Multivariate and Multiscale Approaches for Interpreting the

Jul 9, 2018 - (27−29) Pig manure (10.0 kg), wheat straw (1.0 kg), and deionized water (1.0 kg) ..... The composting samples of the early phase (days...
1 downloads 0 Views 2MB Size
Subscriber access provided by UOW Library

Environmental Processes

Multivariate and multi-scale approaches for interpreting the mechanisms of nitrous oxide emission during pig manure–wheat straw aerobic composting Jinyi Ge, Guangqun Huang, Junbao Li, Xiaoxi Sun, and Lujia Han Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b02958 • Publication Date (Web): 09 Jul 2018 Downloaded from http://pubs.acs.org on July 10, 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 40

Environmental Science & Technology

1

Multivariate and multi-scale approaches for interpreting the mechanisms of nitrous oxide

2

emission during pig manure–wheat straw aerobic composting

3 4

Jinyi Ge, Guangqun Huang, Junbao Li, Xiaoxi Sun, Lujia Han*

5

Biomass Resources and Utilization Laboratory, College of Engineering, China Agricultural

6

University, Beijing 100083, China

7 8 9

* Corresponding author.

10

China Agricultural University (East Campus), Box 191, Beijing 100083, China

11

Fax: 86-10-6273-6778

12

Tel: 86-10-6273-6313

13

E-mail: [email protected]

1 Environment ACS Paragon Plus

Environmental Science & Technology

14

ABSTRACT: Nitrous oxide (N2O) emission during composting causes nitrogen loss and air

15

pollution. The interpretation of N2O emission mechanisms will help to customize composting

16

strategies that mitigates climate change. At pile and particle scales, this study characterized N2O

17

emission-related variables (gases, ions, and microbes) and their correlations during pig manure–

18

wheat straw aerobic composting. Pile-scale results showed that N2O emission mainly occurred in

19

mesophilic, thermophilic, and cooling phases; the nitrification by ammonia-oxidizing bacteria

20

(AOB) and nitrite-oxidizing bacteria (NOB) coexisted with the denitrification by denitrificans

21

(DEN); the major NOB and DEN were Nitrobacter (NOB_Nba) and Thiobacillus denitrificans

22

(DEN_Tb), respectively. The mechanisms of nitrification, nitrifier denitrification, and anaerobic

23

denitrification in composting particles were initially visualized by confocal laser scanning

24

microscopy: Betaproteobacteria (AOB_Beta) sporadically distributed on the outer area of the

25

particles, NOB_Nba internally attached to AOB_Beta, and Nitrosomonas europea/Nitrosomonas

26

eutropha (AOB_eu) and DEN_Tb concentrated in the interior. Correlation analysis of the

27

variables showed that the distribution area of AOB_eu was proportional to N2O emission (R2 =

28

0.84); AOB not only participated in nitrification but also nitrifier denitrification; N2O formation

29

was mainly from nitrifier denitrification by AOB_eu during the mesophilic–thermophilic phase

30

and from denitrification by AOB_eu and DEN during the cooling phase.

31 32

KEYWORDS: aerobic composting, nitrous oxide emission mechanism, multi-scale,

33

multivariate

2 Environment ACS Paragon Plus

Page 2 of 40

Page 3 of 40

Environmental Science & Technology

34 35

TOC Art

3 Environment ACS Paragon Plus

Environmental Science & Technology

36

INTRODUCTION

37

The latest World Meteorological Organization (WMO) Greenhouse Gas Bulletin reports

38

that the globally averaged mole fraction of nitrous oxide (N2O) reached a new high of 328.9 ±

39

0.1 ppb in 2016.1 As a significant source of greenhouse gas emissions, N2O has a global

40

warming potential 265 times that of an equal mass of carbon dioxide (CO2), and it can be

41

transported to the stratosphere, which will cause ozone depletion. Accumulative N2O emission

42

during composting is an important concern, as it could account for 10% of the total nitrogen

43

content of nitrogen-rich composting materials such as manure,2,3 which not only exacerbates air

44

pollution but also results in nitrogen loss in fertilizer.4,5 Therefore, an understanding of the

45

mechanisms of N2O emission during composting is critical to provide theoretical guidance for

46

reducing the environmental impact of N2O.

