• Nie Znaleziono Wyników

Operation-driven heterogeneity and overlooked feed-associated populations in global anaerobic digester microbiome

N/A
N/A
Protected

Academic year: 2021

Share "Operation-driven heterogeneity and overlooked feed-associated populations in global anaerobic digester microbiome"

Copied!
29
0
0

Pełen tekst

(1)

Delft University of Technology

Operation-driven heterogeneity and overlooked feed-associated populations in global anaerobic digester microbiome

Mei, Ran; Nobu, Masaru K.; Narihiro, Takashi; Kuroda, Kyohei; Muñoz Sierra, Julian; Wu, Zhuoying; Ye, Lin; Lee, Patrick K.H.; Lee, Po Heng; van Lier, Jules B.

DOI

10.1016/j.watres.2017.07.050 Publication date

2017

Document Version

Accepted author manuscript Published in

Water Research

Citation (APA)

Mei, R., Nobu, M. K., Narihiro, T., Kuroda, K., Muñoz Sierra, J., Wu, Z., Ye, L., Lee, P. K. H., Lee, P. H., van Lier, J. B., McInerney, M. J., Kamagata, Y., & Liu, W. T. (2017). Operation-driven heterogeneity and

overlooked feed-associated populations in global anaerobic digester microbiome. Water Research, 124, 77-84. https://doi.org/10.1016/j.watres.2017.07.050

Important note

To cite this publication, please use the final published version (if applicable). Please check the document version above.

Copyright

Other than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons. Takedown policy

Please contact us and provide details if you believe this document breaches copyrights. We will remove access to the work immediately and investigate your claim.

(2)

Operation-driven Heterogeneity and Overlooked Feed-associated Populations in Global

1

Anaerobic Digester Microbiome

2 3

Ran Meia, Masaru K. Nobua,b, Takashi Narihirob, Kyohei Kurodaa,c, Julian Muñoz Sierrad, 4

Zhuoying Wue, Lin Yef, Patrick K. H. Leeg, Po-Heng Leee, Jules B. van Lierd, Michael J. 5

McInerneyh, Yoichi Kamagatab, Wen-Tso Liua* 6

7

a. Department of Civil and Environmental Engineering, University of Illinois at Urbana-Champaign,

8

Urbana, IL, USA

9

b. Bioproduction Research Institute, National Institute of Advanced Industrial Science and Technology

10

(AIST), Tsukuba, Ibaraki, Japan

11

c. Department of Environmental Systems Engineering, Nagaoka University of Technology, Kami-tomioka,

12

Niigata, Japan

13

d. Section Sanitary Engineering, Department of Water Management, Delft University of Technology,

14

Delft, The Netherlands

15

e. Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University, Hung

16

Hom, Kowloon, Hong Kong

17

f. School of the Environment, Nanjing University, Nanjing, Jiangsu, China

18

g. School of Energy and Environment, City University of Hong Kong, Kowloon, Hong Kong

19

h. Department of Botany and Microbiology, University of Oklahoma, Norman, OK, USA

20 21

*Corresponding author: Wen-Tso Liu

22 Email: wtliu@illinois.edu 23 24 25 26 27 28 29 30 31 32 *Manuscript

Click here to view linked References

© 2017 Manuscript version made available under CC-BY-NC-ND 4.0 license https://creativecommons.org/licenses/by-nc-nd/4.0/

Postprint of Water Research

Volume 124, 1 November 2017, Pages 77-84

(3)

Abstract

33

Anaerobic digester (AD) microbiomes harbor complex, interacting microbial populations to 34

achieve biomass reduction and biogas production, however how they are influenced by operating 35

conditions and feed sludge microorganisms remain unclear. These were addressed by analyzing 36

the microbial communities of 90 full-scale digesters at 51 municipal wastewater treatment plants 37

from five countries. Heterogeneity detected in community structures suggested that no single AD 38

microbiome could be defined. Instead, the AD microbiomes were classified into eight clusters 39

driven by operating conditions (e.g., pretreatment, temperature range, and salinity), whereas 40

geographic location of the digesters did not have significant impacts. Comparing digesters 41

populations with those present in the corresponding feed sludge led to the identification of a 42

hitherto overlooked feed-associated microbial group (i.e., the residue populations). They 43

accounted for up to 21.4% of total sequences in ADs operated at low temperature, presumably 44

due to ineffective digestion, and as low as 0.8% in ADs with pretreatment. Within each cluster, a 45

core microbiome was defined, including methanogens, syntrophic metabolizers, fermenters, and 46

the newly described residue populations. Our work provides insights into the key factors shaping 47

full-scale AD microbiomes in a global scale, and draws attentions to the overlooked residue 48 populations. 49 50 Keywords 51

Anaerobic digester, microbiome, operation, feed sludge 52

53 54

(4)

1. Introduction

55

Wastewater treatment processes, including primary treatment for solids separation and 56

secondary treatment for carbon and nutrients removal, produce substantial amount of waste 57

sewage sludge. For example, the amount of waste sludge generated in European Union is 58

estimated to exceed 13 million dry solid tons in 2020 (Kelessidis and Stasinakis 2012). 59

Anaerobic digestion (AD) has been used worldwide to simultaneously degrade waste sludge and 60

produce methane, and is an promising solution to treat the increasing global growth of organic 61

solid waste (Appels et al. 2011). Meanwhile, the microbial community involved in AD is 62

complex (Narihiro et al. 2015) and a better understanding of the AD ecosystem would optimize 63

existing processes and enhance the engineering application (Vanwonterghem et al. 2014). 64

To identify critical populations responsible for the AD process, multiple researches have 65

tried to define the core AD microbiome. Campanaro et al. (Campanaro et al. 2016) and Treu et al. 66

(Treu et al. 2016) analyzed metagenomic sequences of mesophilic and thermophilic lab-scale 67

digesters treating cattle manure, and concluded that 77 out of 265 genome bins could be 68

considered as the core essential microbial groups in biogas production. Our recent study 69

analyzed the microbial communities of three full-scale digesters in the a wastewater treatment 70

plant and observed a core microbiome that accounted for 59% of the total 16S rRNA gene 71

sequences (Mei et al. 2016a). Studies investigating multiple full-scale plants reported that core 72

populations constituted 36.4% of the total 16S rRNA gene sequences in seven digesters from 73

Seoul, South Korea (Lee et al. 2012), and 28% of the total 16S rRNA gene sequences in seven 74

digesters from France, Germany, and Chile (Riviere et al. 2009). De Vrieze et al. (De Vrieze et al. 75

2015) evaluated the microbial communities of 29 AD installations whose locations were not 76

specified, and reported that Clostridiales and Bacteroidales were part of the core microbiome as 77

(5)

they were shared by each sample with >0.1% abundance. So, if a large number of digesters are 78

sampled and multiple operating parameters are considered, such as temperature, ammonia 79

concentration, and system configuration that are known to influence AD community (De Vrieze 80

et al. 2015, Smith et al. 2017), would it be still possible to define a core AD microbiome? 81

Furthermore, geographical differences in microbiomes have been observed for waste-treating 82

ecosystems like activated sludge (Zhang et al. 2012) and solid waste landfill (Stamps et al. 2016). 83

Would a similar difference be observed with the AD microbiome? 84

A classic categorization of microorganisms in AD consists of fermenting bacteria 85

