-0.6 mol H2 mol-1 xylose, and in both cases H2

228

production stopped within 48 hours. Freezing and thawing was the least effective 229

pretreatment method, resulting in a yield of only 0.2 mol H2 mol-1 xylose, similar to the 230

yield obtained with the untreated sludge (Figure 3).

231 232

Chang et al. (2011) compared various pretreatment methods (including acidic and 233

alkaline shock, heat shock, aeration, and chemical treatment) on waste activated sludge 234

and reported that an acidic shock (pH 3, 24 hours) was the most effective pretreatment, 235

yielding 1.5 mol H2 mol-1 glucose at 35 °C and at an initial pH of 7. Furthermore, 236

Chang et al. (2011) reported a 20% increase of the H2 yield from glucose after five 237

consecutive mesophilic batch cultures with acidic shock-pretreated activated sludge 238

(11)

11 (without repeating the acidic shock) suggesting that the pretreatment effect can be 239

sustained for extended time periods.

240 241

Figure 3 242

243

3.2.2 Effect of pretreatments on the xylose oxidation pathways at 37 °C 244

In all the mesophilic batch cultures, more than 90% of the xylose was consumed (Figure 245

4). Acetate was always produced, regardless of the pretreatment applied, with a final 246

concentration ranging between 10 and 21 mM. In dark fermentation, the acetate 247

pathway yields the highest amount of H2 (3.3 mol H2 mol-1 xylose). However, the H2

248

partial pressure may cause a shift of the dark fermentation pathway to butyrate 249

production [35], which yields only 1.67 mol H2 mol-1 xylose. This seems to be also the 250

case in this study, as butyrate was found in all the batch cultures (Figure 4) with a 251

concentration proportional to the H2 yield (Figure 5a). The H2 yield obtained from the 252

mesophilic batch cultures with the acidic shock-, alkaline shock-, and heat shock- 253

pretreated activated sludge is, however, too high to be entirely associated to the butyrate 254

pathway, which stoichiometrically yields 2 mol H2 mol-1 butyrate. This suggests that H2

255

was produced through both the acetate and the butyrate pathway. Acetate could be 256

produced also through non-hydrogenic pathways, such as homoacetogenesis, lowering 257

the net H2 production [3].

258 259

At 37 °C, ethanol was produced in all cultures regardless of the pretreatment, with a 260

maximum of 16 mM in the batch cultures with the heat shock-pretreated activated 261

sludge (Figure 4). Solventogenesis can occur as a detoxification process in case of low 262

(12)

12 pH (< 5) and accumulation of undissociated volatile fatty acids [36]. In fact, regardless 263

the pretreatment applied, the pH of the batch medium decreased to below 5.0 after 24- 264

48 hours incubation (Figure 4). The low pH was likely associated to lactate production.

265

In fact, the highest concentration of lactate of 14 and 10 mM was obtained in the batch 266

cultures with the untreated and alkaline shock-pretreated activated sludge, respectively 267

(Figure 4), resulting in the lowest observed pH of 4.0. Chaganti et al. (2012) as well 268

reported lactate production from glucose at 37 °C when using untreated and alkaline 269

shock-pretreated (pH 11, 24 hours) anaerobic sludge as inoculum.

270 271

Figure 4 272

273

3.2.3 Active microbial communities after batch culture at 37 °C 274

Cultivation at 37 °C resulted in a lower diversity of the active microbial community 275

from pretreated than untreated activated sludge, according to the diversity indexes 276

(Table 2). Methane was not detected in the mesophilic batch cultures, including the 277

untreated control, suggesting the absence or inactivity of methanogenic archaea. In fact, 278

the relative abundance of active archaea was < 1% in all the batch cultures. Yin et al.

279

(2014) and Chang et al. (2011) as well did not detect any methane in their mesophilic 280

batch cultures with the pretreated and untreated activated sludge. However, Ren et al.

281

(2008) reported methanogenic activity in mesophilic batch cultures with alkaline shock- 282

pretreated (pH 11, 24 hours) activated sludge.

283 284

Based on the data of this study, mesophilic H2 production correlated reversely with the 285

diversity of the active microbial community (Figure 5b). All mesophilic batch cultures 286

(13)

13 with pretreated activated sludge were dominated by microorganisms belonging to the 287

phylum Firmicutes, present with a relative abundance of 76-98% (Figure S1). Yin and 288

Wang (2016) pretreated anaerobically digested sewage sludge by gamma irradiation, 289

obtaining > 99% of Firmicutes in the microbial community after 36 hours of mesophilic 290

(33 °C) batch incubation with glucose (94.4 mM) as the substrate for dark fermentation.