47

Based on the pile scale, previous research analyzed N2O emission-related variables (gases,

48

ions, and microbes) and established preliminary mechanisms of N2O emission during

49

composting. The results showed that there were two pathways of N2O emission, i.e., aerobic

50

nitrification and anaerobic denitrification.5-7 During aerobic nitrification, ammonia-oxidizing

51

bacteria (AOB) and/or ammonia-oxidizing archaea (AOA) oxidize ammonium (NH4+) to nitrite

52

(NO2–); subsequently, nitrite-oxidizing bacteria (NOB) oxidize NO2– to nitrate (NO3–); as an

53

intermediate product of ammonia oxidation, hydroxylamine (NH2OH) is oxidized to N2O.8-10

54

During denitrification, NO2– is reduced by denitrificans (DEN) anaerobically, where N2O is an

55

intermediate.11 However, according to the research in the area of wastewater treatment, nitrifiers

56

such as Nitrosomonas europea/Nitrosomonas eutropha (AOB_eu) belonging to the AOB group

57

could also execute denitrification and generate N2O, which is called nitrifier denitrification.12-14

58

Does nitrifier denitrification occur during aerobic composting and contribute to N2O emission?

4 Environment ACS Paragon Plus

Page 4 of 40

Page 5 of 40

59

Environmental Science & Technology

This question needs to be further addressed.15

60

The above research based on the pile scale established a theoretical foundation for

61

interpreting the mechanisms of N2O emission during composting. However, these studies

62

assuming completely aerobic conditions cannot be used to analyze N2O generated from

63

anaerobic denitrification.16,17 The availability of oxygen (O2) is a determining factor in N2O

64

production. Manure composts have been considered a microaerobic environment featuring a high

65

mineral nitrogen availability and a high concentration of organic carbon resource, which would

66

provide favourable conditions for N2O production.18 The aerobic layer–anaerobic core structure

67

of composting particles could be a starting point for illumination of the coupling of aerobic

68

nitrification, anaerobic denitrification, and N2O emission during composting.19-22 Based on

69

Fourier transform infrared microspectroscopy, Ge et al.23 determined the dynamic distribution of

70

lipids and aromatics in the composting particles and characterized the aerobic layer–anaerobic

71

core structure of the composting particles. Furthermore, Wang et al.24 examined the depth

72

profiles of the concentration of dissolved O2 and NO3– in manure particles using microelectrodes.

73

The results supported the finding concerning the aerobic layer–anaerobic core structure of

74

manure particles and showed a positive correlation between the concentration of NO3– and

75

dissolved O2. However, as composting is a biochemical process, the distribution of the chemical

76

compounds alone cannot reveal the microbial reactions involved in N2O emission. Fortunately, a

77

method for micro-zone analysis, fluorescent in situ hybridization–confocal laser scanning

78

microscopy (FISH–CLSM), could provide reference for studying the N2O emission-related

79

microbial reactions in the composting particles.25,26

80

The aims of this study were to interpret the mechanisms of N2O emission during pig

81

manure–wheat straw aerobic composting by characterizing the multiple variables at pile and

5 Environment ACS Paragon Plus

Environmental Science & Technology

82

particle scales and their correlations, including quantification of gases [O2, ammonia (NH3), and

83

N2O], ions (NH2OH, NH4+, NO2–, and NO3–), and microbes (AOB, AOA, NOB, and DEN) at the

84

pile scale and visualization of the microbial distribution (AOB, NOB, and DEN) at the particle

85

scale. The research could be useful to reduce nitrogen loss during composting and customize

86

composting strategies that mitigate greenhouse gas emissions.

87

MATERIALS AND METHODS

88

Aerobic composting experiments. The schematic of the aerobic composting reactor, the

89

details of composting experiments, and the methods for determining basic physicochemical

90

parameters can be found in previous studies.27-29 Pig manure (10.0 kg), wheat straw (1.0 kg), and

91

deionized water (1.0 kg) were raw materials and mixed thoroughly by hand. The mixture was

92

divided equally into two composting reactors (height 0.40 m, diameter 0.25 m, and effective

93

volume 16.00 L) and noted as experiments A and B. The duration of the experiments was 16

94

days. Based on the results of previous research,27 intermittent aeration (1 h on/1 h off) with a rate

95

of 0.35 L min–1 was employed.

96

The sampling was performed on days 0, 2, 4, 6, 8, 12, and 16. Moisture content and organic

97

matter content of the samples were determined by standard Test Methods for the Examination of

98

Composting and Compost (TMECC).30 Total carbon content and total nitrogen content were

99

tested with an elemental analyzer (Vario EL CHNOS, Elementar Analysensysteme GmbH,

100

Germany) to calculate the carbon to nitrogen ratio. Each measurement was performed in

101

triplicate, and the mean values were presented as the results. The results of the physicochemical

102

parameters of the initial composting mixtures are presented in Table 1. The values of the

103

moisture content, the organic matter content, and the carbon to nitrogen ratio met the

104

requirements of composting facilities.6 As the parameters except for the organic matter content

6 Environment ACS Paragon Plus

Page 6 of 40

Page 7 of 40

Environmental Science & Technology

105

showed no significant difference between experiments A and B, the related data of the

106

experiments were averaged for the purposes of the rest of the analysis.