(fermenters), syntrophic metabolizers (syntrophs), and methanogenic archaea (methanogens) 86

(Schink and Stams 2006). However, it has been realized that AD microbiome embraces a large 87

proportion of prokaryotes with unrecognized ecophysiology (Narihiro 2016). For example, our 88

recent study (Mei et al. 2016a) revealed that 25% of the AD populations in one wastewater 89

treatment plant migrated from the upstream activated sludge process and remained as residue 90

populations in AD. The presence of those non-anaerobic residue populations has not been widely 91

examined to test whether it is a common phenomenon in all digesters under different operating 92

conditions from different geographical locations. Furthermore, the microbial populations in 93

activated sludge can vary considerably due to differences in process configuration and 94

geographical locations (Zhang et al. 2012). Thus, it is not clear whether such variations of 95

microbial populations in the feed sludge impacts the AD microbiome. 96

In this study, we used high-throughput sequencing technologies to characterize 97

microbiomes in digesters around the world by sampling 90 full-scale digesters with diverse 98

operating conditions and feed sludge characteristics from 51 municipal wastewater treatment 99

plants. The impacts of operating conditions and geographical locations on AD microbiome were 100

(6)

examined. Clustering of samples was performed and cluster-specific core populations were 101

identified. Within the AD microbiome, feed-derived populations were investigated and the 102

distribution in different digesters was characterized. 103

104

2. Materials and methods

105

2.1. Sample collection 106

In total, 148 digester sludge samples were collected from 90 full-scale ADs in 51 107

municipal wastewater treatment plants. Feed sludge in 27 plants were collected prior to entering 108

ADs, and feed sludge in the rest plants were not collected due to sampling difficulties. All 109

operation-related information was provided by the plant operators. Besides the volatile solids 110

reduction (VSR) provided by plant operators, we calculated VSR values using the Van Kleeck 111

equation according to the USEPA regulation (Regulations 2003), which were further used in the 112

downstream analyses. Most plants were operated with the conventional primary-secondary 113

(activated sludge) treatment scheme, while three plants were only configured with primary 114

treatment before AD (plant CAII, CALG, and USRA). Seven plants (JPHW, JPMU, JPNA, JPST, 115

JPTB, JPYS, and USDV) used a two-stage anaerobic digestion process with similar sludge 116

retention time (the first digester treating sludge from primary/secondary clarifiers and the second 117

digester treating sludge from the first digester). Seven plants (JPHG, JPNA, JPNG, USST, 118

USUR, NEAV, and USCA) introduced external solid wastes into digesters, such as food waste, 119

green waste, and sludge from other sources. Wastewater to two Hong Kong plants (HKST and 120

HKTP) had approximately 1/4 to 1/5 of seawater of high salinity. Due to its high saline nature 121

with high sulfate content, these two AD digesters dosed ferric chloride (FeCl3) to suppress

(7)

2001, Zhang et al. 2012). Wastewater to another Hong Kong plant (HKYL) had effluent from the 124

tannery industry and contained high concentrations of Zn and Cr (Wong et al. 2001). Digester 125

NEAV1 had both high salinity influent (electrical conductivity about 30-35 mS/cm) and external 126

food waste sludge simultaneously. Digesters from Hong Kong and US (except for USWA and 127

USSF) were sampled at multiple time points with at least one-month interval. These multiple 128

tome points samples were considered as different samples. Fifty milliliters of sludge were 129

collected from the recirculation lines of digesters, transported to laboratory in UIUC on ice 130

(including international samples), and stored at -80°C until DNA extraction. 131

2.2. 16S rRNA gene sequencing 132

Genomic DNA was extracted from 2 mL of well-mixed sludge using the FastDNA SPIN 133

Kit for Soil (MP Biomedicals, Carlsbad, CA, USA), and quantified using a Nanodrop 2000c 134

spectrophotometer. For PCR amplification, 60 ng of genomic DNA was added into a total 135

reaction volume of 25 µL as template. With a dual-indexing approach (Kozich et al. 2013), a 136

universal primer set 515F (5’-GTGCCAGCMGCCGCGGTAA-3’)/909R(5’-137

CCCCGYCAATTCMTTTRAGT-3’) targeting the V4-V5 region of both bacterial and archaeal 138

16S rRNA gene was used for PCR amplification. PCR was performed with the thermal cycling 139

protocol consisting of initial denaturation (94°C, 3 min), 25 cycles of denaturation (94 °C, 30 s), 140

annealing (55 °C, 45 s) and extension (72 °C, 1 min), and a final extension (72 °C, 10 min) (Mei 141

et al. 2016b). The PCR amplicons were purified using the Wizard SV Gel and PCR Clean-Up 142

system (Promega, Fichburg, WI, USA) and quantified by Qubit 2.0 Fluorometer. Library 143

preparation and sequencing on Illumina Miseq Bulk 2 × 300 nt paired-end system was performed 144

at the Roy J. Carver Biotechnology Center at the University of Illinois at Urbana-Champaign, IL, 145

USA. 146

(8)

2.3. Microbial community analyses 147

Paired-end raw sequences were assembled, screened, and trimmed using Mothur 1.33.3 148

(Schloss et al. 2009) with a maximum sequence length of 400 bp and a quality score of 20. The 149

output data were analyzed using QIIME 1.9.1 (Caporaso et al. 2010b) for OTU (operational 150

taxonomic unit, 97% sequence similarity) picking with the de novo strategy, which included 151

OTU grouping by UCLUST (Edgar 2010), alignment by PyNAST (Caporaso et al. 2010a), 152

chimera identification by ChimeraSlayer (Haas et al. 2011), taxonomy assignment by BLAST 153

using reference sequences in the GreenGene 2013 database. After removing singletons (OTUs 154

that only had one sequence in the entire dataset), all samples were rarefied to an even depth of 155

20,957 sequences (determined by the sample with fewest sequences). Shannon index ( 156

is the relative abundance of an individual population) calculation, UniFrac 157

distance matrix calculation, Bray-Curtis distance matrix calculation, principal coordinate 158

analysis (PCoA), and unweighted pair group method with arithmetic mean (UPGMA) with 100 159

iterations were all performed using QIIME. Relative abundance was calculated from OTU table. 160

Phylogenetic trees was constructed using the methods of neighbor joining and parsimony 161

provided in ARB program (Ludwig et al. 2004). 162

Statistical differences of principal components between samples from different locations 163

were evaluated using Mann Whitney U test with Bonferroni correction with R (Ihaka and 164

Gentleman 1996). A p-value < 0.01 was considered as statistical significance. Correlations 165

between microbial groups, alpha diversity, and VSR were determined using the Spearman's Rank 166

Order Correlation test with R. Evaluation of normality of the data using Shapiro Wilk Normality 167

test, and preparation of box plot and histogram were also performed using R. Distance-based 168

linear model (DistLM) and analysis of similarity (ANOSIM) were performed with Primer 6 169

(9)

(Clarke 1993). Raw Illumina sequences obtained in this study have been deposited in 170

DDBJ/NCBI/EMBL-EBI under the accession number DRA005150. 171

172

3. Results 173

3.1. Operation-driven heterogeneity of AD microbiome 174

In total, over 7 million quality-filtered, non-chimeric sequences were obtained from 148 175