291

This suggests that Firmicutes dominate the mesophilic dark fermentative microbial 292

communities, regardless of the pretreatment applied to the inoculum.

293 294

The family of Clostridiaceae was found regardless of the pretreatment applied, and 295

even in the batch cultures with the untreated activated sludge (Figure 1). The relative 296

abundance of Clostridiaceae had a positive correlation with both the H2 and butyrate 297

yield (Figure 5a). In fact, the acidic shock, which was the most effective pretreatment in 298

terms of H2 yield (Figure 3), promoted a high relative abundance of Clostridiaceae, 299

which accounted for 71-75% of the active microbial community after 144 hours 300

incubation at 37 °C (Figure 1). Clostridium sp. have been widely reported to dominate 301

mesophilic dark fermentative microbial communities, both in batch and continuous 302

reactors [27,38,39]. Liu et al. (2009) applied similar pretreatments to those applied in 303

this study to an intertidal marine sediment to perform mesophilic (37 °C) dark 304

fermentation of glucose. They reported similar results to this study in terms of 305

pretreatment efficacy and dominant microbial communities. However, their DNA-level 306

analysis, performed by PCR-DGGE, only allowed the detection of the dominant 307

species, whereas the RNA-level approach used in this study describes more accurately 308

the active microbial communities involved in the dark fermentative process.

309 310

(14)

14 Figure 5

311 312

Lactate producers of the Lactobacillaceae family were the main competitors to H2

313

producing microorganisms in all mesophilic batch cultures (Figure 1). Despite the 314

relatively high H2 yield, these microorganisms were found even in the batch cultures 315

with the acidic shock-pretreated activated sludge, with a relative abundance of 8-17%.

316

Kim et al. (2014) sequenced DNA from the microbial community developed upon 317

mesophilic (35 °C) dark fermentation of acidic shock-pretreated (pH 3, 12 hours) food 318

waste, without an external inoculum. They reported a relative abundance of 70% of 319

Clostridium and 20% Lactobacillus, similar to the relative abundances obtained in this 320

study. Furthermore, Kim et al. (2014) reported a higher relative abundance of 321

Clostridium (up to 90 %) and a higher H2 yield after performing an acidic shock at pH 2 322

and 1 for 12 hours, suggesting that the pretreatment conditions used in this study can be 323

further optimised to increase the H2 yield.

324 325

In contrast, an acidic shock (pH 3, 24 hours) of activated sludge has been also reported 326

to inhibit H2 producing microorganisms and favour the establishment of lactic acid 327

producing bacteria (Ren et al., 2008). In the batch cultures with the alkaline shock- 328

pretreated activated sludge, the low final pH of 4.0 (Figure 4) likely inhibited H2

329

producers and favoured lactic acid bacteria, resulting in the highest relative abundance 330

of Lactobacillaceae (53-61% of the relative abundance), which produced lactate at the 331

expenses of H2. Lactic acid bacteria may also inhibit Clostridium by excreting toxins 332

[9].

333 334

(15)

15 Freezing and thawing resulted in a high relative abundance of Proteobacteria (Figure 335

S1), such as Veillonellaceae (36-46%), which were found also in the batch cultures with 336

the untreated sludge (Figure 1). The representative sequence of Veillonellaceae matched 337

the genus Megamonas (Table 1), which has been previously reported to produce 338

propionate and acetate from the fermentation of glucose [41]. Both propionate (8 mM) 339

and acetate (10 mM) were indeed detected in the batch cultures with the freezing and 340

thawing-pretreated activated sludge (Figure 4). Both propionate production and the low 341

relative abundance (12 %) of Clostridium were likely the main cause for the low H2

342

yield obtained in these batch cultures (Figure 1).

343 344

3.3 Effect of pretreatments on thermophilic (55 °C) dark fermentation 345

3.3.1 Effect of pretreatments on H2 yield at 55 °C 346

At 55°C, the alkaline shock resulted in the highest yield of 1.2 mol H2 mol-1 xylose, two 347

times higher than the freezing and thawing pretreatment (Figure 3), which was similar 348

to the H2 yield obtained by the untreated activated sludge (0.5-0.6 mol H2 mol-1 xylose).