107

Table 1. Physicochemical parameters of the initial composting mixtures. Physicochemical parameter

Experiment A

Experiment B

p

Moisture content (%)

56.17 ± 2.59

56.47 ± 2.84

0.897

Organic matter content (%, DM) 83.25 ± 1.35

79.25 ± 1.21

0.019

Total carbon (%, DM)

42.02 ± 0.15

43.87 ± 2.19

0.216

Total nitrogen (%, DM)

2.55 ± 0.21

2.78 ± 0.09

0.160

Carbon to nitrogen ratio

16.56 ± 1.41

15.80 ± 1.00

0.489

108

Note: values are presented as the mean ± standard deviation determined from three replicates; p,

109

significance level; p < 0.05 indicates significant difference between experiments A and B; DM,

110

dry weight basis.

111

Measurement of temperature, O2 concentration, NH3 emission, and N2O emission at

112

the pile scale. The composting temperature and ambient temperature were monitored daily using

113

thermocouples (Pt100, Omega Engineering Inc., CT, USA) located in the middle of the

114

composting mixtures and the surrounding atmosphere, respectively. Data from the

115

thermocouples were documented using a programmed data acquisition system (DT85; DataTaker

116

Pty Co., Ltd., Australia). The O2 concentration (volume fraction) in the upper part of the reactors

117

was measured using an O2 sensor (O2S-FR-T2-18X, Apollo Electronics Co., Ltd., China).

118

Simultaneously, gaseous samples in the upper part of the reactors were collected in 1-L gas bags.

119

The NH3 concentration in the bags was immediately analyzed by gas detector tubes (GASTEC,

120

Japan). The N2O concentration in the bags was immediately analyzed by gas chromatography

7 Environment ACS Paragon Plus

Environmental Science & Technology

Page 8 of 40

121

using a greenhouse gas analyzer (GC-2014; Shimadzu, Japan). Each measurement was

122

performed in triplicate, and the results were presented as the mean values.

123

Measurement of NH2OH, NH4+, NO2–, and NO3– at the pile scale. A spectrophotometric

124

method was verified by Frear et al. for determining the NH2OH concentration in soybean

125

leaves.31 However, in the case of crude samples such as composting particles, specific extraction

126

is demanded. The extraction of composting samples was conducted according to the TMECC

127

standard.30 Briefly, the composting samples were diluted by deionized water (1:10) and the

128

solutions were shaken for 2 h at room temperature; the solutions were centrifuged at 9000 × g

129

(4°C, 15 min) and then filtered by 0.45-µm Millipore filter papers. The NH2OH concentration of

130

the filtrates was then quantified by a spectrophotometer (UV-2550, Shimadzu, Japan) according

131

to the study of Frear et al.31 The concentration of NH4+, NO2–, and NO3– of the filtrates was

132

determined by a flow injection analyzer (FIAstar5000, Foss, Denmark). The unit was µmol N g–1.

133

Each measurement was performed in triplicate, and the results were presented as the mean values.

134

Real-time quantitative polymerase chain reaction (qPCR) analysis of the bacterial

135

abundance of AOB, AOA, NOB, and DEN at the pile scale. According to previous research,

136

the main AOB present during composting was Betaproteobacteria (AOB_Beta);8 the main NOB

137

included Nitrobacter (NOB_Nba) and Nitrospira (NOB_Nsp);10,32-34 the main DEN included

138

Azoarcus-Thauera

139

Thiobacillus denitrificans (DEN_Tb);11,35 the key functional genes of denitrification included

140

nitrite reductase (nirS + nirK) and nitrous oxide reductase (nosZ).6 The details of the qPCR

141

analysis and the primers used for AOB_Beta, NOB_Nba, NOB_Nsp, DEN_AT, DEN_P, DEN_Tb,

142

nirS + nirK, and nosZ are listed in Section 1 and Table S1 in Supporting Information. Each

143

qPCR measurement was performed in triplicate, and the results were presented as the mean

denitrificans

(DEN_AT),

Paracoccus

8 Environment ACS Paragon Plus

denitrificans

(DEN_P),

and

Page 9 of 40

Environmental Science & Technology

144

values. The qPCR results showed that there was no AOA, NOB_Nsp, DEN_AT, or DEN_P during

145

pig manure–wheat straw composting.

