Insight into the Aggregation Capacity of Anammox Consortia during

Mar 5, 2018 - Department of Environmental Engineering, Peking University , Beijing 100871 , China. ‡ Key Laboratory of Water and Sediment Sciences, ...
0 downloads 9 Views 2MB Size
Subscriber access provided by UNIV OF DURHAM

Remediation and Control Technologies

Insight into aggregation capacity of anammox consortia during reactor start-up Yunpeng Zhao, Ying Feng, Jianqi Li, Yongzhao Guo, Liming Chen, and Sitong Liu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b06553 • Publication Date (Web): 05 Mar 2018 Downloaded from http://pubs.acs.org on March 8, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 43

Environmental Science & Technology

1

Insight into aggregation capacity of anammox consortia

2

during reactor start-up

3

Yunpeng Zhao 1, 2, Ying Feng 1, 2, Jianqi Li 2, 3, Yongzhao Guo 2, 3, Liming Chen 1, 2,

4

Sitong Liu 1, 2 *

5

1

Department of Environmental Engineering, Peking University, Beijing 100871, China

6

2

Key Laboratory of Water and Sediment Sciences, Ministry of Education of China, Beijing

7

100871, China

8

3

9

518055, China

School of Environment and Energy, Peking University Shenzhen Graduate School, Shenzhen

10

*Corresponding author: Sitong Liu

11

Address: College of Environmental Science and Engineering, Peking University, Yiheyuan Road,

12

No.5, Haidian District, Beijing 100871, China.

13

E-mail: [email protected]

14

Tel/Fax: 0086-10-62754290.

15

1

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 43

16

Abstract

17

Anammox aggregates have been extensively observed in high-efficiency

18

nitrogen-removal reactors, yet the variation and inherent cause of its aggregation

19

capacity related to reactor operation are still unknown. Here, we used microbial

20

detection, metabolomics, extended DLVO theory and multivariate statistical analysis

21

to address this issue. The aggregation capacity of anammox consortia varied

22

periodically during reactor operation, which was determined by the hydrophobic force

23

and the ratio of extracellular protein (PN) to extracellular polysaccharides (PS).

24

Fundamentally, it related to the variation of polysaccharides degradation bacteria

25

abundance and the discrepancy of consortia metabolism. Specifically, the

26

distinguishable up-regulation of the amino acids Phe, Leu, Ala, Thr, Gly, Glu, and Val

27

potentially contributed to the high biosynthesis of extracellular PN. Together with the

28

reduced extracellular PS production that was regulated via the

29

UDP-N-acetyl-D-glucosamine and UDP-N-acetyl-D-galactosamine pathways, the

30

elevated extracellular PN/PS resulted in the obviously increased extracellular

31

hydrophobicity and aggregation capacity. Additionally, the overtly enriched

32

phosphatidylethanolamine biosynthesis pathway was also vital to increasing

33

extracellular hydrophobicity to accelerate aggregation. Understanding aggregation

34

capacity variation is useful for advancing anammox aggregation for its application in

35

wastewater treatment.

36

2

ACS Paragon Plus Environment

Page 3 of 43

Environmental Science & Technology

37

Introduction

38

The anaerobic ammonium-oxidizing (anammox) process, which converts ammonium

39

and nitrite to dinitrogen gas,1 has been applied as a promising technology to treat

40

ammonium wastewater with its inherent advantages of high nitrogen removal, no

41

requirement for oxygen, an additional carbon source and low sludge yield.2 The slow

42

growth rate of functional bacteria with two weeks of doubling time3 makes the high

43

sludge retention a primary requirement for anammox reactor operation. Researchers

44

have applied lots of different strategies to improve sludge retention, such as adding

45

functional carriers, using membrane filters in MBR4 and rapid granulation by

46

signaling molecules, among others. Actually, no matter the consortia granulation or

47

biofilm formation, the intrinsic aggregation capacity of anammox consortia is very

48

important, which varies in different conditions. To investigate this issue, it is

49

important not only to advance biofilm or aggregation formation but also to clarify the

50

potential metabolism conversion in the bacteria along with the different phases of

51

reactor operation.

52

The extracellular polymeric substance (EPS) matrix plays a core role in

53

determining the bacterial aggregation capacity. There were three different concepts

54

existing in the EPS compositions determining the microbial aggregation: (i) The “PS”

55

mechanism: the gel-forming alginate-like PS could form a cross-network that favors

56

sludge aggregation or is a trigger for sludge granulation;5–7 (ii) The “PN” mechanism:

57

the electronegativity and hydrophobicity of PN could promote and stabilize the

58

aggregate structure;8,9 and (iii) The “PN/PS” mechanism: increasing the PN/PS ratio 3

ACS Paragon Plus Environment

Environmental Science & Technology

59

could induce an increase in sludge hydrophobicity and bioflocculation ability.10,11

60

Previous reports have demonstrated that the EPS of anammox sludge has a strong

61

tendency to form biofilm or adhere to inert surfaces,12,13 as it has a larger amount of

62

hydrophobic proteins and more loose secondary structure compared to activated

63

sludge, nitrifying sludge and denitrifying sludge.8,14 The EPS is closely related to the

64

microbial survival ability and aggregation morphology. There are some factors that

65

influence EPS production, such as growth phase, substrate type, and external

66

conditions.15 In fact, the EPS levels and chemical composition vitally depend on the

67

bacteria metabolism. Therefore, the metabolism profile of the bacteria significantly

68

determines its aggregation capacity.

Page 4 of 43

69

EPS consists of different types of biopolymers, such as PN, PS, and phospholipids.

70

They are all organic macromolecules that are biosynthesized by polymerization of the

71

specific building blocks.5 For example, 20 amino acids, such as Ala, Arg, Asn, and

72

others, are reported for bacterial protein synthesis.16 Bacterial polysaccharides are

73

found to be converted from monosaccharides that could be biosynthesized from the

74

Embden–Meyerhoff–Parnas gluconeogenesis (EMP) pathway or amino sugars as

75

precursors.17 The major composition of extracellular phospholipids includes

76

phosphatidylethanolamine (PtE) and phosphatidylcholine (PtC).16,17 Recently,

77

researchers found many generation genes and pathways for bacterial EPS. For

78

example, the biosynthesis of tetratricopeptide repeat (TPR) protein is coded by the

79

asnB gene for activated sludge bacteria.18 Polysaccharide biosynthesis is encoded by

80

the cps3E-F region in some types of lactic acid bacteria.19 Phosphoethanolamine (PE) 4

ACS Paragon Plus Environment

Page 5 of 43

Environmental Science & Technology

81

and phosphocholine (PC) can be formed by a bifunctional

82

cardiolipin/phosphoethanolamine synthase and glycerolphosphocholine (GPC)

83

pathway, respectively.20 Although the hydrophobicity of anammox consortia EPS for

84

its aggregation has been described previously, how bacterial metabolism regulates

85

EPS production and further influences the bacterial aggregation capacity remain to be

86

clarified.

87

Therefore, this study focused on understanding the variation in bacterial

88

aggregation capacity during anammox reactor operation and interpreting the inherent

89

cause using recently developed metabolome biotechnology. Bacterial aggregation

90

capacity, EPS components, and community composition combined with the extended

91

DLVO theory for the bacteria samples harvested across 210 days of reactor operation

92

were used to go insight into consortia aggregation capacity variation during anammox

93

reactor start-up.

94

Materials and methods

95

Reactor start-up and operation. A membrane bioreactor (MBR) with a 5.0-L

96

working volume was operated for 210 d. The polyvinylidene fluoride (PVDF)

97

membrane was placed in the center of the reactor. Then, anammox consortia that had

98

been stored at 4 °C for 60 days was inoculated with the initial VSS concentration of

99

about 0.06 g /L.21 The influent ammonium and nitrite concentration was gradually

100

increased from 20 mg–N/L that was favorable for inoculum adaption to 300 mg–N/L.

101

The details of the medium composition are presented in Supporting Information (SI)

5

ACS Paragon Plus Environment

Environmental Science & Technology

102

Page 6 of 43

Text S1.

103

The reactor was operated in continuous mode at a constant flow rate (2.3 mL/min,

104

HRT=36 h) and was maintained in an anaerobic condition during the reactor operation.

105

According to the microbial community structure, the start-up period could be divided

106

into two stages: phase I (days 1–90) and phase II (days 91–210). The reactor pH was

107

kept in the range of 7.5 to 8.0, and the temperature was maintained at 37 ± 1 °C.