349

Both the acidic shock and heat shock resulted in the lowest H2 yield of 0.2-0.3 mol H2

350

mol-1 xylose. Interestingly, H2 was continuously produced for 96 hours in the 351

thermophilic batch cultures with the alkaline shock-pretreated activated sludge, whereas 352

it ceased in the first 24-48 hours in the other batch cultures. To our knowledge, only two 353

studies have been performed to compare the effect of pretreatments on thermophilic 354

dark fermentation [24,25], and none of them reported alkaline shock as the best 355

pretreatment. In the previous studies, granular anaerobic sludge [25] or anaerobic 356

digester sludge from a palm oil mill plant [24] was used as inoculum, whereas activated 357

sludge treating municipal wastewater was used in this study. O-Thong et al. (2009) 358

(16)

16 reported a higher thermophilic (60 °C) H2 yield from sucrose (1.96 mol H2 mol-1

359

hexose) with loading shock-pretreated (2 days incubation with 83.25 gCOD L-1 sucrose) 360

than acidic shock-, alkaline shock-, chemical treatment-, and heat shock-pretreated 361

anaerobic sludge.

362 363

Luo et al. (2010) obtained a higher thermophilic (60 °C) H2 yield from cassava stillage 364

by untreated than pretreated anaerobic granular sludge. This suggests that, depending on 365

the origin of the inoculum, thermophilic conditions can be sufficient to favour H2

366

producing bacteria over competitors. Luo et al. (2010) also claimed that the effect of 367

pretreatments was not significant in the long term, as the H2 yield was similar after 28 368

days of continuous dark fermentation of the cassava stillage regardless the pretreatment 369

applied.

370 371

It should be noted that, even if a low H2 yield is obtained in a batch culture, an efficient 372

H2 producing microbial community may develop in the long term. For example, heat 373

shock-pretreated activated sludge may switch its metabolism to the butyrate pathway 374

resulting in an increased H2 yield: Dessì et al. (2017) reported a 320% H2 yield increase 375

when performing four consecutive thermophilic (55 °C) batch cultures of heat shock- 376

pretreated (90 °C, 15 min) activated sludge. Therefore, to assess the potential of a 377

pretreated inoculum for dark fermentation, additional long-term experiments need to 378

accompany the initial screening test.

379 380

3.3.2 Effect of pretreatments on the xylose oxidation pathways at 55 °C 381

(17)

17 Thermophilic batch cultures with the untreated and alkaline shock- or freezing and 382

thawing-pretreated activated sludge resulted in a higher (> 90%) xylose consumption 383

than batch cultures with the acidic shock- (75%) and heat shock- (66%) pretreated 384

activated sludge (Figure 4). Interestingly, the alkaline shock was the only pretreatment 385

which suppressed lactate production, as most of the xylose was converted to acetate and 386

butyrate up to a final concentration of 6 and 13 mM, respectively (Figure 4).

387 388

In the thermophilic batch cultures with the alkaline shock-pretreated activated sludge, 389

the pH decreased at a lower rate than in the other batch cultures, resulting in the highest 390

final pH (4.4), which could have favoured H2 production. However, the H2 yield of 1.2 391

H2 mol-1 xylose was 30% higher than the maximum expected of 0.9 mol H2 mol-1 392

xylose, according to the acetate and butyrate concentrations produced (Figure 4). H2

393

overproduction has been previously reported in thermophilic dark fermentation [42], 394

and attributed to an unusual H2 production pathway, such as acetate oxidation.

395

However, in the batch cultures with alkaline-shock pretreated activated sludge, the 396

acetate concentration decreased from 9 to 6 mM only in the last 48 hours of incubation 397

(Figure 4), when the cumulative H2 yield was stable (Figure 3), suggesting that acetate 398

oxidation did not occur. Also Zheng et al. (2016) obtained an average of 14 % H2

399

overproduction from glucose by anaerobic sludge, but an isotope feeding assay with 400

13C-labeled acetate excluded the involvement of the acetate oxidation pathway.

401 402

In the thermophilic batch cultures with the freezing and thawing-pretreated inoculum, 403

lactate (8 mM) and ethanol (10 mM) were produced together with acetate (12 mM) and 404

butyrate (8 mM), suggesting that none of the metabolic pathways was prevailing over 405

(18)

18 the other. In the batch cultures with untreated activated sludge, xylose was initially 406

fermented to acetate and butyrate, which both reached a concentration of 5-6 mM after 407

24 hours incubation. Then, a switch from the butyrate to the lactate pathway occurred, 408

and lactate was produced together with acetate in the subsequent 48 hours, resulting in a 409

final concentration of 15 mM lactate and acetate (Figure 4). In the batch cultures with 410

the acidic shock- and heat shock-pretreated sludge, lactate was the main metabolite, 411

with a concentration of 8 and 9 mM, respectively (Figure 4), resulting in the lowest H2

412

yield (Figure 3). Lactate production resulted in a fast pH decrease: the final pH was 4.2- 413

4.3 in the batch cultures with the acidic shock-, heat shock- and freezing and thawing- 414

pretreated activated sludge, and even lower (3.9) in the batch cultures with the untreated 415

activated sludge (Figure 4), in which lactate production was the highest (Figure 4). Such 416

a low pH likely inhibited thermophilic H2 production [44].