146

FISH–CLSM analysis of AOB_Beta, AOB_eu, NOB_Nba, and DEN_Tb at the particle

147

scale. The preparation of particle samples for FISH–CLSM analysis is detailed in Section 2 in

148

Supporting Information. In preliminary experiments, we embedded the particles in an optimal

149

cutting temperature (OCT) compound (Sakura Finetek, Torrance, CA) and tried to section them

150

into 10-, 20-, and 50-µm slices with a CM 3050-S cryostat (Leica Microsystems GmbH, Wetzlar,

151

Germany). Sectioning was technically difficult due to the fragile structure of the particles. Some

152

previous FISH–CLSM research of sludge flocs and mangrove roots did not adopt sectioning.36-38

153

Szilveszter et al.37 indicated that the CLSM images of sludge flocs without sectioning were

154

similar to those after being sectioned (5 µm), and fluorescent probes penetrated well into the

155

uncut flocs. The reason might be that, unlike biological tissues that are relatively dense and

156

constitute diffusion barriers hindering the penetration of probes, the loose structure of sludge

157

flocs would allow a better penetration, even though they were not sectioned. These results

158

implied that sectioning might not be mandatory for the FISH–CLSM analysis of delicate samples.

159

Therefore, sectioning was not applied to the composting particles in this study.

160

The thickness of the particles lying on the slides was examined by a micro-CT system

161

(Skyscan 1275, Bruker microCT, Belgium). A total of 20 particles were investigated, and the

162

thickness was measured to be 86.3 ± 24.1 µm (Supporting Information, Figure S1). Given that

163

the average diameter of manure particles was 506 ± 5 µm,28 the thickness of the particles lying

164

on the slides was much lower than the average diameter. This suggests that the particles may

165

have been shaved, in effect due to the steps of the preparation process, such as centrifugation and

166

rinse, and then as a result presented a cross section that gave us access to the interior of the

9 Environment ACS Paragon Plus

Environmental Science & Technology

167

particles.

168

The qPCR results showed that there was no AOA, NOB_Nsp, DEN_AT, or DEN_P in the

169

composting samples. Therefore, the research objects in FISH–CLSM analysis were AOB_Beta,

170

AOB_eu, NOB_Nba, and DEN_Tb. In case multiple hybridization cycles on the same slide cause

171

damage to the particles, three groups of double hybridization (i.e., NSO190–NSO1225 probe for

172

AOB_Beta and Nse1472 probe for AOB_eu, NSO190–NSO1225 probe for AOB_Beta and NIT3

173

probe for NOB_Nba, and NSO190–NSO1225 probe for AOB_Beta and TBD1419 probe for

174

DEN_Tb) were conducted separately. The details of the fluorescent probes and hybridization

175

conditions are listed in Table S2 and Section 3 in Supporting Information.

176

Fluorescent images of the hybridized composting particles on the slides were taken by a

177

confocal microscope (TCS SP5, Leica Microsystems GmbH, Wetzlar, Germany). Lasers at 488

178

nm and 543 nm were used to excite the FITC and Cy3 fluorophores, respectively. Image

179

segmentation and analysis were performed with Image Pro Plus 6.0 software (Media Cybernetics,

180

Rockville, MD, USA). Outputs included the distribution area of AOB_Beta (AAOB_Beta), the

181

distribution area of AOB_eu (AAOB_eu), the distribution area of NOB_Nba (ANOB_Nba), and the

182

distribution area of DEN_Tb (ADEN_Tb). Two particles were examined for each composting

183

sample.

184

The particles were also observed by an optical microscope (DM 2500, Leica Microsystems

185

GmbH, Wetzlar, Germany). The particle area (Apart) was quantified using Image Pro Plus 6.0

186

software. The area ratios of the microbes to the particles were calculated according to

187

AAOB_Beta/Apart, AAOB_eu/Apart, ANOB_Nba/Apart, and ADEN_Tb/Apart.

188

Statistical analysis and data processing. One-way analysis of variance and least

189

significant difference tests were performed using SPSS version 15.0 software (SPSS, Inc.,

10 Environment ACS Paragon Plus

Page 10 of 40

Page 11 of 40

Environmental Science & Technology

190

Chicago, IL, USA), with the significance level set at p < 0.05, to determine the significant

191

difference in the physicochemical parameters between experiments A and B.