108

The influent and effluent concentrations of ammonium, nitrite and nitrate were

109

determined every two days using an APHA standard engineering method. 22

110

Suspended granular samples were obtained from the reactor every 30 days in two

111

phases (on day 30 60, 90, 120, 150, 180 and 210) and samples on day 165 and 195

112

were also harvested to get more information about the variation of aggregation

113

capacity in the late stage of the reactor operation. Inoculum and these samples were

114

labeled in order as S1–S10, respectively. These samples were used to determine

115

biomass concentration, aggregation capacity, diameter, EPS composition and structure,

116

contact angle, zeta potential and bacterial community structure. Meanwhile, the

117

representative samples were also applied to detect metabolites: (i) the samples

118

obtained every 60 days [on day 60 (S3), 120 (S5) and 180 (S8), respectively]; (ii) the

119

samples obtained on the final day of the reactor operation [on day 210 (S10)].

120

Aggregation capacity and diameter size assay. Aggregation capacity was

121

determined using attachment-crystal violet staining.23 The consortia samples with

122

synthetic medium were added to a 12-well plate and then incubated on a shaker. After

6

ACS Paragon Plus Environment

Page 7 of 43

Environmental Science & Technology

123

that, the microbial suspension that could not be immobilized onto the plate wall was

124

removed. Anammox consortia that were immobilized onto the plate wall were stained

125

with crystal violet before elution by ethanol. Finally, the aggregation capacity was

126

expressed in terms of the absorbance of ethanol solution. Higher absorbance indicated

127

higher aggregation capacity.23 Additionally, the turbidity assays were used to support

128

the results. Detailed methods are presented in SI Texts S2 and S3. The diameter of

129

samples was determined using the Nano ZS instrument (Malvern. Inc., USA), based

130

on the assumption that the aggregates were spherical.24 Five parallel samples were

131

determined for each test.

132

EPS Extraction and Determination. EPS was extracted using the cation exchange

133

resin (CER) method proposed by Hou et al.8 The extracellular polysaccharides and

134

proteins were determined using the Anthrone method25 and the BCA assay26

135

respectively, and was normalized by the biomass concentration. The biomass

136

concentrations were determined in terms of volatile suspended solids (VSS)

137

concentration22 using the vigorously mixed consortia by magnetic stirring and N2

138

gas.27 The PN or PS percentage was defined as the ratio of PN or PS to the sum of PN

139

and PS concentrations. The extraction EPS solution was also lyophilized for the FTIR

140

and XPS analyses.8 Pearson correlation analysis and Kruskal–Wallis tests were carried

141

out using SPSS v20.0.

142

Extended DLVO theory. To determine the interaction energy between sludge cells

143

influencing anammox aggregation, the extended DLVO analysis was performed.28-30

144

The total interaction energy (WT) is the summation of the Van der Waal energies (WA), 7

ACS Paragon Plus Environment

Environmental Science & Technology

145

electric double layer (WR), and acid–base interaction (WAB), or WT = WA + WR +

146

WAB. Calculation of the functional forces of WR, WA, and WAB was performed

147

according to the surface thermodynamic calculation approach based on the contact

148

angle and zeta potential of samples.29,30 The contact angle and zeta potential were

149

measured as described previously.29,30 Detailed methodology and calculations can be

150

found in SI Texts S4 and S5, respectively.

151

Page 8 of 43

Microbial community analysis. Microbial DNA from consortia samples was

152

extracted using the FastDNA @ SPIN Kit for Soil (MP Biomedicals, USA), as per the

153

manufacturer’s instructions. Next, the V4 regions of bacterial 16S rRNA gene were

154

amplified with the primer pair 5158F/806R.31 Finally, the extracted and purified PCR

155

products were sequenced using the Illumina MiSeq platform at Majorbio Co., Ltd.

156

(Shanghai, China). Thereafter, raw sequences (NCBI SRA: SRP126708) were quality

157

filtered using the following criteria: sequence length shorter than 200 bp, average

158

sequence quality score less than 25, or sequences containing any primer mismatches,

159

barcode mismatches, ambiguous bases, or homopolymer runs exceeding two bases

160

were excluded from further sequence analysis.31 Furthermore, the biomarkers among

161

different phases were identified using the filtered data with linear discriminant

162

analysis (LDA) effect size (LEfSe) as describe.32 The Fluorescence in situ

163

hybridization (FISH) was used to validate the abundance of anammox species. Details

164

were presented in SI Text S6.

165

Metabolomic Profiling and Quantitation Analysis. The sample was harvested

166

after centrifugation at 4°C at 6000×g for 3 min and washing with 1× PBS buffer three 8

ACS Paragon Plus Environment

Page 9 of 43

Environmental Science & Technology

167

times. The sample was resuspended in 1.5 mL of cold ultrapure water (stored at 4 °C),

168

and the mixture was sonicated with a probe tip sonicator on ice (3–s intervals between

169

3-s pulses; 30 min, 30% of maximum intensity; Scientz, JY–92–IIN, 650W).33 Then,

170

6 mL of pre-chilled methanol/water (4:1, v/v) was subsequently added to the

171

samples.34 The samples were incubated at -80 °C for 2 h to precipitate the cellular

172

proteins. After centrifugation (20 min at 10,000× g at 4°C), the supernatants were

173

collected and dried under a gentle steam of nitrogen gas at room temperature.

174

Afterwards, the samples were stored in a -80 °C freezer and prepared for metabolites

175

that including intracellular and extracellular compounds determination using

176

high-performance liquid chromatography with tandem mass spectrometry

177

(HPLC–MS/MS). Furthermore, untargeted metabolomic analysis is used for global

178

metabolome analysis, while targeted analysis is more accurate for quantitation.35

179

Detailed metabolomics analysis methods are presented in SI Text S7. All tests were

180

repeated three times. Additionally, the detailed methods to determine the hydrolysates

181

of extracellular proteins and polysaccharides were presented in SI Text S8.

182

The metabolites that were identified in all four samples were used for statistical,

183

principal components analysis (PCA), and pathway analyses. Two–tail paired t-tests

184

were conducted in Excel. All the heatmaps and the PCA analysis in this study were

185

completed in R using the ggbiplot packages respectively. Prior to analysis, the

186

metabolic data were pretreated using the autoscaling method.36 Pathway analysis was

187

finished using Metaboanalyst (www.metaboanalyst.ca).

188

Results 9

ACS Paragon Plus Environment

Environmental Science & Technology

189

Page 10 of 43

Bacterial aggregation capacity and diameter size variations response to

190

operational phases. During reactor start-up, there were discrepant nitrogen

191

conversion ratios and nitrogen removal rates (NRRs) between two phases (Figure 1

192

and SI Figure S1). During phase I, both the nitrite to ammonium consumption ratio

193

(△NO2-–N/△NH4+–N) and the nitrate to ammonium conversion ratio

194

(△NO3-–N/△NH4+–N) varied considerably. And the reactor performance gradually

195

increased corresponding to the highest NRR of 124.52 g–N/(m3 d) in the end. In phase

196

II, the average △NO2-–N/△NH4+–N and △NO3-–N/△NH4+–N was 1.07±0.03 and

197

0.22±0.04, respectively. The total NRR increased sharply from 131.46 to 356.73

198

g–N/(m3 d) with the increased influent nitrogen concentrations from 120 mg/L to 300

199

mg/L.

200

To determine whether the aggregation behavior of anammox consortia varied with

201

the reactor operation, a comparison of aggregation capacity was conducted.

202

Interestingly, the aggregation capacity exhibited the periodic law of first increase, and

203

subsequently decreased during the reactor start-up. For example, it decreased to the

204

minimum in the beginning of phase II (S5) and then increased. It decreased again for

205

S8 and increased until the end of phase II (S10). The diameters of anammox

206

aggregates that were the intuitive reflection of aggregate capacity were shown in SI

207

Figure S2. Importantly, its variation tendency was all agreed with the variations of

208

aggregation capacity (r = 0.947, P = 0.000). Turbidity assay was also conducted to

209

support these results (SI Figure S3) (details were presented in SI Text S9).