AD samples in 51 municipal wastewater treatment plants (Fig. S1, Table S1 in the 176

Supplementary material). After removing singletons and subsampling to an even depth (20,957 177

sequences per sample, determined by the sample with fewest sequences), each AD sample on 178

average contained 1,844 OTUs with a high standard deviation of 595 OTUs. The Shannon index 179

that characterized both richness and evenness of a community showed large variations (Fig. S2), 180

with the highest value being 2.5 times higher than the lowest value (9.12 vs. 3.68). Dissimilarity 181

between AD communities was also reflected in the large variations in the relative abundance of 182

major phyla (Fig. S3). For example, the abundance of Bacteroidetes varied from 5% to 71% in 183

different samples, and the abundance of Thermotogae varied from 0 to 56%. 184

Principal coordinate analysis (PCoA) performed on beta-diversity (weighted UniFrac 185

distance) showed that there were different types of AD communities (Fig. S4). However, the 186

variance could not be explained by geographical locations, as only North America samples 187

significantly differed from Hong Kong samples in PC1 and from Japan samples in PC2. In 188

addition, only small portions of the variance could be explained by single environmental 189

parameters such as temperature (9.63%), pH (3.22%), and sludge retention time (SRT) (1.63%) 190

(Table S2A). 191

(10)

To identify shaping factors of the heterogeneous AD communities, the dissimilarity based 192

on weighted Unifrac was further analyzed using unweighted pair group method with arithmetic 193

mean (UPGMA), a clustering method that could fully reveal the variance in beta diversity. Eight 194

clusters were observed (Fig. 1). Cluster A contained six samples from saline digesters in two 195

Hong Kong plants due to flushing toilet with sea water. Cluster B contained two samples from 196

digesters (one from the US and one from the Netherlands) that received feed sludge after 197

pretreatment with thermal hydrolysis. Cluster C contained three samples from the digester 198

treating wastewater partially from the tannery industry in a Hong Kong plant. Cluster D 199

contained 14 samples from thermophilic digesters (>50°C) located in Japan, US, Canada, and the 200

Netherlands. Cluster E contained seven samples from two Japanese plants and one USA plant 201

that operated digesters at temperatures < 30°C for at least three months. Cluster F contained six 202

samples from one non-saline Hong Kong plant (HKSW, digester temperature at 36.0°C) and one 203

USA plant (USNO, digester temperature at 30.3°C), but the operating conditions that determined 204

high community similarity of these two plants are still not clear. Cluster G contained 16 samples 205

from seven Japanese plants, with slightly high operation temperatures between 38 and 42°C, 206

except for plant JPSS at 36.5°C. The largest cluster (H) contained 91 samples of from 16 USA 207

plants, six Japanese plants, two Canadian plants, and four Netherlandish plants, which operated 208

digesters mainly under mesophilic conditions. Samples from plant USLA and NEAV were not 209

assigned to any cluster due to lack of clear association with operating conditions. Within each 210

cluster, samples that originated from the same plant generally clustered together, even though 211

they might be collected from different reactors or on different dates. The clustering of the AD 212

microbiomes into eight clusters was confirmed by ANOSIM, which gave global R-values close 213

to 1, showing that the between-cluster distances were significantly larger than the within-cluster 214

(11)

distances (Table S2B). In contrast, the clustering solely based on the geographical location of the 215

samples generated much smaller global R-values (less than 0.6) (Table S2C). A UPGMA-based 216

clustering on Bray-Curtis distance matrix produced very similar results (Fig. S5C), where only 217

two samples diverged from cluster G and three samples diverged from cluster H compared to the 218

results based on weighted UniFrac. 219

3.2. Characterization of feed-derived residue populations 220

Our previous study revealed that, in a single wastewater treatment plant, AD microbial 221

communities could contain exogenous populations (i.e., residue populations) that migrated from 222

the feed sludge, resisted to digestion, and not actively involved in anaerobic metabolism (Mei et 223

al. 2016a). In the present study with a much broader sampling scale, we identified such residue 224

populations by comparing the upstream feed sludge and the corresponding AD. To be stringent, 225

we first defined an OTU as being more abundant in feed sludge in a plant only when its feed/AD 226

abundance ratio was over 2, and, conversely, an OTU as being more abundant in AD when the 227

feed/AD abundance ratio was below 0.5. Further, we defined OTUs as residue populations if 228

they were frequently more abundant in feed sludge (minimum five plants) and rarely more 229

abundant in AD (maximum five plants) (Fig. 2). Using these criteria, 1,464 OTUs were 230

identified as residue populations. In agreement, only 172 of them were associated with known 231

obligate anaerobic taxa based on family-level phylogeny (TableS3) obtained from literature 232

(Rosenberg et al. 2014, Vos et al. 2011). In total, 704 residue OTUs were associated with 233

Proteobacteria and 298 OTUs with Bacteroidetes, accounting for 20.8% and 13.4% of 234

sequences in feed sludge, respectively (Fig. 3A). Abundances of these OTUs in the AD 235

community decreased drastically to 3.6% and 1.6%, respectively. Other phyla including 236

Firmicutes, Planctomycetes, and Chloroflexi also contained residues populations but were 237

(12)

presented by a small number of OTUs (<100) and low relative abundance. Detailed phylogenetic 238

analysis of the top 50 abundant residue OTUs indicated that 21 of them were associated with 239

Proteobacteria (excluding Deltaproteobacteria) and 18 OTUs with Bacteroidetes (excluding 240

Bacteroidales) (Fig. S6), which were mostly known as aerobes or facultative anaerobes and were 241

consistent with our previous study (Mei et al. 2016a). On the other hand, known anaerobic 242

populations in AD were not assigned as residue in our analysis, although they were detected in 243

the feed sludge. These populations included for example methanogens (e.g., Methanobacteriales, 244

Methanomicrobiales, and Methanosarcinales), fermenters (e.g., Anaerolineales), and syntrophs 245

(e.g., Syntrophobacterales). Their abundance increased after entering AD, and no residue OTU 246

was related to these taxa (Fig. S7). 247

We further observed that the presence of residue populations was a universal 248

phenomenon in all the digesters sampled (Fig. 3B). The lowest relative abundance of residue 249

populations in a sample was 0.02% in USSF1 that received feed sludge after pretreatment, and 250

the majority (117 out of 148 AD samples) were less than 10%. High residue populations were 251

less common, with 26 samples between 10%-20%, and five samples between 20-30%. The 252

highest abundance was observed with JPYS1 (27.3%) that was operated below 20°C. 253

Furthermore, we observed a clear positive correlation (rho=0.846, p<0.01) between residue 254

populations and alpha diversity (Shannon index) of the AD community (Fig. 3C), indicating the 255

migration of residue populations increased both species richness and evenness of the AD 256

microbial community. In contrast, varying abundance of endogenous populations, such as 257

methanogens or syntrophs, did not correlate with Shannon index of the community (small rho 258

values, Fig. S8). Also a higher residue population abundance was observed to coincide with a 259

(13)

lower digestion efficiency (volatile solids reduction) (Fig. S9), but the correlation was weak as 260

indicated by a low coefficient (rho=-0.361, p<0.01). 261

The presence of residue populations was also influenced by operating conditions (Fig. 262