192

The correlation between the variables at the pile and particle scales was quantified by

193

canonical correspondence analysis (CCA) using Canoco 5.0 software (Microcomputer Power,

194

USA) to calculate the following: (1) the correlation between the functional genes of AOB_Beta,

195

NOB_Nba, DEN_Tb, nirS + nirK, and nosZ and the concentration of NH4+, NO3–, NO2–, and N2O;

196

(2) the correlation between the distribution area of the microbes (i.e., AAOB_Beta/Apart,

197

AAOB_eu/Apart, ANOB_Nba/Apart, and ADEN_Tb/Apart) and the concentration of NH4+, NO3–, NO2–, and

198

N2O; and (3) the Pearson correlation between N2O emission and other variables (i.e., O2, NH3,

199

NH4+, NO2–, NO3–, AOB_Beta abundance, NOB_Nba abundance, DEN_Tb abundance, nirS +

200

nirK abundance nosZ abundance, AAOB_Beta/Apart, AAOB_eu/Apart, ANOB_Nba/Apart, and ADEN_Tb/Apart).

201

RESULTS AND DISCUSSION

202

Evolution of temperature, O2 concentration, NH3 emission, and N2O emission at the

203

pile scale. As shown in Figure 1a, both the composting temperature of experiments A and B

204

followed the classic four-phase pattern, i.e., the mesophilic (days 0–4), thermophilic (days 4–6),

205

cooling (days 6–12), and maturation (days 12–16) phases. The ambient temperature was

206

maintained at 23°C. As shown in Figure 1b, the O2 concentration declined dramatically during

207

the mesophilic phase because of vigorous metabolism. During the thermophilic phase, the

208

composting temperature was higher than 50°C, which could kill most pathogens and thereby

209

meet sanitary standards.6 Meanwhile, the O2 concentration was at a low level that provided ideal

210

conditions for anaerobic denitrification. The trend of NH3 emission is similar to that of the

211

composting temperature. The volatilization of NH3 was enhanced during the mesophilic phase,

212

and NH3 emission peaked at around 1150 ppmv during the thermophilic phase. In the later

11 Environment ACS Paragon Plus

Environmental Science & Technology

213

phases, NH3 emission gradually decreased. Briefly, NH3 was formed throughout the mesophilic

214

and thermophilic phases, which is consistent with the previous research on household waste

215

composting.39 N2O emission increased in the mesophilic, thermophilic, and cooling phases,

216

respectively. N2O emission started around the middle stage of the composting period when NH3

217

emission and the composting temperature began to decline, which agrees with the observations

218

of the study of swine manure composting.40 There was N2O emission during the cooling phase

219

even though the O2 concentration in the reactor was high, suggesting that N2O emission was not

220

only related to the aerobic/anaerobic conditions in the macro-environment but also the

221

aerobic/anaerobic conditions in the composting particles.

12 Environment ACS Paragon Plus

Page 12 of 40

Page 13 of 40

Environmental Science & Technology

222 223

(a)

224 225

(b)

226

Figure 1. Evolution of temperature and gas concentration during pig manure–wheat straw

227

aerobic composting: (a) composting temperature and ambient temperature; (b) oxygen (O2)

228

concentration, ammonia (NH3) emission, and nitrous oxide (N2O) emission.

229

Note: the values of composting temperature, ambient temperature, and O2 are the means of two

13 Environment ACS Paragon Plus

Environmental Science & Technology

230

replicates; the error bars of NH3 and N2O are the standard deviations of six replicates.

231

Evolution of NH2OH, NH4+, NO2–, and NO3– at the pile scale. The analytical results

232

showed that NH2OH was not detected in any of the composting samples. Wunderlin et al.14

233

added NH2OH to sludge reactors and examined the resulting variations of nitrogen-containing

234

ions. Their results showed that most of the NH2OH was converted into NO3–, which might be the

235

reason for the lack of detection of NH2OH in this study. The evolution of NH4+, NO2–, and NO3–

236

during pig manure–wheat straw aerobic composting is illustrated in Figure 2. The increase of the

237

composting temperature during the mesophilic phase (days 0–4) accelerated the transfer from

238

NH4+ to NH3 and the volatilization of NH3. As a result, the NH4+ concentration decreased

239

significantly, which agrees with the research on pig slurry–wheat straw composting.41 The

240

correlation between the composting temperature and NH3 emission might occur because the

241

equilibrium between NH4+ and NH3 was proportional to the Henry constant that is depending on

242

temperature.41 The decline of NO2– during the thermophilic phase (days 4–6) and the late phase

243

(days 8–16) implied that NO2– oxidation was promoted, because an increase of NO3– was

244

observed between days 4 and 16. The variation of NO3– observed in this study is consistent with

245

the study of cattle and sheep manure composting, where an increase in the NO3– concentration

246

occurred during the maturation phase, when the composting temperature was close to the

247

ambient temperature.42

14 Environment ACS Paragon Plus

Page 14 of 40

Page 15 of 40

Environmental Science & Technology

248 249

Figure 2. Evolution of the concentration of ammonium (NH4+), nitrite (NO2–), and nitrate (NO3–)

250

during pig manure–wheat straw aerobic composting.