210

Variation in EPS chemical structure and composition. To uncover the variation 10

ACS Paragon Plus Environment

Page 11 of 43

Environmental Science & Technology

211

in the EPS composition of anammox consortia, the concentrations of PN and PS were

212

determined. Changes in the concentration of PN, PS, and the ratio of PN/PS showed

213

different trends as follows [Figure 2(b)]: (i) PN concentration was relatively lower for

214

inoculum and the samples in the end of phase I (S4) and in the initial period of phase

215

II (S5 and S6). The average concentration of PN in phase II was significantly higher

216

than that in phase I (Kruskal–Wallis test, P < 0.03). (ii) Intriguingly, PS concentration

217

exhibited the periodic variation law of first increase, and subsequently decreased

218

during the reactor start-up. (iii) Integrating the change rules of these two indices, the

219

ratio of PN/PS changed periodically in two phases. It was relatively lower for the

220

inoculum (S1), at the end of phase I (S4), the initial period of phase II (S5) and S8.

221

Furthermore, the correlation between the aggregation ability and the concentration

222

of PN was significantly positive (r = 0.547, P < 0.003) [Figure 2(c)]. Importantly,

223

though there was no significant relationship with the concentration of PS [Figure 2(d)],

224

a relatively stronger and significantly more positive correlation was observed between

225

the aggregation capacity and ratio of PN/PS (r = 0.670, P = 0.000) than that observed

226

between the aggregation capacity and PN concentration [Figure 2(e)]. Consequently,

227

the PN/PS ratio, and not the individual concentration of PN or PS, showed a stronger

228

positive correlation with the aggregation capacity.

229

To unravel the functional groups of the anammox consortia surface, the EPS was

230

further analyzed by FTIR spectroscopy and XPS analysis. The FTIR results showed

231

that the variation in the chemical structure of EPS was in line with the composition

232

change (SI Figure S4). In addition, the absorption bands of lipids reached a maximum 11

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 43

233

in the end of phase I (S4) and afterwards became invisible in phase II (S5-S10). Most

234

importantly, the XPS analysis showed that the ratio of hydrophobic functional groups

235

also changed periodically along with the variation of aggregation capacity during the

236

reactor operation (SI Figure S5). Details are presented in SI Text S10.

237

Various bacterial interaction energies. To determine the key interaction energy

238

influencing the anammox aggregation, the contact angles and zeta potential combined

239

with an extended DLVO theory analysis were performed. The results of the contact

240

angles against the three types of liquids (water, ethanediol and propanetriol) and the

241

zeta potential in 0.1 M NaCl solution are shown in SI Tables S1 and S2, respectively.

242

Combining these experimental results, the interaction energies can be calculated using

243

the formula presented in SI Text S5. Figure 3 shows the curves for total energy (WT)

244

as a function of separation distance among bacteria of different samples. Interestingly,

245

the energy barriers that would prevent the bacteria from aggregating also varied

246

periodically. This finding was in line with the aggregation capacity results [Figure

247

2(a)].

248

In fact, the contributions of WA, WR and WAB in extend DLVO were significantly

249

dependent and discrepant. Here, though the WA and WR curves were nearly similar for

250

different consortia samples (SI Figure S6), the WAB curves that depended on the

251

hydrophobic interaction followed the periodical rule during the reactor start-up.

252

Importantly, there were obvious energy barriers while the WAB values of consortia

253

samples were higher, such as S1, S4 and S5. Thus, the WAB that on behalf on the

254

hydrophobic force played a decisive role in determining the aggregation process, 12

ACS Paragon Plus Environment

Page 13 of 43

Environmental Science & Technology

255

which agreed with the EPS content results. Therefore, it is important to elaborate on

256

the biosynthesis and degradation of anammox EPS that led to the variation in bacterial

257

hydrophobicity and aggregation ability.

258

Bacterial community shift. To identify the microbiological community shift

259

during the reactor operation, 16S rRNA gene sequencing and FISH combined with

260

PCA analysis were conducted. Though the abundance of 47 community members

261

(species with more than 1% relative abundance) changed markedly between phase I

262

and phase II [Figures 4(a) and 4(b)] (details were presented in SI Text S11 and

263

Figures S7 and S8), the PCA analysis showed the colonization-specific clustering of

264

phase I and phase II, respectively. Therefore, the bacterial community structures were

265

relatively stable in phase I and phase II, respectively.

266

Importantly, the variation in the Phycisphaerae classes and the Anaerolineaceae

267

families followed a periodic law [Figure 4(a)], which was the same as the variation in

268

PS [Figure 2(b)]. To explore whether there was a correlation between these species

269

and the PS in anammox consortia, a Pearson correlation analysis was carried out.

270

Surprisingly, the relationship between the ratio of PS and the sum of the relative

271

abundance of these two species was significantly negative (P < 0.03) [Figure 4(c)].

272

Metabolite profile related to anammox consortia EPS shift. To provide insight

273

into the possible metabolic pathways of EPS, the metabolites of one sample in phase I

274

(S3) and three samples in phase II (S5, S8 and S10) were examined. In all, the 383

275

metabolites that were detected in all four samples belonged to the following Kyoto

13

ACS Paragon Plus Environment

Environmental Science & Technology

276

Encyclopedia of Genes and Genomes metabolic pathways: amino acids,

277

carbohydrates, lipids, peptides, nucleotides, cofactors and vitamins.

278

Page 14 of 43

Figure 5(a) depicts the intensities of the detected metabolites in each sample. As

279

whole, dramatic changes of amino acids, sugar and glycerophospholipid were

280

observed in the metabolic activities of anammox consortia. PCA of the metabolic

281

profiles brought insight into the colonization-specific clustering of the four samples

282

[Figure 5(b)]. The detail differentiation among colonization states has been shown in

283

Figures 5(c)-(e). From these figures, we could distinguish specific biosynthetic

284

pathways for the predominant EPS components.

285

Figure 5(c) shows the biosynthetic pathway of amino acids. These amino acid

286

intensities of four samples were all significantly distinct, especially for phenylalanine

287

(Phe), leucine (Leu), alanine (Ala), threonine (Thr), glycine (Gly), glutamate (Glu),

288

valine (Val), and others. However, not every amino acid can be the predominant

289

component of extracellular proteins. The relationships of 20 amino acids and PN were

290

established using Pearson correlation analysis. The results showed that only seven

291

amino acids exhibited strongly positive correlations (P < 0.003), including Phe, Leu,

292

Ala, Thr, Gly, Glu and Val (SI Figure S9). Importantly, these seven amino acids were

293

indeed involved in the hydrolysates of extracellular PN (SI Figure S10). Therefore,

294

these seven amino acids might be the main synthetic monomers of extracellular

295

proteins.

296

Unlike the complex biosynthesis pathway of PN, the sugar metabolism of

14

ACS Paragon Plus Environment

Page 15 of 43

Environmental Science & Technology

297

anammox consortia is relatively straightforward. PS was biosynthesized mainly via

298

the uridine diphosphate-N-acetylglucosamine (UDP-GlcNAc) and uridine

299

diphosphate-N-acetylgalactosamine (UDP-GalNAc) biosynthetic pathways, since

300

these metabolites and their biosynthetic precursors (Fru-6P and Glu-6P) were all

301

detected in the hydrolysates of extracellular polysaccharides (SI Figure S10).

302

Additionally, it also can be supposed by the varied regulation of sugar metabolites

303

intensities and PS concentrations in different samples (Details were presented in Text

304

S12). Furthermore, the catabolism of UDP-hexosamine was more active for S10 than

305

that for S8 and the lower concentration of PS might be degraded for S10 (details were

306

presented in Text S12).

307

Except for PN and PS, phospholipids are the other important components of

308

microorganism EPS and have an important role in determining the hydrophobicity.5

309

Although there were many types of lipids detected, only phosphatidylcholine (PtC)

310

and phosphatidylethanolamine (PtE) biosynthesis pathways were relatively intact. The

311

intensities of phosphocholine (PC) and phosphoethanolamine (PE), which are the

312

precursors of PtC and PtE, respectively, varied unceasingly with the reactor start-up.

313

Integrating the results of FTIR analysis, the PtE that was formed by the reaction of PE

314

decarboxylation was supposed to be the dominant content of extracellular

315

phospholipids instead of PtC (details were presented in SI Text S13).

316

Discussion

317

Increased extracellular PN/PS improves anammox sludge aggregation capacity.