3D). The highest residue population abundance was 21.4%, observed in cluster E (low operating 263

temperature), followed by 13.9% in cluster F. Correspondingly, clusters E and F had the highest 264

alpha diversity. The abundance of residue populations in cluster H, which represented most of 265

the digesters studied, was 6.0%. In comparison, clusters B (pretreatment), D (thermophilic), and 266

G (>40°C) contained residue population at relative abundances of 0.8%, 3.3%, and 1.6%, 267

respectively. In addition, residue populations could be more abundant than syntrophs (1.2-7.1%) 268

and methanogens (0.3-2.6%), such as in cluster E and F (Fig. S10). We also tested whether 269

residue populations affected beta-diversity by removing residue OTUs from each community. 270

Based on weighted UniFrac distance, clusters A to G remained intact. Seven samples that were 271

originally in cluster H were separated from the cluster(Fig. S5A and B). Based on Bray-Curtis 272

distance, samples in cluster E were split (Fig. S5C and D). 273

3.3. Identification of cluster-specific core populations 274

The heterogeneity revealed by the occupancy distribution of OTUs among all 148 AD 275

samples precluded the ability to define a universal core AD microbiome (Fig. S11A). No OTU 276

was present in 147 or 148 samples. Only 14 OTUs were detected in more than 136 samples, and 277

they only accounted for 4.8% of total sequences. In contrast, within each cluster, OTUs shared 278

by all the samples accounted for a large portion of the total sequences (>50% in each cluster, Fig. 279

S11B), indicating that samples in the same cluster tended to have highly similar microbiomes. 280

Thus, we defined cluster-specific core populations (Fig. 4) by including OTUs that were both 281

prevalent and abundant (top 15 abundant bacterial and top three abundant archaeal OTUs that 282

(14)

were detected in all samples of that cluster). The phylogeny of core OTUs was confirmed by 283

building phylogenetic trees (Fig. S12). 284

In the core communities, OTUs related to known syntrophs were limited to Smithella and 285

Syntrophomonas (Fig. S12A), known to syntrophically oxidize propionate. Smithella related 286

OTUs were observed in clusters B, E, F, G, and H, whereas Syntrophomonas related OTUs were 287

observed in clusters with high salinity (cluster A), industrial influent (cluster C), and high 288

operating temperature (cluster D). For the methanogenic core populations, there was a similar 289

trend that an OTU related to Candidatus Methanofastidiosa (hydrogenotrophic methanogen) and 290

an OTU related to Methanosaeta (aceticlastic methanogen) were consistently observed in 291

clusters B, E, F, G, and H. The high-temperature cluster D contained two unique core OTUs 292

related to Methanothermobacter and Methanoculleus. The low-temperature cluster E contained 293

one unique core OTU related to Methanoregula. Cluster C with industrial influent contained two 294

core OTUs related to Methanosarcina, absent in the core communities of other clusters. Cluster 295

A with high salinity contained an OTU related to Methanolinea but at low abundance (<0.05%) 296

compared with other hydrogenotrophic methanogens. The core community of cluster A also 297

contained an OTU related to Methanosaeta, but likely a different species from the one shared by 298

other clusters based on phylogenetic analysis (Fig. S12B). 299

With regards to residue populations, the core communities of cluster B (plants with 300

pretreatment) and cluster G (plants operated at ~ 40°C) did not contain any OTU identified as 301

residue population. For the core communities of other clusters, Proteobacteria were the major 302

taxa, and the core residue populations were generally related to Zoogloea, Decholomonas, 303

Azospira, and Acidovorax (Fig. S12C). Cluster F contained residue populations mainly related to 304

Sphingobacteria in Bacteroidetes, likely because the feed sludge of cluster F had highest 305

(15)

abundance of Bacteroidetes and lowest abundance Proteobacteria in comparison to other 306

clusters (Fig. S13). 307

The remaining core populations were classified as fermenters. Bacteroidetes, as the most 308

diverse, abundant, and ubiquitous phylum, contained 30 core OTUs, all related to the order 309

Bacteroidales (Fig. S12D). All clusters contained multiple Bacteroidetes-related core OTUs, 310

except for cluster B (plants operated at thermophilic conditions) with only one Bacteroidetes-311

related core OTU. Other major phyla were Firmicutes, Candidatus Cloacimonetes (WWE1), 312

Spirochaetes, and Thermotogae. The majority of fermenters were only assigned to a taxonomic 313

level at order or phylum, as a few known closely isolates were available including Mesotoga, 314

Defluviitoga, Anaerobaculum, Sedimentibacter, and Coprothermobacter. Last, we observed core 315

populations related to phyla without cultivated representatives, including Candidatus 316

Aminicenantes (OP8), Candidatus Fermentibacteria (Hyd24-12), Candidatus Atribacteria (OP9) 317

and Candidatus Marinimicrobia (SAR406). 318

319

4. Discussion

320

Determining the core microbiome for an ecosystem is an effective approach to delineate 321

how microbes drive biochemical processes (Consortium 2012, Gilbert et al. 2014, Sunagawa et 322

al. 2015). This study demonstrated heterogeneity in AD microbial communities, and rejected the 323

possibility to define a universal core microbiome for all digesters that differed in operational 324

conditions. This was contradictory to studies using a small number of digesters (Campanaro et al. 325

2016, Lee et al. 2012, Mei et al. 2016a, Riviere et al. 2009), but consistent with the previous 326

report that when a relatively large number of digesters were sampled, different types of 327

(16)

was linked to diversity in operating conditions, which further led to the discovery of cluster-329

specific core microbiomes. For example, in digesters operated at high temperature (those in 330

cluster D), core OTUs related to thermophiles, including Methanoculleus (Cheng et al. 2008), 331

Methanothermobacter (Cheng et al. 2011) Defluviitoga (Hania et al. 2012), Coprothermobacter 332

(Etchebehere et al. 1998), and Anaerobaculum (Rees et al. 1997) were uniquely detected. OTUs 333

related to zinc-tolerant Sedimentibacter (Burkhardt et al. 2011) were detected in digesters 334

(cluster C) receiving tannery industry wastewater that had high Zn concentration. OTUs related 335

to sulfur-utilizing Mesotoga (Nesbø et al. 2012) were detected in digesters (cluster D) receiving 336

sea water. These sulfur-utilizing microorganisms could compete for hydrogen and suppress 337

hydrogenotrophic methanogens in cluster D. An OTU related to Methanoregula that could grow 338

at 10°C was detected in digesters in cluster E operated under 30°C (Yashiro et al. 2011). It could 339

be expected that if more digesters with more diverse operating conditions are included, the 340

heterogeneity and the clustering complexity will keep increasing as niche diversity increases. 341

Although there was no shared population among all the eight clusters, some populations 342

were frequently observed in clusters B, E, F, G, and H. These populations included OTUs related 343

to the novel archaeal clade Candidatus Methanofastidiosa that is predicted to perform 344

hydrogenotrophic methanogenesis through methylated thiol reduction (Nobu et al. 2016), and 345

Smithella that syntrophically oxidize propionate (Liu et al. 1999). Possibly methylated thiol 346

compounds (e.g., methanethiol and dimethylsulfide) and propionate are critical intermediates 347

prevalent in most ADs. We also observed abundant and diverse OTUs affiliated with the phyla 348