251

Note: the error bars are the standard deviations of six replicates.

252

Evolution of bacterial abundance (AOB_Beta, NOB_Nba, DEN_Tb, nirS + nirK, and

253

nosZ) at the pile scale and related CCA analysis. The qPCR analysis showed that there was no

254

AOA in the composting samples, which is accordant with the study of cattle manure composting;

255

the reason might be a high level of O2 concentration or moisture content that would harm the

256

metabolism of AOA.8,43,44 Moreover, NOB_Nsp, DEN_AT, and DEN_P were not detected.

257

Therefore, NOB_Nba and DEN_Tb were the major NOB and DEN, respectively, during pig

258

manure–wheat straw composting. The evolution of AOB_Beta, NOB_Nba, DEN_Tb, nirS + nirK,

259

and nosZ and the related CCA analysis is depicted in Figure 3. The concurrence of AOB_Beta,

260

NOB_Nba, and DEN_Tb suggested that both the aerobic nitrification of AOB and NOB and the

261

anaerobic denitrification of DEN existed in the composting piles. The magnitude of AOB_Beta

262

approximates the results found in cattle and pig manure composting, where the AOB_Beta

263

abundance ranged between 105–106 copies g–1.43 The NOB_Nba was three orders of magnitude

15 Environment ACS Paragon Plus

Environmental Science & Technology

264

lower than AOB_Beta. The bacterial abundance of NOB_Nba dropped between days 4 and 8 and

265

soon rebounded, probably because the high temperature led to the inactivation of NOB_Nba

266

followed by recovery of the bacteria with decreasing temperature. The trends of nirS, nirK, and

267

nosZ agree with the results of cattle manure composting, where nirS generally increased and

268

nirK and nosZ decreased first and later increased.5 As shown in Figure 3b, the first two CCA

269

axes explained 48.9% of the cumulative variance of the correlation between the bacterial

270

abundance (AOB_Beta, NOB_Nba, DEN_Tb, nirS + nirK, and nosZ) and the variables (NH4+,

271

NO3–, NO2–, and N2O) with p = 0.002, which demonstrated that these variables could

272

significantly explain the variation of bacterial abundance.45 The distance between AOB_Beta and

273

NOB_Nba is shorter than that between AOB_Beta and others, indicating a smaller difference

274

between AOB_Beta and NOB_Nba abundance. This might occur because both AOB_Beta and

275

NOB_Nba are nitrifiers. The abundance of AOB_Beta was related to NH4+ levels because

276

AOB_Beta are the important bacteria in NH4+ oxidation. The high correlation between NOB_Nba

277

and NO2– was reasonable because NOB_Nba consumed NO2– in the process of nitrification.

278

Recall that nirS + nirK is produced by AOB_eu and/or DEN while nosZ is associated with DEN,

279

so the fact that both nirS + nirK and nosZ are closely related to N2O emission (Figure 3b)

280

suggested that the denitrification accomplished by AOB_eu and/or DEN was the key pathway of

281

N2O emission. As shown in Figure 3b, the distribution of composting samples varied according

282

to different composting time. The composting samples of the early phase (days 0–6) featured

283

higher NH4+ concentration and AOB_Beta abundance, which demonstrated strong nitrification

284

took place at that time. The samples of the late phase (days 12–16) showed higher NO3–

285

concentration, N2O emission, and abundance of nirS + nirK and nosZ suggested that both

286

nitrification and denitrification were robust during the late phase of composting, and AOB_eu

16 Environment ACS Paragon Plus

Page 16 of 40

Page 17 of 40

287

Environmental Science & Technology

and/or DEN were the dominant bacteria in denitrification.

288 289

(a) qPCR analysis

290 291

(b) CCA analysis

292

Figure 3. Evolution of bacterial abundance during pig manure–wheat straw aerobic composting:

17 Environment ACS Paragon Plus

Environmental Science & Technology

Page 18 of 40

293

(a) real-time quantitative polymerase chain reaction (qPCR) analysis; (b) related canonical

294

correspondence analysis (CCA).