15

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 43

318

To our best knowledge, this is the first to present the aggregation capacity of

319

anammox consortia through the microbial and metabolome analysis during reactor

320

start-up, as aggregation capacity is important for anammox sludge retention and

321

accelerating anammox reactor start-up. The anammox process can be proved through

322

the bacterial community structure and nitrogen conversion ratios. Although both of

323

△NO2-–N/△NH4+–N and△NO3-–N/△NH4+–N did not completely conform with the

324

theoretical value of anammox process, they both agreed well with the anammox

325

process during bioreactor start–up.37 This was because an almost pure free-cells

326

suspension of highly active anammox was used to calculate the stoichiometry of

327

anammox process as described previously, but several other species of bacteria were

328

present in the anammox consortia community in the bioreactor.38 EPS, a complex

329

organic macromolecule mixture of polymers, is accountable for organism cohesion.5,15

330

Since PN and PS were predominantly present in EPS and the summation of these two

331

components concentration was regarded as the EPS concentration, the PN and PS

332

concentrations of EPS could focally affect the surface characteristics of anammox

333

consortia.14

334

Here, we first presented that the ratio of extracellular PN/PS showed a stronger

335

positive correlation with aggregation capacity of anammox consortia, which indicated

336

that it played more potentially important role in affecting the aggregation capacity.

337

Though it has been reported that PN, PS and PN/PS ratio all could determine the

338

microorganism aggregation,5-11 here, the aggregation capacity had a more strongly

339

and significantly positive relationship with the PN/PS ratio. Furthermore, the 16

ACS Paragon Plus Environment

Page 17 of 43

Environmental Science & Technology

340

aggregation capacity increased while the extracellular hydrophobic compositions and

341

interaction increased. Therefore, we supposed that the anammox aggregation was

342

determined by the hydrophobic components and force, which agreed with the results

343

of previous studies.8,39 As the aggregate caused by hydrophobic force that is a

344

hydrogen band in the micro-scale28 had more tight structure that could prevent oxygen

345

pervasion into the interior of aggregates1 than that caused by the alginate PS that

346

contains the diamond-shaped holes.6 The abundance of PN, as well as PS, affected the

347

hydrophobicity. Hence, the higher PN/PS ratio resulted from the metabolism of EPS

348

(detailed discussion in 4.2, 4.3 and 4.4), and not the individual PN or PS

349

concentration, contributed more to the anammox aggregates.

350

Active PN metabolism increases the bacterial aggregation capacity.

351

Considering that the bacterial community structure was relatively stable in different

352

phases and the aggregation capacity and EPS components varied constantly, the shift

353

in microbial community was not the only cause for the variations observed in the

354

aggregation capacity and EPS composition. Variations in the EPS composition also

355

corresponded with the metabolites. Additionally, all metabolites from the entire

356

biomass were extracted. Some identified metabolites from the inner of the aggregates

357

that contain the inert biomass may not contribute towards surface aggregation.

358

However, it has been reported that in microorganisms, the cell-free channel structures

359

were a strategy to transport metabolites from inner aggregation to the surface.40

360

Therefore, all identified metabolites from different layers of aggregates were assumed

361

to have the same contribution for surface aggregation. 17

ACS Paragon Plus Environment

Environmental Science & Technology

362

Page 18 of 43

In this study, the concentration of PN changed greatly, and those of S1 (inoculum),

363

S4 (end of Phase I), S5and S6 (initial of phase II) consortia were relatively lower than

364

those of other samples. The low protein concentration of S1 may have been caused by

365

enzyme hydrolysis when stored at 4 °C.41 The lower content of the other samples may

366

be attributed to the operational stage. Previous studies have shown that the EPS

367

composition of the same sludge is dynamic in different operational phases.42

368

Interestingly, the PN concentration of anammox consortia was relative lower for S4,

369

S5 and S6. Since there was a marked increase in the biomass concentration from S4 to

370

S6, this period was considered as the propagation phase (SI Figure S11). There are

371

different theories about the variation in bacterial extracellular PN in this phase. It has

372

been found that the PN concentration increased for hydrogenotrophic methanogens,43

373

while it was decreased for Rhodopseudomonas acidophila.44 In this study, the

374

reduction in the PN composition of anammox consortia was caused by the amino

375

acids for PN biosynthesis being used for cell growth and reproduction. Here, we

376

proposed that the main amino acids to synthesize protein included Phe, Leu, Ala, Thr,

377

Gly, Glu, and Val based on the Pearson correlation analysis. In addition, the

378

abundance of Gly, Glu, and Thr, which were selected on the basis of the correlation

379

analysis, was relatively lower in the hydrolysates of extracellular proteins. This may

380

be because amino acids were used as carbon and energy sources by heterotrophic

381

bacteria.1 Firstly, the intensities of these seven amino acids were significantly lower

382

for S5 compared to the other three samples (S3, S8 and S10) and led to a decrease in

383

PN production at this stage. Secondly, the possible participant in other roles of these 18

ACS Paragon Plus Environment

Page 19 of 43

Environmental Science & Technology

384

amino acids in metabolism including the biosynthesis of nucleotide sugar45-47 and

385

phospholipids48,49 could also cause a further decrease in excretion in propagation

386

phase (detailed discussion in 4.3, 4.4 and SI Text S14). Therefore, Phe, Leu, Ala, Thr,

387

Gly, Glu and Val were supposed to be the dominant residues of extracellular PN, and

388

their low concentrations for the synthesis of extracellular PN contributed to the low

389

aggregation capacity.

390

Furthermore, the average PN concentration of phase I was significantly lower than

391

that of phase II, especially than that of the end of phase II. These results might be

392

caused by the energy metabolism of anammox consortia to adapt to the environment

393

changes. Analysis of the purine metabolic pathway revealed that the intensities of

394

metabolites required for ATP synthesis, such as guanosine and AMP, were all higher

395

in S3 (phase I) than those in S8 and S10 (phase II) (SI Table S5), which indicated that

396

the energy metabolism during phase I was more active than that during phase II.

397

Importantly, it has been shown that the microorganism have a distinct upregulation of

398

energy metabolism to adapt to the new environment, and maintain steady-state

399

metabolism during the stationary phase.50 Hence, compared to phase II, the anammox

400

consortia need more energy in phase I. Additionally, it has been reported that EPS

401

synthesis costs energy, and decreased EPS secretion potentially conserves energy for

402

bacterial growth.51 Consequently, the energy that is produced might be preferentially

403

used for microbial growth, rather than for PN production in phase I, which may result

404

in lower PN concentration in phase I.

405

Polysaccharides degradation and inactive UDP-hexosamine biosynthesis 19

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 43

406

metabolism increases the bacterial aggregation capacity. Though the carbohydrate

407

structure in EPS was relatively stable, their content changed greatly and influenced

408

the characteristics of EPS.52 The results of this study agree with this finding.

409

Interestingly, the variation in PS concentration followed a periodic law, and we

410

proposed that one of the crucial reasons was the shift of PS degradation bacteria. It

411

has been shown that the Phycisphaerae classes can grow on cellobiose and

412

monosaccharides.53 Additionally, the Anaerolineaceae families have the ability to

413

degrade polysaccharides.54,55 Here, the variation in the sum of the abundance of these

414

two species also followed a periodic law and lagged behind the variation in PS. Most

415

importantly, there was a significantly negative relationship between the sum

416

abundance of these two species and the PS ratio. Just as the sugar metabolism for S10

417

was more active compared to S8, the sum of the relative abundance of these two

418

species was also higher for S10 than that for S8, which caused the PS concentration of

419

S10 to be lower than that of S8 and aggregation capacity of S10 to be higher than that

420

of S8. Therefore, these kinds of species could potentially increase the aggregation

421

capacity of anammox consortia through degrading PS. It’s also the possible reason

422

that the polysaccharides concentration and aggregation capacity changed periodically.

423

Furthermore, the biosynthesis of PS can be divided into two classes based on the

424

synthesis pathway.5,56 The first class comprises the production of

425

monopolysaccharides that are biosynthesized by the polymerization reaction of

426

monosaccharides including glucose and fructose. However, most PS is

427

heteropolysaccharides with irregular repeating units that are synthesized from a 20

ACS Paragon Plus Environment

Page 21 of 43

Environmental Science & Technology

428

variety of intracellular sugar nucleotides including glucose, GlcNAc and GalNAc,

429

among others.5,56 Here, we proposed that the PS of anammox consortia was mainly

430

biosynthesized through the UDP-GlcNAc and UDP-GalNAc synthesis pathways from

431

GlcNAc and GalNAc as precursors, which is the same as the lactic acid bacteria.56 At

432

the same time, the higher PS concentration of S5 was causd by the active metabolism

433

of sugar to biosynthesize cell tissues for anammox growth and reproduction (detailed

434

discussion in SI Text S15).5,57 Additionally, some types of amino acids might also be

435

translated into glucose and hexosamine via the EMP pathway for synthesis of

436

cell-building organisms for supplementation as described previously (detailed

437

discussion in SI Text S15).17,45,46,58 Taken together, the active biosynthesis of

438

UDP-GlcNAc and UDP-GalNAc was the fundamental cause of the high content of PS

439

and also led to the low aggregation capacity.