Bacteroidetes and Candidatus Cloacimonetes, whose ecological functions in AD are still 349

difficult to discern. For example, isolates of Bacteroidetes from anaerobic reactors could be 350

saccharolytic (Su et al. 2014, Sun et al. 2016) or proteolytic (Abe et al. 2012, Chen and Dong 351

(17)

2005), but the vast majority of the members in this phylum remain uncultivated and, thus, their 352

metabolism is unknown (Wu et al. 2011). Candidatus Cloacimonetes-related populations have 353

been proposed to perform amino acids fermentation (Pelletier et al. 2008), syntrophic propionate 354

oxidation (Nobu et al. 2015), or extracellular cellulose hydrolysis (Limam et al. 2014). Given 355

that the core OTUs in this phylum were associated with distinct uncultivated phylogenetic clades 356

(e.g., W22, SHA-116, BHB21, and W5), one can only speculate about their metabolisms until 357

more genomics information becomes available or until representatives of these clades are 358

cultured. 359

Previous studies detected core AD populations related to known aerobic and facultative 360

microorganisms including Thauera, Brachymonas, and Rhodobacter (Nelson et al. 2011, Riviere 361

et al. 2009) that were reported as predominant microorganisms in activated sludge (Zhang et al. 362

2012). Their appearance as core populations in AD is likely due to incomplete digestion, in 363

contrast to other core populations such as methanogens, syntrophs, and fermenters. It is known 364

that activated sludge processes sometimes contain anaerobic zones supporting the growth of 365

anaerobic microorganisms in (Kämpfer et al. 1996). Based on the change in abundance before 366

and after entering AD, our analysis could effectively distinguish microorganisms in feed sludge 367

as residue populations (i.e., decreasing abundance) from those contributing to digestion (i.e., 368

increase in abundance) in AD. Thus, the residue populations we define here were unlikely to 369

involve in the essential functions in AD, i.e., waste degradation and biogas production. Further 370

investigations are necessary to elucidate the exact survival mechanisms of the residue 371

populations in AD. For example, some of them could survive on accumulated carbon reserve like 372

polyhydroxyalkanoates (Liu et al. 2001) or carry out anaerobic metabolism with different 373

(18)

electron acceptors (e.g., nitrate reduction by Zoogloea (Shao et al. 2009) and chlorate reduction 374

by Dechloromonas (Achenbach et al. 2001)). 375

On the basis of our previous study of AD in a single plant (Mei et al. 2016a), we showed 376

here that the presence of residue populations was a common phenomenon among all the sampled 377

digesters. Residue populations could account for at least 6% of total sequences obtained from 378

digesters under normal conditions (i.e., cluster H community) and were more abundant than 379

methanogens and syntrophs. Higher abundance of residue populations (i.e., 21.4%) was observed 380

with cluster E likely due to low operating temperature at <30°C. In addition, pretreatment such 381

as thermal hydrolysis could successfully reduce residue populations in AD (i.e., 0.8% abundance 382

in cluster B community). 383

We observed that the presence of residue populations only contributed to the increase of 384

alpha diversity of the AD microbiome. By removing residue populations from each community, 385

we observed almost no change on the beta-diversity, and the topology of the clustering remained 386

almost the same based on either weighted UniFrac (only seven samples split from cluster H) or 387

Bray-Curtis distance (only cluster E split). This is likely due to the fact that most residue 388

populations were affiliated with Proteobacteria and Bacteroidetes, which only represented a 389

small fraction of the vast phylogenetic diversity of AD microbiome. Moreover, the abundances 390

of residue populations were generally less than 10% in most digesters, thus their impacts on the 391

beta diversity calculation were marginal. Only when the abundance of residue populations was 392

high (i.e., in cluster E), a major impact was observed. Finally, a very weak correlation was 393

observed between the abundance of residue populations and overall digestion efficiency, likely 394

because the presence of residue populations could only indicate inefficient cell lysis, the first 395

step of AD process (Amani et al. 2010). The digestion efficiency of full-scale systems is 396

(19)

collectively influenced by other factors including compositions of the feed sludge, mixing 397

condition of the reactor, and monitoring approaches. 398

399

5. Conclusion

400

The analyses of microbial communities of 90 full-scale anaerobic digesters around the 401

world lead to the following conclusions: 402

The differences of microbial community structures were determined by the operating 403

conditions of digesters, whereas geographical location of the digesters did not have a 404

significant impact. 405

Residue populations associated with undigested feed sludge were commonly observed in 406

all the AD samples, with the highest abundance observed in low-temperature digesters 407

and lowest abundance in digesters with pretreatment. 408

There was no population shared by all the sampled digesters due to the operation-driven 409

heterogeneity. The cluster-specific core microbiome contained methanogens, syntrophs, 410

fermenters, and residue populations. 411

412

Acknowledgements

413

We are sincerely grateful to the numerous operation staff in the 51 wastewater treatment 414

plants for their generous help on providing samples and sharing information on operating 415

parameters and efficiencies. This research is part of an ongoing project supported by Joint 416

Genomic Institute. 417

(20)

References

419

Abe, K., Ueki, A., Ohtaki, Y., Kaku, N., Watanabe, K. and Ueki, K. (2012) Anaerocella delicata gen. nov.,

420

sp. nov., a strictly anaerobic bacterium in the phylum Bacteroidetes isolated from a methanogenic reactor

421

of cattle farms. The Journal of General and Applied Microbiology 58(6), 405-412.

422

Amani, T., Nosrati, M. and Sreekrishnan, T. (2010) Anaerobic digestion from the viewpoint of

423

microbiological, chemical, and operational aspects-a review. Environmental Reviews 18(NA), 255-278.

424

Appels, L., Lauwers, J., Degrève, J., Helsen, L., Lievens, B., Willems, K., Van Impe, J. and Dewil, R.

425

(2011) Anaerobic digestion in global bio-energy production: potential and research challenges.

426

Renewable and Sustainable Energy Reviews 15(9), 4295-4301.

427

Burkhardt, E.-M., Bischoff, S., Akob, D.M., Büchel, G. and Küsel, K. (2011) Heavy metal tolerance of Fe

428

(III)-reducing microbial communities in contaminated creek bank soils. Applied and environmental

429

microbiology 77(9), 3132-3136.

430

Campanaro, S., Treu, L., Kougias, P.G., Francisci, D., Valle, G. and Angelidaki, I. (2016) Metagenomic

431

analysis and functional characterization of the biogas microbiome using high throughput shotgun

432

sequencing and a novel binning strategy. Biotechnology for biofuels 9(1), 26.

433

Caporaso, J.G., Bittinger, K., Bushman, F.D., DeSantis, T.Z., Andersen, G.L. and Knight, R. (2010a)

434

PyNAST: a flexible tool for aligning sequences to a template alignment. Bioinformatics 26(2), 266-267.

435

Caporaso, J.G., Kuczynski, J., Stombaugh, J., Bittinger, K., Bushman, F.D., Costello, E.K., Fierer, N.,

436

Pena, A.G., Goodrich, J.K. and Gordon, J.I. (2010b) QIIME allows analysis of high-throughput

437

community sequencing data. Nature Methods 7(5), 335-336.

438

Chen, S. and Dong, X. (2005) Proteiniphilum acetatigenes gen. nov., sp. nov., from a UASB reactor

439

treating brewery wastewater. International journal of systematic and evolutionary microbiology 55(6),

440

2257-2261.