417 418

3.3.3 Active microbial communities after batch culture at 55 °C 419

After incubation at 55 °C, regardless of the pretreatment applied, > 98% of the active 420

microbial community was composed of microorganisms of the phylum Firmicutes 421

(Figure S1). In general, the thermophilic active microbial communities were less diverse 422

than the mesophilic ones (Table 2). The active microbial communities from 423

thermophilic batch cultures with the untreated and acidic shock-, alkaline shock-, and 424

freezing and thawing-pretreated activated sludge, were all similar, and dominated by the 425

families of Clostridiaceae and Peptostreptococcaceae (Figure 2).

426 427

A different active microbial community developed in the thermophilic batch culture 428

with the heat shock-pretreated activated sludge (Figure 1), in which a family of 429

(19)

19 Thermoanaerobacterales (Figure 1), closely related to T. thermosaccharolyticum (Table 430

1), was prevailing with a relative abundance up to 94%. Similarly, O-Thong et al.

431

(2009) compared various pretreatments on thermophilic (60 °C) dark fermentation of 432

sucrose in batch experiments, and reported the predominance of T.

433

thermosaccharolyticum from heat shock-pretreated anaerobic sludge, whereas 434

Clostridium became dominant from both acidic and alkaline shock-pretreated anaerobic 435

sludge. The family of Clostridiaceae includes thermophilic species, such as C.

436

thermosaccharolyticum [45] and C. thermopalmarium [46], able to convert sugars to 437

H2, acetate and butyrate. T. thermosaccharolyticum effectively produces H2 from xylose 438

at 55-60 °C and in the pH range 5.5-7.0 (Ren et al., 2008). In the batch cultures with the 439

heat shock-pretreated activated sludge of this study, T. thermosaccharolyticum was thus 440

likely inhibited by the low pH, which dropped to < 4.5 within 48 hours (Figure 4), 441

resulting in a low H2 yield (Figure 3) and poor xylose consumption (Figure 4).

442 443

Bacillales, found in the thermophilic batch cultures with the heat shock- (3-30%) and 444

acidic shock- (3-11%) pretreated activated sludge (Figure 1), might be responsible for 445

lactate production. In fact, the order of Bacillales includes lactic acid producers such as 446

Sporolactobacillus, a spore-forming acidophilic microorganism previously reported in 447

thermophilic dark fermentation of sugars [48].

448 449

4. Conclusions 450

H2 production at 37 °C was dependent on the relative abundance of Clostridiaceae in 451

the microbial community, and an acidic shock favoured their establishment at the 452

expense of competing microorganisms. At 55 °C, an alkaline shock suppressed lactate 453

(20)

20 production, resulting in a high H2 yield, and is thus a suitable pretreatment for starting- 454

up thermophilic H2 producing bioreactors. Inoculum pretreatment appears necessary for 455

mesophilic dark fermentation, whereas less invasive strategies (such as pH control at a 456

pH of 5 or 6) could be considered to select H2 producing microorganisms at the expense 457

of lactic acid bacteria in thermophilic bioreactors.

458 459

Acknowledgements 460

The authors gratefully thank Leena Ojanen (Tampere University of Technology, 461

Finland) for analysing the gas samples, Camilla Thorn (National University of Ireland 462

Galway, Ireland) for helping with the microbial community analysis, and the Mutton 463

island wastewater treatment plant (Galway, Ireland) for providing the activated sludge 464

inoculum.

465 466

Funding 467

This work was supported by the Marie Skłodowska-Curie European Joint Doctorate 468

(EJD) in Advanced Biological Waste-To-Energy Technologies (ABWET) funded from 469

Horizon 2020 under grant agreement no. 643071.

470 471

Declaration of interest 472

The authors declare no conflicts of interest.

473 474

References 475

[1] Palomo-Briones R, Razo-Flores E, Bernet N, Trably E. Dark-fermentative 476

biohydrogen pathways and microbial networks in continuous stirred tank 477

(21)

21 reactors: Novel insights on their control. Appl Energy 2017;198:77–87.

478

[2] Nikolaidis P, Poullikkas A. A comparative overview of hydrogen production 479

processes. Renew Sustain Energy Rev 2017;67:597–611.