440

Active phosphatidylethanolamine (PtE) biosynthesis metabolism benefits the

441

bacterial aggregation capacity. It is known that hydrophobic phospholipids also

442

have effects on the bacterial aggregation ability.5 As the dominant compositions of

443

phospholipids, phosphatidylcholine (PtC) and phosphatidylethanolamine (PtE) can be

444

synthesized through three routes: CDP-choline and CDP-ethanolamine (Kennedy)

445

pathways and the CDP-diacylglycerol pathway. Therein, the Kennedy pathways

446

biosynthesized PtC and PtE from phosphocholine (PC) and phosphoethanolamine (PE)

447

as precursors, respectively, while the CDP-diacylglycerol pathway synthesized

448

phosphatidylserine from CDP-diacylglycerol as a precursor.59,60 Considering the

449

variation in phospholipids metabolites and extracellular phospholipids, we proposed 21

ACS Paragon Plus Environment

Environmental Science & Technology

450

that PtE might be the predominant component of extracellular phospholipids for

451

anammox consortia.

Page 22 of 43

452

Furthermore, the decreased intensities of phospholipids for consortia in phase II

453

were caused by the role and metabolism of PE and PC. Previous studies have shown

454

that PE and PC are the major components of phospholipids in most bacterial

455

membranes.20 It has been found that either PE or PC is the head-group of ladderane

456

lipids that play an important role in translocation of hydrazine in the anammoxosome

457

membrane to complete the anammox process.58,61 Hence, the metabolism of this type

458

of lipid should be active to transport the increasing hydrazine while the NRR

459

increased. Here, compared to PE, the metabolism and catabolism pathway of PC

460

became more active while the influent nitrogen concentration increased. Hence, we

461

supposed that PC instead of PE was the dominant head-group of the anammoxosome

462

membrane for anammox consortia and that the active metabolism of PC contributed to

463

the increased NRR for phase II, which was also one of the reasons for the lower PN

464

concentration.

465

Most importantly, the active metabolism of PtE could increase the aggregation

466

capacity of anammox consortia. Taking the case of S4 and S5, S4 consortia had a

467

lower PN concentration and a similar PN/PS ratio to S5. However, the aggregation

468

capacity of S4 was significantly higher than that of S5. This result was likely because

469

the active PtE metabolism in S4 contributed to the higher concentration of

470

hydrophobic lipids in EPS. Then, it contributed stronger hydrophobic interactions and

471

higher aggregation capacity for S4, which also agreed with results of the extended 22

ACS Paragon Plus Environment

Page 23 of 43

Environmental Science & Technology

472

DLVO theory. Therefore, the metabolism of phospholipids was also critical to

473

bacterial aggregation capacity.

474

Significance of this study. The determination of the aggregation capacity of

475

anammox consortia during reactor start-up has an important role in achieving rapid

476

biofilm formation, reactor start-up and advancing the anammox technology

477

applications for wastewater treatment. This study, for the first time, showed that the

478

anammox aggregation capacity changed periodically and was strongly correlated to

479

the extracellular PN/PS ratio and hydrophobic functional groups that are related to the

480

periodical variation of microbial community structure and discrepant metabolism. As

481

there was a great difference in bacterial aggregation capacity among the different

482

operational phases, optimal bio-carriers with special surface characteristics such as

483

hydrophobicity and zeta potential could be designed for different phases.

484

Importantly, active metabolism of the seven amino acids and PtE biosynthesis and

485

the inactive metabolism of the two UDP-hexosamine biosynthetic pathways have the

486

potential to improve the aggregation capacity. Therefore, the accommodation capacity

487

of anammox biomass can be improved through stimulating some metabolism

488

pathways. This study is first to elucidate the variation in aggregation capacity and

489

EPS composition of anammox consortia across the operational phases of the reactor

490

based on the metabolism profile, and it provides a methodology reference for similar

491

studies.

492

23

ACS Paragon Plus Environment

Environmental Science & Technology

493

Acknowledgments

494

The authors are grateful to the Foundations (NSFC No.51478006 and

495

No.JSGG20160429162015597) for financial support.

496

Supporting Information

497

Tables showing contact angle and zeta potential assay results, main bacterial taxa of

498

anammox consortia, metabolites intensities of purine pathway and pyruvate and

499

abbreviations of partial metabolites. Details of two aggregation capacity assays,

500

contact angle and zeta potential analysis, extended DLVO theory, FISH method,

501

untargeted and targeted metabolomic method, hydrolysates of extracellular proteins

502

and polysaccharides measurement methods, FTIR and XPS results, microbial

503

community shifts, sugar and glycerophospholipid vary, and discussions of amino

504

acids and nucleotide sugars. Figures illustrating the nitrogen conversion ratio,

505

aggregation capacity, aggregate diameter, FTIR and XPS analysis, extended DLVO

506

theory, FISH analysis, correlation between the amino acids and extracellular PN,

507

hydrolysates of extracellular PN and PS and biomass concentration.

Page 24 of 43

508

24

ACS Paragon Plus Environment

Page 25 of 43

Environmental Science & Technology

509

References

510

(1) Lawson, C. E.; Wu, S.; Bhattacharjee, A. S.; Hamilton, J. J.; McMahon, K. D.; Goel, R.;

511

Noguera, D. R. Metabolic network analysis reveals microbial community interactions in anammox

512

granules. Nat. Commun. 2017, 8, 15416.

513

(2) Lackner, S.; Gilbert, E. M.; Vlaeminck, S. E.; Joss, A.; Horn, H.; van Loosdrecht, M. C. M.

514

Full-scale partial nitritation/anammox experiences–An application survey. Water Res. 2014, 55,

515

292–303.

516

(3) Strous, M.; Fuerst, J. a; Kramer, E. H.; Logemann, S.; Muyzer, G.; van de Pas-Schoonen, K. T.;

517

Webb, R.; Kuenen, J. G.; Jetten, M. S. Missing lithotroph identified as new planctomycete. Nature

518

1999, 400 (6743), 446–449.

519

(4) Niu, Z.; Zhang, Z. T.; Liu, S. T.; Miyoshi, T.; Matsuyama, H.; Ni, J. R. Discrepant membrane

520

fouling of partial nitrification and anammox membrane bioreactor operated at the same nitrogen

521

loading rate. Bioresour. Technol. 2016, 214, 729–736.

522

(5) Flemming, H.; Wingender, J. The biofilm matrix. Nat. Rev. Microbiol. 2010, 8 (9), 623–633.

523

(6) Li, X. L.; Luo, J. H.; Guo, G.; Mackey, H. R.; Hao, T. W.; Chen, G. H. Seawater-based

524

wastewater accelerates development of aerobic granular sludge: A laboratory proof-of-concept.

525

Water Res. 2017, 115, 210–219.

526

(7) Tay, J. H.; Liu, Q. S.; Liu, Y. The role of cellular polysaccharides in the formation and

527

stability of aerobic granules. Lett. Appl. Microbiol. 2001, 33 (3), 222–226.

528

(8) Hou, X. L.; Liu, S. T.; Zhang, Z. T. Role of extracellular polymeric substance in determining 25

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 43

529

the high aggregation ability ofanammox sludge. Water Res. 2015, 75 (15), 51–62.

530

(9) Laspidou, C. S.; Rittmann, B. E. A unified theory for extracellular polymeric substances,

531

soluble microbial products, and active and inert biomass. Water Res. 2002, 36 (11), 2711–2720.

532

(10) Zhang, H. F.; Yu, H. H.; Zhang, L. H.; Song, L F.. Stratification structure of polysaccharides

533

and proteins in activated sludge with different aeration in membrane bioreactor. Bioresour.

534

Technol. 2015, 192, 361–366.

535

(11) Liao, B. Q.; Allen, D. G.; Droppo, I. G.; Leppard, G. G.; Liss, S. N. Surface properties of

536

sludge and their role in bioflocculation and settleability. Water Res. 2001, 35 (2), 339–350.

537

(12) Ni, S. Q.; Lee, P. H.; Fessehaie, A.; Gao, B. Y.; Sung, S. Enrichment and biofilm formation

538

of Anammox bacteria in a non-woven membrane reactor. Bioresour. Technol. 2010, 101 (6),

539

1792–1799.