441

Cheng, L., Dai, L., Li, X., Zhang, H. and Lu, Y. (2011) Isolation and characterization of

442

Methanothermobacter crinale sp. nov., a novel hydrogenotrophic methanogen from the Shengli oil field.

443

Applied and environmental microbiology 77(15), 5212-5219.

444

Cheng, L., Qiu, T.L., Li, X., Wang, W.D., Deng, Y., Yin, X.B. and Zhang, H. (2008) Isolation and

445

characterization of Methanoculleus receptaculi sp. nov. from Shengli oil field, China. FEMS

446

microbiology letters 285(1), 65-71.

447

Clarke, K.R. (1993) Non‐parametric multivariate analyses of changes in community structure. Australian

448

journal of ecology 18(1), 117-143.

449

Consortium, H.M.P. (2012) Structure, function and diversity of the healthy human microbiome. Nature

450

486(7402), 207-214.

451

De Vrieze, J., Saunders, A.M., He, Y., Fang, J., Nielsen, P.H., Verstraete, W. and Boon, N. (2015)

452

Ammonia and temperature determine potential clustering in the anaerobic digestion microbiome. Water

453

Research 75, 312-323.

(21)

Edgar, R.C. (2010) Search and clustering orders of magnitude faster than BLAST. Bioinformatics 26(19),

455

2460-2461.

456

Etchebehere, C., Pavan, M., Zorzopulos, J., Soubes, M. and Muxi, L. (1998) Coprothermobacter platensis

457

sp. nov., a new anaerobic proteolytic thermophilic bacterium isolated from an anaerobic mesophilic

458

sludge. International journal of systematic and evolutionary microbiology 48(4), 1297-1304.

459

Gilbert, J.A., Jansson, J.K. and Knight, R. (2014) The Earth Microbiome project: successes and

460

aspirations. BMC biology 12(1), 1.

461

Haas, B.J., Gevers, D., Earl, A.M., Feldgarden, M., Ward, D.V., Giannoukos, G., Ciulla, D., Tabbaa, D.,

462

Highlander, S.K. and Sodergren, E. (2011) Chimeric 16S rRNA sequence formation and detection in

463

Sanger and 454-pyrosequenced PCR amplicons. Genome research 21(3), 494-504.

464

Hania, W.B., Godbane, R., Postec, A., Hamdi, M., Ollivier, B. and Fardeau, M.-L. (2012) Defluviitoga

465

tunisiensis gen. nov., sp. nov., a thermophilic bacterium isolated from a mesothermic and anaerobic whey

466

digester. International journal of systematic and evolutionary microbiology 62(6), 1377-1382.

467

Ihaka, R. and Gentleman, R. (1996) R: a language for data analysis and graphics. Journal of

468

computational and graphical statistics 5(3), 299-314.

469

Kämpfer, P., Erhart, R., Beimfohr, C., Böhringer, J., Wagner, M. and Amann, R. (1996) Characterization

470

of bacterial communities from activated sludge: culture-dependent numerical identification versus in situ

471

identification using group-and genus-specific rRNA-targeted oligonucleotide probes. Microbial ecology

472

32(2), 101-121.

473

Kelessidis, A. and Stasinakis, A.S. (2012) Comparative study of the methods used for treatment and final

474

disposal of sewage sludge in European countries. Waste Management 32(6), 1186-1195.

475

Koenig, A. and Bari, Q. (2001) Vane shear strength of dewatered sludge from Hong Kong. Water science

476

and technology 44(2-3), 389-397.

477

Kozich, J.J., Westcott, S.L., Baxter, N.T., Highlander, S.K. and Schloss, P.D. (2013) Development of a

478

dual-index sequencing strategy and curation pipeline for analyzing amplicon sequence data on the MiSeq

479

Illumina sequencing platform. Applied and environmental microbiology 79(17), 5112-5120.

480

Lee, S.-H., Kang, H.-J., Lee, Y.H., Lee, T.J., Han, K., Choi, Y. and Park, H.-D. (2012) Monitoring

481

bacterial community structure and variability in time scale in full-scale anaerobic digesters. Journal of

482

Environmental Monitoring 14(7), 1893-1905.

483

Limam, R.D., Chouari, R., Mazéas, L., Wu, T.D., Li, T., Grossin‐Debattista, J., Guerquin‐Kern, J.L.,

484

Saidi, M., Landoulsi, A. and Sghir, A. (2014) Members of the uncultured bacterial candidate division

485

WWE1 are implicated in anaerobic digestion of cellulose. MicrobiologyOpen 3(2), 157-167.

486

Liu, W.T., Nielsen, A.T., Wu, J.H., Tsai, C.S., Matsuo, Y. and Molin, S. (2001) In situ identification of

487

polyphosphate‐and polyhydroxyalkanoate‐accumulating traits for microbial populations in a biological

488

phosphorus removal process. Environmental Microbiology 3(2), 110-122.

(22)

Liu, Y., Balkwill, D.L., Aldrich, H.C., Drake, G.R. and Boone, D.R. (1999) Characterization of the

490

anaerobic propionate-degrading syntrophs Smithella propionica gen. nov., sp. nov. and Syntrophobacter

491

wolinii. International journal of systematic bacteriology 49(2), 545-556.

492

Ludwig, W., Strunk, O., Westram, R., Richter, L., Meier, H., Buchner, A., Lai, T., Steppi, S., Jobb, G.

493

and Förster, W. (2004) ARB: a software environment for sequence data. Nucleic acids research 32(4),

494

1363-1371.

495

Mei, R., Narihiro, T., Nobu, M.K., Kuroda, K. and Liu, W.-T. (2016a) Evaluating digestion efficiency in

496

full-scale anaerobic digesters by identifying active microbial populations through the lens of microbial

497

activity. Scientific reports 6, 34090.

498

Mei, R., Narihiro, T., Nobu, M.K. and Liu, W.-T. (2016b) Effects of heat shocks on microbial community

499

structure and microbial activity of a methanogenic enrichment degrading benzoate. Letters in Applied

500

Microbiology 63(5), 356-362.

501

Narihiro, T. (2016) Microbes in the Water Infrastructure: Underpinning Our Society. Microbes and

502

Environments 31(2), 89-92.

503

Narihiro, T., Nobu, M.K., Kim, N.K., Kamagata, Y. and Liu, W.T. (2015) The nexus of

syntrophy-504

associated microbiota in anaerobic digestion revealed by long-term enrichment and community survey.

505

Environmental Microbiology 17(5), 1707-1720.

506

Nelson, M.C., Morrison, M. and Yu, Z. (2011) A meta-analysis of the microbial diversity observed in

507

anaerobic digesters. Bioresource Technology 102(4), 3730-3739.

508

Nesbø, C.L., Bradnan, D.M., Adebusuyi, A., Dlutek, M., Petrus, A.K., Foght, J., Doolittle, W.F. and Noll,

509

K.M. (2012) Mesotoga prima gen. nov., sp. nov., the first described mesophilic species of the

510

Thermotogales. Extremophiles 16(3), 387-393.

511

Nobu, M.K., Narihiro, T., Kuroda, K., Mei, R. and Liu, W.-T. (2016) Chasing the elusive Euryarchaeota

512

class WSA2: genomes reveal a uniquely fastidious methyl-reducing methanogen. ISME J 10(10),

2478-513

2487.