480

[3] Bundhoo MAZ, Mohee R. Inhibition of dark fermentative bio-hydrogen 481

production: A review. Int J Hydrogen Energy 2016;41:6713–33.

482

[4] Arimi MM, Knodel J, Kiprop A, Namango SS, Zhang Y, Geißen S-U. Strategies 483

for improvement of biohydrogen production from organic-rich wastewater: A 484

review. Biomass and Bioenergy 2015;75:101–18.

485

[5] Lee DJ, Show KY, Su A. Dark fermentation on biohydrogen production: Pure 486

culture. Bioresour Technol 2011;102:8393–402.

487

[6] Elsharnouby O, Hafez H, Nakhla G, El Naggar MH. A critical literature review 488

on biohydrogen production by pure cultures. Int J Hydrogen Energy 489

2013;38:4945–66.

490

[7] Li C, Fang HHP. Fermentative Hydrogen Production From Wastewater and Solid 491

Wastes by Mixed Cultures. Crit Rev Environ Sci Technol 2007;37:1–39.

492

[8] Etchebehere C, Castelló E, Wenzel J, del Pilar Anzola-Rojas M, Borzacconi L, 493

Buitrón G, et al. Microbial communities from 20 different hydrogen-producing 494

reactors studied by 454 pyrosequencing. Appl Microbiol Biotechnol 495

2016;100:3371–84.

496

[9] Noike T, Takabatake H, Mizuno O, Ohba M. Inhibition of hydrogen fermentation 497

of organic wastes by lactic acid bacteria. Int J Hydrogen Energy 2002;27:1367–

498

71.

499

[10] Collins MD, Lawson PA, Willems A, Cordoba JJ, Fernandez-Garayzabal J, 500

Garcia P, et al. The Phylogeny of the Genus Clostridium: Proposal of Five New 501

(22)

22 Genera and Eleven New Species Combinations. Int J Syst Bacteriol

502

1994;44:812–26.

503

[11] Galperin MY. Genome Diversity of Spore-Forming Firmicutes. Microbiol Spectr 504

2013;1:1–27. doi:10.1128/microbiolspectrum.TBS-0015-2012.Genome.

505

[12] Wang J, Yin Y. Principle and application of different pretreatment methods for 506

enriching hydrogen-producing bacteria from mixed cultures. Int J Hydrogen 507

Energy 2017;42:4804–23.

508

[13] Appels L, Baeyens J, Degrève J, Dewil R. Principles and potential of the 509

anaerobic digestion of waste-activated sludge. Prog Energy Combust Sci 510

2008;34:755–81.

511

[14] Sawicka JE, Robador A, Hubert C, Jørgensen BB, Brüchert V. Effects of freeze–

512

thaw cycles on anaerobic microbial processes in an Arctic intertidal mud flat.

513

ISME J 2010;4:585–94.

514

[15] Rafieenia R, Lavagnolo MC, Pivato A. Pre-treatment technologies for dark 515

fermentative hydrogen production: Current advances and future directions. Waste 516

Manag 2017. doi:10.1016/j.wasman.2017.05.024.

517

[16] Misturini Rossi D, Berne da Costa J, Aquino de Souza E, Ruaro Peralba M do C, 518

Samios D, Záchia Ayub MA. Comparison of different pretreatment methods for 519

hydrogen production using environmental microbial consortia on residual 520

glycerol from biodiesel. Int J Hydrogen Energy 2011;36:4814–9.

521

[17] Guo L, Lu M, Li Q, Zhang J, She Z. A comparison of different pretreatments on 522

hydrogen fermentation from waste sludge by fluorescence excitation-emission 523

matrix with regional integration analysis. Int J Hydrogen Energy 2015;40:197–

524

208.

525

(23)

23 [18] Jeong DY, Cho SK, Shin HS, Jung KW. Application of an electric field for 526

pretreatment of a seeding source for dark fermentative hydrogen production.

527

Bioresour Technol 2013;139:393–6.

528

[19] Chang S, Li J-Z, Liu F. Evaluation of different pretreatment methods for 529

preparing hydrogen-producing seed inocula from waste activated sludge. Renew 530

Energy 2011;36:1517–22.

531

[20] Chaganti SR, Kim D-H, Lalman JA. Dark fermentative hydrogen production by 532

mixed anaerobic cultures: Effect of inoculum treatment methods on hydrogen 533

yield. Renew Energy 2012;48:117–21.