540

(13) Meng, F. G.; Su, G. Y.; Hu, Y. F.; Lu, H.; Huang, L. N.; Chen, G. H. Improving nitrogen

541

removal in an ANAMMOX reactor using a permeable reactive biobarrier. Water Res. 2014, 58,

542

82–91.

543

(14) Yin, C. Q.; Meng, F. G.; Chen, G.-H. Spectroscopic characterization of extracellular

544

polymeric substances from a mixed culture dominated by ammonia-oxidizing bacteria. Water res.

545

2015, 68, 740–749.

546

(15) Sheng, G. P.; Yu, H. Q.; Li, X. Y. Extracellular polymeric substances (EPS) of microbial

547

aggregates in biological wastewater treatment systems: A review. Biotechnology Adv. 2010, 28 (6),

548

882–894.

26

ACS Paragon Plus Environment

Page 27 of 43

Environmental Science & Technology

549

(16) Gottschalk G. Bacterial metabolism, 2nd, ed.; Springer: New York, 1986.

550

(17) Michal, G.; Schomburg, D. Biochemical pathways: an atlas of biochemistry and molecular

551

biology, 2nd, ed.; Wiley: New York, 2012.

552

(18) An, W. X.; Guo, F. ; Song, Y. L.; Gao, N.; Bai, S. J.; Dai, J. C.; Wei, H. H.; Zhang, L. P.; Yu,

553

D. Z.; Xia, M.; Yu, Y.; Qi, M.; Tian, C. Y.; Chen, H. F.; Wu, Z. B.; Zhang, T.; Qiu, D. R.

554

Comparative genomics analyses on EPS biosynthesis genes required for floc formation of

555

Zoogloea resiniphila and other activated sludge bacteria. Water Res. 2016, 102, 494–504.

556

(19) Gangoiti, M. V.; Puertas, A. I.; Hamet, M. F.; Peruzzo, P. J.; Llamas, M. G.; Medrano, M.;

557

Prieto, A.; Due?as, M. T.; Abraham, A. G. Lactobacillus plantarum CIDCA 8327: an α-glucan

558

producing-strain isolated from kefir grains. Carbohyd. Polym. 2017, 170, 52–59.

559

(20) Sohlenkamp, C.; Geiger, O. Bacterial membrane lipids: Diversity in structures and pathways.

560

FEMS Microbiol. Rev. 2016, 40 (1), 133–159.

561

(21) Tang, X.; Liu, S. T.; Zhang, Z. T.; Zhuang, G. Q. Identification of the release and effects of

562

AHLs in anammox culture for bacteria communication. Chem. Eng. J. 2015, 273, 184–191.

563

(22) APHA. Standard Methods for the Examination of Water and Wastewater, 20th ed.; American

564

Public Health Association: Washington, DC, 1999.

565

(23) Xiong, Y. H.; Liu, Y. Importance of extracellular proteins in maintaining structural integrity

566

of aerobic granules. Colloid. Surface B. 2013, 112, 435–440.

567

(24) Zhang, W. J.; Chen, Z.; Cao, B. Di; Du, Y. I.; Wang, C. X.; Wang, D. S.; Ma, T.; Xia, H.

568

Improvement of wastewater sludge dewatering performance using titanium salt coagulants (TSCs) 27

ACS Paragon Plus Environment

Environmental Science & Technology

Page 28 of 43

569

in combination with magnetic nano-particles: Significance of titanium speciation. Water Res. 2017,

570

110, 102–111.

571

(25) Loewus, F. Improvement in anthrone method for determination of carbohydrates. Anal. Chem.

572

1952, 24 (1), 219–219.

573

(26) Osnes, T.; Sandstad, O.; Skar, V.; Osnes, M.; Kierulf, P. Total protein in common duct bile

574

measured by acetonitrile precipitation and a micro bicinchoninic acid (BCA) method. Scand. J.

575

Clin. Lab. Invest. 1993, 53 (7), 757–763.

576

(27) Sudmalis, D.; Gagliano, M. C.; Pei, R.; Grolle, K.; Plugge, C. M.; Rijnaarts, H. H. M.;

577

Zeeman, G.; Temmink, H. Fast anaerobic sludge granulation at elevated salinity. Water Res. 2018,

578

128, 293–303.

579

(28) Grasso, D.; Subramaniam, K.; Butkus, M.; Strevett, K.; Bergendahl, J. A review of

580

non-DLVO interactions in environmental colloidal systems. Rev. Environ. Sci. Biotechnol. 2002, 1

581

(1), 17–38.

582

(29) Liu, X. M.; Sheng, G. P.; Luo, H. W.; Feng, Z.; Yuan, S. J.; Wu, J. G.; Zeng, R. J.; Yu, H. Q.

583

Contribution of extracellular polymeric substances (EPS) to the sludge aggregation. Environ. Sci.

584

Technol. 2010, 44 (11), 4355–4360.

585

(30) Liu, X. M.; Sheng, G. P.; Yu, H. Q. DLVO approach to the flocculability of a photosynthetic

586

H2–producing bacterium, Rhodopseudomonas acidophila. Environm. Sci. Technol. 2007, 41 (13),

587

4620–4625.

588

(31) Bhattacharjee, A. S.; Wu, S.; Lawson, C. E.; Jetten, M. S. M.; Kapoor, V.; Domingo, J. W. S.;

28

ACS Paragon Plus Environment

Page 29 of 43

Environmental Science & Technology

589

McMahon, K. D.; Noguera, D. R.; Goel, R. Whole–community metagenomics in two different

590

anammox configurations: Process performance and community structure. Environ. Sci. Technol.

591

2017, 51 (8), 4317–4327.

592

(32) Segata, N.; Izard, J.; Waldron, L.; Gevers, D.; Miropolsky, L.; Garrett, W. S.; Huttenhower,

593

C. Metagenomic biomarker discovery and explanation. Genome Biol. 2011, 12 (6), R60.

594

(33) Worth, A. J.; Basu, S. S.; Snyder, N. W.; Mesaros, C.; Blair, I. A. Inhibition of neuronal cell

595

mitochondrial complex I with rotenone increases lipid β-oxidation, supporting acetyl-coenzyme a

596

levels. J. Biol. Chem. 2014, 289 (39), 26895–26903.

597

(34) Yuan, M.; Breitkopf, S. B.; Yang, X.; Asara, J. M. A positive/negative ion-switching,

598

targeted mass spectrometry-based metabolomics platform for bodily fluids, cells, and fresh and

599

fixed tissue. Nat. protoc. 2012, 7 (5), 872–881.

600

(35) Griffiths, W. J.; Koal, T.; Wang, Y. Q.; Kohl, M.; Enot, D. P.; Deigner, H. P. Targeted

601

metabolomics for biomarker discovery. Angew. Chem. Int. Ed. 2010, 49 (32), 5426–5445.

602

(36) van den Berg, R. a; Hoefsloot, H. C. J.; Westerhuis, J. a; Smilde, A. K.; van der Werf, M. J.

603

Centering, scaling, and transformations: improving the biological information content of

604

metabolomics data. BMC genomics 2006, 7, 142.

605

(37) van der Star, W. R. L.; Abma, W. R.; Blommers, D.; Mulder, J. W.; Tokutomi, T.; Strous, M.;

606

Picioreanu, C.; van Loosdrecht, M. C. M. Startup of reactors for anoxic ammonium oxidation:

607

Experiences from the first full-scale anammox reactor in Rotterdam. Water Res. 2007, 41 (18),

608

4149–4163.

29

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 43

609

(38) Lotti, T.; Kleerebezem, R.; Lubello, C.; van Loosdrecht, M. C. M. Physiological and kinetic

610

characterization of a suspended cell anammox culture. Water Res. 2014, 60, 1–14.

611

(39) Jia, F. X.; Yang, Q.; Liu, X. H.; Li, X. Y.; Li, B. K.; Zhang, L.; Peng, Y. Z. Stratification of

612

extracellular polymeric substances (EPS) for aggregated anammox microorganisms. Environ. Sci.

613

Technol. 2017, 51 (6), 4149–4163.

614

(40) Gonzalez-Gil, G.; Lens, P. N. L.; Van Aelst, A.; Van As, H.; Versprille, A. I.; Lettinga, G.

615

Cluster structure of anaerobic aggregates of an expanded granular sludge bed reactor. Appl.

616

Environ. Microbiol. 2001, 67 (8), 3683–3692.

617

(41) Adav, S. S.; Lee, D.-J.; Lai, J.-Y. Proteolytic activity in stored aerobic granular sludge and

618

structural integrity. Bioresour. Technol. 2009, 100 (1), 68–73.