295

Note:

296

denitrificans; nirS + nirK, nitrite reductase; nosZ, nitrous oxide reductase; NH4+, ammonium;

297

NO2–, nitrite; NO3–, nitrate; N2O, nitrous oxide; the error bars in Figure 3a are the standard

298

deviations of six replicates; the sample name in Figure 3b is “composting days-experiment A or

299

B-the serial number of the replicates”.

AOB_Beta,

Betaproteobacteria;

NOB_Nba, Nitrobacter;

DEN_Tb,

Thiobacillus

300

Dynamic distribution of AOB_Beta, AOB_eu, NOB_Nba, and DEN_Tb in composting

301

particles. In the FISH–CLSM images (Figure 4), the dynamic distribution of AOB_Beta,

302

AOB_eu, NOB_Nba, and DEN_Tb in composting particles was qualitatively evaluated. As shown

303

in Figure 4a, AOB_Beta was sporadically distributed on the outer area of the composting

304

particles. Earlier research23 showed the possibility that the outer area of composting particles was

305

aerobic, so it was possible that AOB_Beta therefore conducted aerobic reactions, such as

306

ammonia oxidation. The location of AOB_Beta is similar to the distribution of AOB_Beta in

307

biomass granules,46 but the area of AOB_Beta in composting particles is larger than that in

308

biomass granules. The reason might be either that the microbial communities differentiated

309

between different materials, or that the previous research used a NSO190 probe to stain

310

AOB_Beta, while the present study used a combined NSO190–NSO1225 probe that was more

311

comprehensive. Furthermore, the distribution of AOB_Beta was discontinuous (Figure 4a),

312

which is accordant with the previous CLSM analysis of biofilm.47 However, AOB_eu, which also

313

belongs to the AOB category, was located in the interior of the composting particles (Figure 4a).

314

The different location between AOB_Beta and AOB_eu might be attributed to the spatial

315

difference of NH4+ and NO2– in the composting particles and different functions of the bacteria.

18 Environment ACS Paragon Plus

Page 19 of 40

Environmental Science & Technology

316

Matsumoto et al.48 and He et al.49 determined the profiles of NH4+ and NO2– in sludge and

317

biomass granules using microelectrodes. They found that the NH4+ concentration decreased from

318

the granular surface inwards, while the NO2– concentration showed a slight increase. From this it

319

might be inferred that the NH4+ concentration was higher in the outer area of the composting

320

particles, which facilitated the growth of AOB_Beta in the outer area. In contrast, the NO2–

321

concentration was higher in the interior of the particles, which provided hospitable conditions for

322

AOB_eu. However, this conjecture should be further investigated using microelectrodes or other

323

techniques. As shown in Figure 4b, NOB_Nba were internally attached to AOB_Beta, which is

324

consistent with the biofilm structure seen during wastewater treatment.50 NOB_Nba were

325

adjacent to AOB_Beta, probably because NOB_Nba needed to take in the NO2– generated from

326

AOB_Beta.46,51 Moreover, as NOB_Nba and AOB_Beta competed for the available O2,

327

NOB_Nba lost out and was forced to attach to the AOB_Beta internally.52 As shown in Figure 4c,

328

DEN_Tb was concentrated in the interior of the composting particles, which suggested the

329

denitrification due to DEN was an anaerobic reaction. The coexistence of AOB_Beta and

330

DEN_Tb in the composting particles illustrates a link between the mechanisms of aerobic

331

nitrification and anaerobic denitrification during pig manure–wheat straw aerobic composting.

19 Environment ACS Paragon Plus

Environmental Science & Technology

332 333

(a) AOB_Beta (red) and AOB_eu (green)

20 Environment ACS Paragon Plus

Page 20 of 40

Page 21 of 40

Environmental Science & Technology

334 335

(b) AOB_Beta (red) and NOB_Nba (green)

21 Environment ACS Paragon Plus

Environmental Science & Technology

336 337

(c) AOB_Beta (red) and DEN_Tb (green)

22 Environment ACS Paragon Plus

Page 22 of 40

Page 23 of 40

Environmental Science & Technology

338

Figure 4. Dynamic distribution of nitrifiers and denitrifiers in composting particles: (a) Betaproteobacteria (AOB_Beta) and

339

Nitrosomonas europea/Nitrosomonas eutropha (AOB_eu); (b) AOB_Beta and Nitrobacter (NOB_Nba); (c) AOB_Beta (red) and

340

Thiobacillus denitrificans (DEN_Tb).