534

[21] Yin Y, Hu J, Wang J. Enriching hydrogen-producing bacteria from digested 535

sludge by different pretreatment methods. Int J Hydrogen Energy 536

2014;39:13550–6.

537

[22] Liu H, Wang G, Zhu D, Pan G. Enrichment of the hydrogen-producing microbial 538

community from marine intertidal sludge by different pretreatment methods. Int J 539

Hydrogen Energy 2009;34:9696–701.

540

[23] Ren N, Guo W-Q, Wang X-J, Xiang W-S, Liu B-F, Wang X-Z, et al. Effects of 541

different pretreatment methods on fermentation types and dominant bacteria for 542

hydrogen production. Int J Hydrogen Energy 2008;33:4318–24.

543

[24] O-Thong S, Prasertsan P, Birkeland N. Evaluation of methods for preparing 544

hydrogen-producing seed inocula under thermophilic condition by process 545

performance and microbial community analysis. Bioresour Technol 546

2009;100:909–18.

547

[25] Luo G, Xie L, Zou Z, Wang W, Zhou Q. Evaluation of pretreatment methods on 548

mixed inoculum for both batch and continuous thermophilic biohydrogen 549

(24)

24 production from cassava stillage. Bioresour Technol 2010;101:959–64.

550

[26] De Vrieze J, Regueiro L, Props R, Vilchez-Vargas R, Jáuregui R, Pieper DH, et 551

al. Presence does not imply activity: DNA and RNA patterns differ in response to 552

salt perturbation in anaerobic digestion. Biotechnol Biofuels 2016;9:244.

553

[27] Dessì P, Lakaniemi A-M, Lens PNL. Biohydrogen production from xylose by 554

fresh and digested activated sludge at 37, 55 and 70 °C. Water Res 555

2017;115:120–9.

556

[28] Griffiths RI, Whiteley AS, O’Donnell AG, Bailey MJ. Rapid Method for 557

Coextraction of DNA and RNA from Natural Environments for Analysis of 558

Ribosomal DNA- and rRNA-Based Microbial Community Composition. Appl 559

Environ Microbiol 2000;66:5488–91.

560

[29] Suzuki R, Shimodaira H. Pvclust: an R package for assessing the uncertainty in 561

hierarchical clustering. Bioinformatics 2006;22:1540–2.

562

[30] Logan BE, Oh SE, Kim IS, Van Ginkel S. Biological Hydrogen Production 563

Measured in Batch Anaerobic Respirometers. Environ Sci Technol 564

2002;36:2530–5.

565

[31] Box GEP, Holohan R, Scott E. Statistics for experimenters: an introduction to 566

design, data analysis, and model building. John Wiley and sons; 1978.

567

[32] Willems A, De Ley J, Gillis M, Kersters K. Comamonadaceae , a new family 568

encompassing the Acidovorans rRNA complex, including Variovorax paradoxus 569

gen. nov., comb. nov., for Alcaligenes paradoxus (Davis 1969). Int J Syst Evol 570

Microbiol 1991;41:445–50.

571

[33] Bouvet PJM, Grimont PAD. Taxonomy of the genus Acinetobacter with the 572

recognition of Acinetobacter baumannii sp. nov., Acinetobacter haemolyticus sp.

573

(25)

25 nov., Acinetobacter johnsonii sp. nov., and Acinetobacter junii sp. nov. and 574

emended descriptio. Int J Syst Bacteriol 1986;36:228–40.

575

[34] Alperi A, Martínez-Murcia AJ, Ko W-C, Monera A, Saavedra MJ, Figueras MJ.

576

Aeromonas taiwanensis sp. nov. and Aeromonas sanarellii sp. nov., clinical 577

species from Taiwan. Int J Syst Evol Microbiol 2010;60:2048–55.

578

[35] Angenent LT, Karim K, Al-dahhan MH, Wrenn BA, Domı R. Production of 579

bioenergy and biochemicals from industrial and agricultural wastewater. Trends 580

Biotechnol 2004;22:477–85.

581

[36] Van Ginkel S, Logan BE. Inhibition of Biohydrogen production by 582

Undissociated Acetic and Butyric Acids. Environ Sci Technol 2005;39:9351–6.

583

[37] Yin Y, Wang J. Changes in microbial community during biohydrogen production 584

using gamma irradiated sludge as inoculum. Bioresour Technol 2016;200:217–

585

22.

586

[38] Si B, Liu Z, Zhang Y, Li J, Xing X-H, Li B, et al. Effect of reaction mode on 587

biohydrogen production and its microbial diversity. Int J Hydrogen Energy 588

2015;40:3191–200.