619

(42) Ding, Z.; Bourven, I.; Guibaud, G.; van Hullebusch, E. D.; Panico, A.; Pirozzi, F.; Esposito,

620

G. Role of extracellular polymeric substances (EPS) production in bioaggregation: application to

621

wastewater treatment. Appl. Microbiol. Biotechnol. 2015, 99 (23), 9883–9905.

622

(43) Jia, X. S.; Furumai, H.; Fang, H. H. P. Extracellular polymers of hydrogen-utilizing

623

methanogenic and sulfate-reducing sludges. Water Res. 1996, 1354 (96), 1439–1444.

624

(44) Sheng, G. P.; Yu, H. Q. Relationship between the extracellular polymeric substances and

625

surface characteristics of Rhodopseudomonas acidophila. Appl. Microbiol. Biotechnol. 2006, 72

626

(1), 126–131.

627

(45) Lazar, C. S.; Baker, B. J.; Seitz, K. W.; Teske, A. P. Genomic reconstruction of multiple

628

lineages of uncultured benthic archaea suggests distinct biogeochemical roles and ecological

30

ACS Paragon Plus Environment

Page 31 of 43

Environmental Science & Technology

629

niches. ISME J. 2017, 11, 1118–1129.

630

(46) Ng, W. V.; Kennedy, S. P.; Mahairas, G. G.; Berquist, B.; Pan, M.; Shukla, H. D.; Lasky, S.

631

R.; Baliga, N. S.; Thorsson, V.; Sbrogna, J.; Swartzell, S.; Weir, D.; Hall, J.; Dahl, T. A.; Welti, R.;

632

Goo, Y. A.; Leithauser, B.; Keller, K.; Cruz, R.; Danson, M. J.; Hough, D. W.; Maddocks, D. G.;

633

Jablonski, P. E.; Krebs, M. P.; Angevine, C. M.; Dale, H.; Isenbarger, T. A.; Peck, R. F.;

634

Pohlschroder, M.; Spudich, J. L.; Jung, K. H.; Alam, M.; Freitas, T.; Hou, S. B.; Daniels, C. J.;

635

Dennis, P. P.; Omer, A. D.; Ebhardt, H.; Lowe, T. M.; Liang, R.; Riley, M.; Hood, L.; DasSarma,

636

S. Genome sequence of Halobacterium species NRC-1. Proc. Natl. Acad. Sci. U. S. A. 2000, 97

637

(22), 12176–12181.

638

(47) van Teeseling, M. C. F.; Mesman, R. J.; Kuru, E.; Espaillat, A.; Cava, F.; Brun, Y. V.;

639

VanNieuwenhze, M. S.; Kartal, B.; van Niftrik, L. Anammox Planctomycetes have a

640

peptidoglycan cell wall. Nat. Commun. 2015, 6, 6878.

641

(48) Boumann, H. A.; Hopmans, E. C.; Van De Leemput, I.; Op Den Camp, H. J. M.; Van De

642

Vossenberg, J.; Strous, M.; Jetten, M. S. M.; Sinninghe Damsté, J. S.; Schouten, S. Ladderane

643

phospholipids in anammox bacteria comprise phosphocholine and phosphoethanolamine

644

headgroups. FEMS Microbiol. Lett. 2006, 258 (2), 297–304.

645

(49) Medema, M. H.; Zhou, M. M.; van Hijum, S. A.; Gloerich, J.; Wessels, H. J.; Siezen, R. J.;

646

Strous, M. A predicted physicochemically distinct sub-proteome associated with the intracellular

647

organelle of the anammox bacterium Kuenenia stuttgartiensis. BMC Genomics 2010, 11, 299.

648

(50) Rolfe, M. D.; Rice, C. J.; Lucchini, S.; Pin, C.; Thompson, A.; Cameron, A. D. S.; Alston, M.;

649

Stringer, M. F.; Betts, R. P.; Baranyi, J.; Peck, M. W.; Hinton, J. C. D. Lag phase is a distinct 31

ACS Paragon Plus Environment

Environmental Science & Technology

Page 32 of 43

650

growth phase that prepares bacteria for exponential growth and involves transient metal

651

accumulation. J. Bacteriol. 2012, 194 (3), 686–701.

652

(51) Raszka, A.; Chorvatova, M.; Wanner, J. The role and significance of extracellular polymers

653

in activated sludge. Part I: Literature review. Acta Hydroch. Hydrob. 2006, 34 (5), 411–424.

654

(52) Rehm, B. H. A. Microbial production of biopolymers and polymer precursors: applications

655

and perspectives; Caister Academic Press: New Zealand, 2009.

656

(53) Wang, X.; Sharp, C. E.; Jones, G. M.; Grasby, S. E.; Brady, A. L.; Dunfield, F. Stable-isotope

657

probing identifies uncultured planctomycetes as primary degraders of a complex

658

heteropolysaccharide in soil. Appl. Environ. Microbol. 2015, 81 (14), 4607–4615.

659

(54) Ahn, J. H.; Song, J.; Kim, B. Y.; Kim, M. S.; Joa, J. H.; Weon, H. Y. Characterization of the

660

bacterial and archaeal communities in rice field soils subjected to long-term fertilization practices.

661

J. Microbiol. 2012, 50 (5), 754–765.

662

(55) Terashima, M.; Yama, A.; Sato, M.; Yumoto, I.; KamagaTa, Y.; Kato, S. Culture-dependent

663

and independent identification of polyphosphate-accumulating Dechloromonas spp.

664

predominating in a full-scale oxidation ditch wastewater treatment plant. Microbes Environ 2016,

665

31 (4), 449–455.

666

(56) Boels, I. C.; van Kranenburg, R.; Hugenholtz, J.; Kleerebezem, M.; de Vos, W. M. Sugar

667

catabolism and its impact on the biosynthesis and engeneering of exopolysaccahrides production

668

in lactic acid bacteria. Int. Diary J. 2001, 11, 723–732.

669

(57) van Teeseling, M. C. F.; Maresch, D.; Rath, C. B.; Figl, R.; Altmann, F.; Jetten, M. S. M.;

32

ACS Paragon Plus Environment

Page 33 of 43

Environmental Science & Technology

670

Messner, P.; Schäffer, C.; van Niftrik, L. The S-layer protein of the anammox bacterium Kuenenia

671

stuttgartiensis is heavily O-Glycosylated. Front. Microbiol. 2016, 7, 1–10.

672

(58) Strous, M.; Pelletier, E.; Mangenot, S.; Rattei, T.; Lehner, A.; Taylor, M. W.; Horn, M.;

673

Daims, H.; Bartol-Mavel, D.; Wincker, P. Barbe, V.; Fonknechten, N.; Vallenet, D.; Segurens, B.;

674

Schenowitz-Truong, C.; Medigue, C.; Collingro, A.; Snel, B.; Dutilh, B. E.; Op den Camp, H. J.

675

M.; van der Drift, C.; Cirpus, I.; van de Pas-Schoonen, K. T.; Harhangi, H. R.; van Niftrik, L.;

676

Schmid, M.; Keltjens, J.; van de Vossenberg, J.; Kartal, B.; Meier, H.; Frishman, D.; Huynen, M.

677

A.; Mewes, H. W.; Weissenbach, J.; Jetten, M. S. M.; Wagner, M.; Le Paslier, D. Deciphering the

678

evolution and metabolism of an anammox bacterium from a community genome. Nature 2006,

679

440 (7085), 790–794.

680

(59) Gibellini, F.; Smith, T. K. The Kennedy pathway-de novo synthesis of

681

phosphatidylethanolamine and phosphatidylcholine. IUBMB Life 2010, 62 (6), 414–428.

682

(60) P Pessi, G.; Kociubinski, G.; Mamoun, C. B. A pathway for phosphatidylcholine biosynthesis

683

in Plasmodium falciparum involving phosphoethanolamine methylation. Proc. Natl. Acad. Sci. U.

684

S. A. 2004, 101 (16), 6206–6211.

685

(61) Chaban, V. V.; Nielsen, M. B.; Kopec, W.; Khandelia, H. Insights into the role of cyclic

686

ladderane lipids in bacteria from computer simulations. Chem. Phys. Lipids 2014, 181, 76–82.

687

33

ACS Paragon Plus Environment

Environmental Science & Technology

Page 34 of 43

688

Figure captions

689

Figure 1. Nitrogen concentration profile and removal rate during anammox reactor

690

start-up: (a) NH4+-N, NO2--N and NO3--N concentrations in the influent and effluent

691

in phase I (days 1-90) and phase II (days 91-210). (b) NH4+-N and NO2--N removal

692

efficiency, as well as nitrogen removal rate in two phases.