514

Nobu, M.K., Narihiro, T., Rinke, C., Kamagata, Y., Tringe, S.G., Woyke, T. and Liu, W.T. (2015)

515

Microbial dark matter ecogenomics reveals complex synergistic networks in a methanogenic bioreactor.

516

ISME J 9(8), 1710-1722.

517

Pelletier, E., Kreimeyer, A., Bocs, S., Rouy, Z., Gyapay, G., Chouari, R., Rivière, D., Ganesan, A.,

518

Daegelen, P. and Sghir, A. (2008) “Candidatus Cloacamonas acidaminovorans”: genome sequence

519

reconstruction provides a first glimpse of a new bacterial division. Journal of Bacteriology 190(7),

2572-520

2579.

521

Rees, G.N., Patel, B.K., Grassia, G.S. and Sheehy, A.J. (1997) Anaerobaculum thermoterrenum gen. nov.,

522

sp. nov., a novel, thermophilic bacterium which ferments citrate. International journal of systematic and

523

evolutionary microbiology 47(1), 150-154.

524

Regulations, E. (2003) Technology: Control of Pathogens and Vector Attraction in Sewage Sludge.

525

USEPA, Offıce of Research and Development.

(23)

Riviere, D., Desvignes, V., Pelletier, E., Chaussonnerie, S., Guermazi, S., Weissenbach, J., Li, T.,

527

Camacho, P. and Sghir, A. (2009) Towards the definition of a core of microorganisms involved in

528

anaerobic digestion of sludge. The ISME Journal 3(6), 700-714.

529

Rosenberg, E., DeLong, E.F., Lory, S., Stackebrandt, E. and Thompson, F. (2014) The Prokaryotes: Other

530

Major Lineages of Bacteria and The Archaea, Springer Berlin Heidelberg.

531

Schink, B. and Stams, A.J. (2006) Syntrophism among prokaryotes, Springer.

532

Schloss, P.D., Westcott, S.L., Ryabin, T., Hall, J.R., Hartmann, M., Hollister, E.B., Lesniewski, R.A.,

533

Oakley, B.B., Parks, D.H. and Robinson, C.J. (2009) Introducing mothur: open-source,

platform-534

independent, community-supported software for describing and comparing microbial communities.

535

Applied and environmental microbiology 75(23), 7537-7541.

536

Smith, A.L., Shimada, T. and Raskin, L. (2017) A comparative evaluation of community structure in

full-537

scale digesters indicates that two-phase digesters exhibit greater microbial diversity than single-phase

538

digesters. Environmental Science: Water Research & Technology.

539

Stamps, B.W., Lyles, C.N., Suflita, J.M., Masoner, J.R., Cozzarelli, I.M., Kolpin, D.W. and Stevenson,

540

B.S. (2016) Municipal solid waste landfills harbor distinct microbiomes. Frontiers in microbiology 7.

541

Su, X.-L., Tian, Q., Zhang, J., Yuan, X.-Z., Shi, X.-S., Guo, R.-B. and Qiu, Y.-L. (2014)

542

Acetobacteroides hydrogenigenes gen. nov., sp. nov., an anaerobic hydrogen-producing bacterium in the

543

family Rikenellaceae isolated from a reed swamp. International journal of systematic and evolutionary

544

microbiology 64(9), 2986-2991.

545

Sun, L., Toyonaga, M., Ohashi, A., Tourlousse, D.M., Matsuura, N., Meng, X.-Y., Tamaki, H., Hanada,

546

S., Cruz, R. and Yamaguchi, T. (2016) Lentimicrobium saccharophilum gen. nov., sp. nov., a strictly

547

anaerobic bacterium representing a new family in the phylum Bacteroidetes, and proposal of

548

Lentimicrobiaceae fam. nov. International journal of systematic and evolutionary microbiology.

549

Sunagawa, S., Coelho, L.P., Chaffron, S., Kultima, J.R., Labadie, K., Salazar, G., Djahanschiri, B., Zeller,

550

G., Mende, D.R. and Alberti, A. (2015) Structure and function of the global ocean microbiome. Science

551

348(6237), 1261359.

552

Treu, L., Kougias, P.G., Campanaro, S., Bassani, I. and Angelidaki, I. (2016) Deeper insight into the

553

structure of the anaerobic digestion microbial community; the biogas microbiome database is expanded

554

with 157 new genomes. Bioresource Technology 216, 260-266.

555

Vanwonterghem, I., Jensen, P.D., Ho, D.P., Batstone, D.J. and Tyson, G.W. (2014) Linking microbial

556

community structure, interactions and function in anaerobic digesters using new molecular techniques.

557

Current opinion in biotechnology 27, 55-64.

558

Vos, P., Garrity, G., Jones, D., Krieg, N.R., Ludwig, W., Rainey, F.A., Schleifer, K.-H. and Whitman, W.

559

(2011) Bergey's Manual of Systematic Bacteriology: Volume 3: The Firmicutes, Springer Science &

560

Business Media.

561

Wong, J., Li, K., Fang, M. and Su, D. (2001) Toxicity evaluation of sewage sludges in Hong Kong.

562

Environment International 27(5), 373-380.

(24)

Wu, G.D., Chen, J., Hoffmann, C., Bittinger, K., Chen, Y.-Y., Keilbaugh, S.A., Bewtra, M., Knights, D.,

564

Walters, W.A. and Knight, R. (2011) Linking long-term dietary patterns with gut microbial enterotypes.

565

Science 334(6052), 105-108.

566

Yashiro, Y., Sakai, S., Ehara, M., Miyazaki, M., Yamaguchi, T. and Imachi, H. (2011) Methanoregula

567

formicica sp. nov., a methane-producing archaeon isolated from methanogenic sludge. International

568

journal of systematic and evolutionary microbiology 61(1), 53-59.

569

Zhang, T., Shao, M.-F. and Ye, L. (2012) 454 Pyrosequencing reveals bacterial diversity of activated

570

sludge from 14 sewage treatment plants. The ISME Journal 6(6), 1137-1147.

571 572

(25)

0.1 US PS4 US OA 4

USMO4 (Feed slud

ge) US D E 3 NE V L 2 J PS T1 U SD V3 USAR1 USPE2 USS T9 USPN 4 U S D E1 JPKH 3 J PS T2 CA LG1 USPE1 USAR2 USMO3 US ST 10 US DV 1 U SR A 1 USS T4 US NP6 NE DR 1 USS T8 J PHG 2 J P TR 2 CALI1 JPT B1 US ST 16 JP MU 2 US DV 4 JPY S2 USUR1 USCA1 USD I1 HKTP3 HKTP2 US TU4 JPN G1 USMO2 USWL2 USPN 3 NEHA1 US OA 5 US D E4 US PS2 US OA6 US R A2 J P TR 1 US U R3 USBL1 U SD E 5 US PS 5 NEVE 1 US PS1 USPE3 USMO1 J PS B1 USS T12 CAA I1 USS T3 USS F2 JP MU 1 N EB R1 U SS F3 JPT R 4 JP HW 2 US LA 1 JPAS3 JPKH 1 H KS W 1 J PT R 3 USST1 JP NA 1 US PN 5 USU R2 NEAV1 US DE 2 US T U 2 JP NA 2 J PHG3 USU R4 J PS S 1 US T U 3 JPKH2 JPAS1 J P TO 2 USPN 1 US TU 1 CAI I! JPM R3 USU R6 U SWA2 HK S W 2 JPY S 1 US OA 1 USSF1 US OA3 HKST3 USS T5 US WA1 US OA 2 USPN 2 USCA2 HKYL 3 USU R5 USBL3 JPM R2 USS T7 J PHG 1 JPM R1 USS T11 HKYL2 US D V2 JPAS2 JP IK1 J PT O 3 U S N O 3 JP NA 4 HKST2 USS T14 JP NA 3 US RA 3