589

[39] Sivagurunathan P, Kumar G, Park J-H, Park J-H, Park H-D, Yoon J-J, et al.

590

Feasibility of enriched mixed cultures obtained by repeated batch transfer in 591

continuous hydrogen fermentation. Int J Hydrogen Energy 2016;41:4393–403.

592

[40] Kim D-H, Jang S, Yun Y-M, Lee M-K, Moon C, Kang W-S, et al. Effect of acid- 593

pretreatment on hydrogen fermentation of food waste: Microbial community 594

analysis by next generation sequencing. Int J Hydrogen Energy 2014;39:16302–

595 596 9.

[41] Chevrot R, Carlotti A, Sopena V, Marchand P, Rosenfeld E. Megamonas 597

(26)

26 rupellensis sp. nov., an anaerobe isolated from the caecum of a duck. Int J Syst 598

Evol Microbiol 2008;58:2921–4.

599

[42] Kotsopoulos TA, Zeng RJ, Angelidaki I. Biohydrogen production in granular up- 600

flow anaerobic sludge blanket (UASB) reactors with mixed cultures under hyper- 601

thermophilic temperature (70°C). Biotechnol Bioeng 2006;94:296–302.

602

[43] Zheng H, Zeng RJ, O’Sullivan C, Clarke WP. Critical analysis of hydrogen 603

production from mixed culture fermentation under thermophilic condition (60 604

°C). Appl Microbiol Biotechnol 2016;100:5165–76.

605

[44] Zhang T, Liu H, Fang HHP. Biohydrogen production from starch in wastewater 606

under thermophilic condition. J Environ Manage 2003;69:149–56.

607

[45] Islam S, Zhang C, Sui K, Guo C, Liu C. Coproduction of hydrogen and volatile 608

fatty acid via thermophilic fermentation of sweet sorghum stalk from co-culture 609

of Clostridium thermocellum and Clostridium thermosaccharolyticum. Int J 610

Hydrogen Energy 2016;42:830–7.

611

[46] Lawson Anani Soh A, Ralambotiana H, Ollivier B, Prensier G, Tine E, Garcia J- 612

L. Clostridium thermopalmarium sp. nov., a Moderately Thermophilic Butyrate- 613

Producing Bacterium Isolated from Palm Wine in Senegal. Syst Appl Microbiol 614

1991;14:135–9.

615

[47] Ren N, Cao G, Wang A, Lee D, Guo W, Zhu Y. Dark fermentation of xylose and 616

glucose mix using isolated Thermoanaerobacterium thermosaccharolyticum 617

W16. Int J Hydrogen Energy 2008;33:6124–32.

618

[48] Abreu AA, Karakashev D, Angelidaki I, Sousa DZ, Alves MM. Biohydrogen 619

production from arabinose and glucose using extreme thermophilic anaerobic 620

mixed cultures. Biotechnol Biofuels 2012;5:6.

621

(27)

27 Table and Figure captions

622

Table 1. Association of a representative 16S rRNA gene sequence of the most abundant 623

12 microbial families obtained in this study (Figure 1) to those collected in the 624

GenBank.

625 626

Table 2. Diversity of the active microbial community (cDNA) after various 627

pretreatments, and after incubation at 37 or 55 °C, measured by the Shannon, Simpson 628

and Pielou’s J’ index. The number in parenthesis is the standard deviation of the 629

triplicate.

630 631

Figure 1. Relative abundance of the active families resulting from MiSeq sequencing of 632

the partial 16S rRNA on microbiological samples collected before and after batch 633

cultures with the untreated and pretreated sludge at 37 and 55 °C. The results are 634

reported in triplicate. “Other” refers to the sum of genera with relative abundance < 1%.

635 636

Figure 2. Hierarchical clustering of active microbial communities from untreated (U) 637

and pretreated (AS, acidic shock; BS, alkaline shock; HS, heat shock; FT, freezing and 638

thawing) inoculum before (0 h) or after mesophilic (37 °C) or thermophilic (55 °C) dark 639

fermentation of xylose. The values reported are the approximately unbiased (au) p- 640

value, bootstrap probability (bp), and cluster label (edge #). Clusters with au > 95% are 641

indicated in the rectangles. The analysis was conducted on the average relative 642

abundance of triplicates.

643 644

(28)

28 Figure 3. H2 yield from mesophilic (37 °C) and thermophilic (55 °C) dark fermentation 645

of xylose by the untreated or pretreated activated sludge. The results shown are the 646

average of three independent batch cultures. The error bars represent the standard 647

deviation of the triplicates.