693

Figure 2. (a) The aggregation capacity for ten anammox consortia samples during the

694

reactor start-up. S1-S10 represent that the inoculum samples and the samples obtained

695

on day 30, 60, 90, 120, 150, 165, 180, 195, and 210, respectively. Red, green and blue

696

columns indicate the samples in inoculum, phase I and phase II, respectively. (b) The

697

extracellular proteins (red columns) and polysaccharides (grey columns) contents and

698

the ratio of protein to polysaccharides (blue columns) of anammox consortia in

699

inoculum, phase I and phase II. Error bars represent standard deviations of triplicate.

700

(c), (d) and (e) show the correlation relationships between the sludge aggregation

701

capacity and the concentration of extracellular PN, PS and PN/PS ratio, respectively.

702

The red open circles depict the value of these three indexes. The dark grey lines

703

indicate fitted linear models with confidence intervals of 0.95 (light grey lines), which

704

were established by the Pearson correlation analysis. P < 0.05 was considered

705

statistically significant in all statistical analysis.

706

Figure 3. Total energy of the interaction (WT) vs. separation distance curves of

707

anammox consortia samples of (a) inoculum (S1), (b) phase I (S2-S4), (c) phase II

708

(S5-S10) calculated by extended DLVO theory. The electric double layer (WR), the 34

ACS Paragon Plus Environment

Page 35 of 43

Environmental Science & Technology

709

van der Waals energies (WA) and the acidbase interaction (WAB) vs. separation

710

distance curves were presented in SI Figure S4.

711

Figure 4. (a)Heatmap showing relative abundance of 47 kinds of predominant

712

bacterial community types (phyla with more than 1% relative abundance) detected in

713

samples from inoculum and two phases. The detailed taxonomy was presented in SI

714

Table S3. (b) Principal component analysis (PCA) plot of bacterial taxa of anammox

715

consortia among different phases. Red, green and blue colonization-specific clustering

716

indicates the samples of inoculum, phase I and phase II, respectively. (c) The

717

correlation relationship between the sum relative abundance of the Phycisphaerae

718

classes and the Anaerolineaceae families and the ratio of extracellular

719

polysaccharides in the sum of extracellular proteins and polysaccharides. Correlation

720

relationship was assessed by the Pearson analysis. The dark grey lines indicate fitted

721

linear models with confidence intervals of 0.95 (light grey lines). P < 0.05 was

722

considered statistically significant.

723

Figure 5. (a) The heatmap showing relative abundance of metabolites in four samples

724

(S3, S5, S8 and S10) determined by the untargeted and targeted metabolomic analysis.

725

S3, S5, S8 and S10 represented the samples obtained on day 60, 120, 180 and 210,

726

respectively. Each column represents one biological replicate and every sample was

727

conducted in triplicate. (b) PCA analysis plot of anammox consortia metabolites in

728

four samples. (c) (d) and (e) represent metabolic pathway of amino acids,

729

polysaccharides and glycerophospholipid. Red, green, blue and purple circle represent

730

the metabolite intensity of the samples obtained on day 60 (S3), 150 (S5), 180 (S8) 35

ACS Paragon Plus Environment

Environmental Science & Technology

Page 36 of 43

731

and 210 (S10), respectively. The different circle’s diameter represents the different

732

metabolite intensity. And the metabolite intensity of S3 was used as a control with the

733

fixed diameter. The circle’s diameters of other three samples were same as that of S3

734

when the metabolites intensities changed no significance, and were bigger or smaller

735

than that of S3 when the metabolites intensities were significantly increased or

736

decreased, respectively (P < 0.05). The solid lines represent the intracellular

737

metabolic pathways, and the dotted lines represent the supposed extracellular

738

metabolic pathways.

739

36

ACS Paragon Plus Environment

Page 37 of 43

Environmental Science & Technology

740 741

Figure 1. Nitrogen concentration profile and removal rate during anammox reactor

742

start-up: (a) NH4+-N, NO2--N and NO3--N concentrations in the influent and effluent

743

in phase I (days 1-90) and phase II (days 91-210). (b) NH4+-N and NO2--N removal

744

efficiency, as well as nitrogen removal rate in two phases.

37

ACS Paragon Plus Environment

Environmental Science & Technology

Page 38 of 43

745 746

Figure 2. (a) The aggregation capacity for ten anammox consortia samples during the

747

reactor start-up. S1-S10 represent that the inoculum samples and the samples obtained

748

on day 30, 60, 90, 120, 150, 165, 180, 195, and 210, respectively. Red, green and blue

749

columns indicate the samples in inoculum, phase I and phase II, respectively. (b) The

750

extracellular proteins (red columns) and polysaccharides (grey columns) contents and

751

the ratio of protein to polysaccharides (blue columns) of anammox consortia in

752

inoculum, phase I and phase II. Error bars represent standard deviations of triplicate.

753

(c), (d) and (e) show the correlation relationships between the sludge aggregation

754

capacity and the concentration of extracellular PN, PS and PN/PS ratio, respectively.

755

The red open circles depict the value of these three indexes. The dark grey lines

756

indicate fitted linear models with confidence intervals of 0.95 (light grey lines), which

757

were established by the Pearson correlation analysis. P < 0.05 was considered

758

statistically significant in all statistical analysis.

759 38

ACS Paragon Plus Environment

Page 39 of 43

Environmental Science & Technology

760

761

Figure 3. Total energy of the interaction (WT) vs. separation distance curves of

762

anammox consortia samples of (a) inoculum (S1), (b) phase I (S2-S4), (c) phase II

763

(S5-S10) calculated by extended DLVO theory. The electric double layer (WR), the

764

van der Waals energies (WA) and the acidbase interaction (WAB) vs. separation

765

distance curves were presented in SI Figure S4.

766

39

ACS Paragon Plus Environment

Environmental Science & Technology

Page 40 of 43

767 768

Figure 4. (a)Heatmap showing relative abundance of 47 kinds of predominant

769

bacterial community types (phyla with more than 1% relative abundance) detected in

770

samples from inoculum and two phases. The detailed taxonomy was presented in SI

771

Table S3. (b) Principal component analysis (PCA) plot of bacterial taxa of anammox

772

consortia among different phases. Red, green and blue colonization-specific clustering

773

indicates the samples of inoculum, phase I and phase II, respectively. (c) The

774

correlation relationship between the sum relative abundance of the Phycisphaerae

775

classes and the Anaerolineaceae families and the ratio of extracellular

776

polysaccharides in the sum of extracellular proteins and polysaccharides. Correlation

777

relationship was assessed by the Pearson analysis. The dark grey lines indicate fitted

778

linear models with confidence intervals of 0.95 (light grey lines). P < 0.05 was

779

considered statistically significant.

780 40

ACS Paragon Plus Environment

Page 41 of 43

Environmental Science & Technology

781 782

Figure 5. (a) The heatmap showing relative abundance of metabolites in four samples

783

(S3, S5, S8 and S10) determined by the untargeted and targeted metabolomic analysis.

784

S3, S5, S8 and S10 represented the samples obtained on day 60, 120, 180 and 210,

785

respectively. Each column represents one biological replicate and every sample was

786

conducted in triplicate. (b) PCA analysis plot of anammox consortia metabolites in

787

four samples. (c) (d) and (e) represent metabolic pathway of amino acids,

788

polysaccharides and glycerophospholipid. Red, green, blue and purple circle represent

789

the metabolite intensity of the samples obtained on day 60 (S3), 150 (S5), 180 (S8) 41

ACS Paragon Plus Environment

Environmental Science & Technology

Page 42 of 43

790

and 210 (S10), respectively. The different circle’s diameter represents the different

791

metabolite intensity. And the metabolite intensity of S3 was used as a control with the

792

fixed diameter. The circle’s diameters of other three samples were same as that of S3

793

when the metabolites intensities changed no significance, and were bigger or smaller

794

than that of S3 when the metabolites intensities were significantly increased or

795

decreased, respectively (P < 0.05). The solid lines represent the intracellular

796

metabolic pathways, and the dotted lines represent the supposed extracellular

797

metabolic pathways.

798

42

ACS Paragon Plus Environment

Page 43 of 43

Environmental Science & Technology

799

TOC/Abstract graphic

800 801

For Table of Contents Only

802

43

ACS Paragon Plus Environment