USMO6 (Feed sludge)

US DV 5 USDI2 H KS W 3 J P TO 1 USWL1 US LA2 US PS 6 HKYL 1 US N O 1 JP HW 1 NE NW 1 USP S3 USS T2 NE VL1 HKTP1 USBL2 HKST1 US ST13 JPT B2 J PK H 4 USS T15 USS T6 US CA3 J PS B2 US N O 2

USMO8 (Feed sludge)

Figure 1. Clustering of digester microbial communities. UPGMA dendrogram was built

USS T US U R* US DV US D E U ST U US VL NE NW US MU JP NA USU R* US PS USP N CAI I JPKH * CALI NEHA USAR USPE JPAS USBL USMO USWL USCA JP IK US DI JP TB NEB R JP H G JP SB JP ST JP TR JP T O JP S S HK S W US N O USR A JP YS JP HW US LA US OA NE DR CA LG CA AI JPN G JPM R HKYL NEVE HKST HKTP NEAV USSF * US SF* JP KH * US W A* USSF * USW A* USUR*

A

B

C

<30°C

E

F

G

~40°C h ig h sa lin ity p re trea tm en t ind u st rial in flue nt >5 0°C

D

H

>90% >70% >50% Figure 1

(26)

Residue

populations

0 5 10 15 20 25 27 0 5 10 15 20 25 27

Number of plants (feed

abundance

/AD

abundance

≥ 2)

N

u

m

b

e

r o

f p

la

n

ts (

fe

e

d

ab u n d a n c e

/A

D

ab u n d a n c e

0

.5

)

Figure 2. Identification of OTUs related to residue populations. For

each OTU, the x value represents the number of plants where the OTU

has more than double abundance in feed sludge than in AD. The y

value represents the number of plants where the OTU has more than

double abundance in AD than in feed sludge. The size of each bubble

represents the number of OTUs (in log scale) sharing the same x-y

coordinate. The color scale represents at a given coordinate the ratio

of the number of OTUs associated with aerobic/facultative anaerobic

taxa over the number of OTUs associated with obligate anaerobic

10 100 >1000 ≥5 2 1 0.5 ≤0.2

Number of OTUs sharing same coordinate

OTUs associated with aerobic/facultative anaerobic taxa vs. OTUs associated with obligate anaerobic taxa

(27)

0 150 300 450 600 750 900 0 5 10 15 20 25 30 O T U n u m b e r R e la ti v e a b u n d a n c e ( % ) Proteo bacte ria Bacte roidete s Planc tomyce tes Firmicut es Verruc omicr obia Chlor oflexi Ca. P arcub acter ia Actin obacter ia GN02 Acido bacte ria Relative abundance in feed sludge community Relative abundance in AD community OTU number (A) rho=0.846, p<0.01 40 50 20 20 30 30 0 0 10 10 Relative abundance (%) N o . o f s a m p le s (B) 0 5 10 15 20 25 30 3 4 5 6 7 8 9 10 R e la ti v e a b u n d a n c e ( % ) Shannon index (C)

Figure 3. Distribution of OTUs identified as residue populations. Panel (A) shows the top ten phyla that contain high numbers of residue OTUs. Dots represent numbers of OTUs of this phylum (primary y axis). Solid bars represent abundance of residue populations of this phylum in feed sludge community and open bars represent abundance of residue populations of this phylum in AD community (secondary y axis). Panel (B)

0 1.5 3 4.5 6 7.5 9 0 5 10 15 20 25 30 A B C D E F G H R e la ti v e a b u n d a n c e ( % ) S h a n n o n i n d ex

Relative abundance of residue populations Shannon index

(D)

Cluster

(28)

D (>50°C) C (industrial influent) A (high salinity) B (pre-treatment) E (<30°C) F G (~40°C) H Taxomony OTUID#

Relative abundance in each cluster 251952 454177 91897 456136 320695 75801 444616 170163 119112 274978 433935 429809 40322 411727 356423 215050 424987 200474 203311 147804 413531 170164 286773 234708 260866 143565 453562 249182 59307 35110 44552 202151 276527 15143 202864 431477 137302 82547 441867 288191 28139 132409 176925 68756 387815 406612 323003 106994 135922 439264 221084 482600 211125 214687 36264 129869 209485 169256 148920 441596 186335 361090 82535 150104 223517 33917 46600 263980 20743 46097 359293 416445 308382 354771 59281 129871 97104 224485 436831 15147 37214 59278 306593 416952 275687 386455 139545 284625 18768 423765 257239 347826 336001 86947 347296 309783 348740 355271 27060 Methanothermobacter Ca. Methanofastidiosa Methanoculleus Methanolinea Methanoregula Methanospirillum Methanosaeta Methanosarcina Syntrophomonas Smithella Flavobacterium Chitinophagaceae Acidovorax Dechloromonas Azospira Zoogloea Gammaproteobacteria Ca. Latescibacteria Bacteroidales Bacteroidetes Bacteroidetes Proteobacteria Proteobacteria Firmicutes Sedimentibacter Coprothermobacter Firmicutes Sphirochaetes Ca. Cloacimonetes 287086 Defluviitoga Mesotoga Thermotogae

Others Ca. Aminicenantes

Caldiserica Choloroflexi Ca. Fermentibacteria Ca. Atribacteria Ca. Marinimicrobia Synergistetes Verrucomicrobia Methanogen Syntroph Residue Fermenter Figure 4

(29)

Electronic Supplementary Material (for online publication only)

Cytaty

Powiązane dokumenty

Impulsem do powstania niniejszego artykułu były moje doświadczenia zdobyte podczas stażów naukowych w Klinikach Chirurgii Piersi Europejskiego Instytutu Onkologii w

Renal dysfunction as a predictor of stroke and systemic embolism in patients with nonvalvular atrial fibrillation: validation of the R(2)CHADS(2) index in the ROCKET AF

The present study investigated the relationship between hypertension and insulin resistance in non-diabetic participants who were referred to Qazvin Metabolic Diseases

The article provides multidimensional comparative analyses of dependent variables (32 European countries, years) and explanatory variables (population and human deaths) in terms of

Dopiero 8 kwietnia 1720 r. w aktach urzędującego ówcześnie burmistrza Ga­ briela a Bomeln znajdujemy następującą wiadomość dotyczącą kasy. Otóż dnia tego

Type 1 &gt; 2 of pes digital form ula shows highest frequencies in all populations studied, including both tribes and castes. Sex and bilateral differences seem

There is also another version o f t-Bonferroni test, designed for non-correlated variables.. Differ­ ences appear only when identifying homogenous groups. Therefore

that evolution by natural selection will take place in any population (a group of causally connected objects) in which there are phenotypic variations, heritability, and differences