648 649

Figure 4. Xylose, metabolite concentration and pH profiles from mesophilic (37 °C) 650

and thermophilic (55 °C) dark fermentation of xylose by the untreated or pretreated 651

activated sludge. The results shown are the average of three independent batch cultures.

652

The error bars represent the standard deviation of the triplicates.

653 654

Figure 5. Correlation between the diversity index (Simpson) and the H2 yield (a), and 655

between the relative abundance of active Clostridiaceae and H2 or butyrate yield (b) 656

after mesophilic (37 °C) dark fermentation of xylose by the untreated (C) and pretreated 657

(AS, acidic shock; BS, alkaline shock; HT, heat treatment; FT, freezing and thawing) 658

inoculum.

659 660

(29)

29 Table 1.

661

Familya Genus and speciesb Accession

number

Matching sequencec

Similarity (%)d

Clostridiaceae Clostridium sp. HF566199 458-749 99

Lactobacillaceae Lactobacillus mucosae MF425117 490-781 99 Peptostreptococcaceae Romboutsia ilealis LN555523 2071187-

2071477

99

Thermoanaerobacterales Family III

Thermoanaerobacterium thermosaccharolyticum

MF405082 414-705 99

Moraxellaceae Acinetobacter sp. KY818302 465-756 99

Comamonadaceae Acidovorax sp. LC279183 425-716 99

Veillonellaceaee Megamonas sp. LT628480 522-812 99

Acidimicrobiales Incertae Sedis

Candidatus Microthrix parvicella

KM052469 91-383 99

Rhodocyclaceaef Dechloromonas KY029047 486-777 99

Myxococcales unclassifiedg

- - - -

Aeromonadaceae Aeromonas MF461171 478-769 100

Chitinophagaceae Terrimonas arctica NR_134213 454-745 99

a The families refer to Figure 4 in the main text; b Closest cultured species in GenBank; c 662

Position (in bp) in which the sequence overlaps the reference sequence; d Percentage of identical 663

nucleotide pairs between the 16S rRNA gene sequence and the closest cultured species in 664

GenBank; e Classified as Selenomonadaceae in the GenBank; e Classified as Azonexaceae in the 665

GenBank; f No match at family, genus and species level.

666 667 668

(30)

30 Table 2.

669

Samplea Pretreatment Shannon

diversity

Simpson diversity

J’ Evenness

0 hours (inoculum)

Untreated 3.10 (± 0.06) 0.92 (± 0.01) 0.73 (± 0.02) Acidic shock 3.08 (± 0.02) 0.92 (± 0.00) 0.72 (± 0.01) Alkaline shock 1.56 (± 0.07) 0.66 (± 0.02) 0.41 (± 0.01) Heat shock 2.86 (± 0.02) 0.88 (± 0.00) 0.66 (± 0.01) Freezing and thawing 3.36 (± 0.03) 0.94 (± 0.00) 0.78 (± 0.01) Batch culture

37 °C

Untreated 2.81 (± 0.07) 0.90 (± 0.01) 0.66 (± 0.02) Acidic shock 1.01 (± 0.10) 0.45 (± 0.03) 0.30 (± 0.03) Alkaline shock 1.21 (± 0.21) 0.56 (± 0.06) 0.35 (± 0.04) Heat shock 1.30 (± 0.07) 0.64 (± 0.02) 0.46 (± 0.04) Freezing and thawing 2.02 (± 0.11) 0.77 (± 0.03) 0.50 (± 0.02) Batch culture

55 °C

Untreated 1.12 (± 0.08) 0.56 (± 0.06) 0.30 (± 0.02) Acidic shock 1.14 (± 0.11) 0.58 (± 0.06) 0.38 (± 0.02) Alkaline shock 1.10 (± 0.20) 0.56 (± 0.08) 0.34 (± 0.02) Heat shock 0.68 (± 0.53) 0.31 (± 0.29) 0.23 (± 0.16) Freezing and thawing 1.01 (0.08) 0.54 (± 0.06) 0.30 (± 0.03)

a All samples were subsampled to the size of the smallest sample (8,654 sequences) 670

671 672 673 674 675 676 677

(31)

31 Figure 1

678

679 680

(32)

32 Figure 2

681

682 683 684 685 686 687 688 689 690 691 692 693

(33)

33 Figure 3

694

695 696 697 698

(34)

34 Figure 4

699

700

(35)

35 Figure 5

701

702

Referências

Documentos relacionados

O caso Merina sugere que Jodensavanne (assim como os demais cemitérios) não é importante apenas do ponto de vista simbólico: é a localização física dos