341

Note: the scale of the images is given on the first image.

23 Environment ACS Paragon Plus

Environmental Science & Technology

342

Evolution of the distribution area of AOB_Beta, AOB_eu, NOB_Nba, and DEN_Tb in

343

composting particles and related CCA analysis. As mentioned above in Materials and

344

Methods, we represent the distribution of the various types of bacteria by calculating the ratio of

345

the area occupied by the bacteria in question to the area of the surface of the particle on which

346

the bacteria has grown. As shown in Figure 5a, AAOB_Beta/Apart is the highest, which agrees with

347

the fact that AOB_Beta had the highest magnitude in the qPCR results above (Figure 3a). The

348

variation of AAOB_eu/Apart in Figure 5a is very similar to the variation of N2O emission shown in

349

Figure 1b, which suggested that AOB_eu was associated with N2O emission. Also, the evolution

350

of ANOB_Nba/Apart shown here agrees with that of NOB_Nba shown in Figure 3a. As shown in

351

Figure 5b, the first two CCA axes explained 50.3% of the cumulative variance of the correlation

352

between the bacterial distribution area (AAOB_Beta/Apart, ANOB_Nba/Apart, AAOB_eu/Apart, and

353

ADEN_Tb/Apart) and the variables (NH4+, NO3–, NO2–, and N2O) with p = 0.002. The distance

354

between AAOB_Beta/Apart and AAOB_eu/Apart was significant although both AOB_Beta and AOB_eu

355

belong to AOB, which quantified the difference in their growth habits and reaction mechanisms.

356

The AAOB_Beta/Apart distribution closely followed that of NH4+ variation while ANOB_Nba/Apart

357

correlated well with variations of NO2– and NO3–, which is consistent to the CCA results of the

358

qPCR analysis (Figure 3b). In addition, the distribution of AAOB_eu/Apart correlated strongly with

359

N2O variation, again suggesting that AOB_eu was the most important pathway of N2O emission.

24 Environment ACS Paragon Plus

Page 24 of 40

Page 25 of 40

Environmental Science & Technology

360 361

(a) FISH–CLSM analysis

362 363

(b) CCA analysis

25 Environment ACS Paragon Plus

Environmental Science & Technology

364

Figure 5. Evolution of the distribution area of Betaproteobacteria (AOB_Beta), Nitrosomonas

365

europea/Nitrosomonas eutropha (AOB_eu), Nitrobacter (NOB_Nba), and Thiobacillus

366

denitrificans (DEN_Tb) in composting particles: (a) fluorescent in situ hybridization–confocal

367

laser scanning microscopy (FISH–CLSM) analysis; (b) related canonical correspondence

368

analysis (CCA).

369

Note: AAOB_Beta, AAOB_eu, ANOB_Nba, and ADEN_Tb represent the distribution area of AOB_Beta,

370

AOB_eu, NOB_Nba, and DEN_Tb, respectively; Apart, the particle area; NH4+, ammonium; NO2–,

371

nitrite; NO3–, nitrate; N2O, nitrous oxide; the error bars in Figure 5a are the standard deviations

372

of four replicates.

373

Regarding the contribution of nitrification, nitrifier denitrification, and anaerobic

374

denitrification to N2O production, the results of multiple linear regression showed N2O = –

375

7.005AAOB_Beta + 544.752AAOB_eu + 159.452ADEN_Tb with a coefficient of determination (R2) of

376

0.87, where the p-values of AAOB_Beta, AAOB_eu, and ADEN_Tb were 0.771, 0.001, and 0.188. This

377

statistics implied that the correlation between nitrifier denitrification and N2O emission was

378

extremely significant (p < 0.01). Although AOB_Beta had the highest abundance (Figure 3a) and

379

distribution area (Figure 5a), the correlation between AAOB_Beta and N2O emission was

380

insignificant. It may be that, as in the case of Wunderlin et al.,14 most of the NH2OH was

381

converted into NO3– but the contribution of nitrification to N2O emission could be relatively low.

382

As shown in Table S3, the Pearson correlation analysis between N2O emission and individual

383

variables (O2, NH3, NH4+, NO2–, NO3–, AOB_Beta abundance, NOB_Nba abundance, DEN_Tb

384

abundance, nirS + nirK abundance nosZ abundance, AAOB_Beta/Apart, AAOB_eu/Apart, ANOB_Nba/Apart,

385

and ADEN_Tb/Apart) also showed that there was a strong positive linear correlation between

386

AAOB_eu/Apart and N2O emission (N2O emission = 1161.7AAOB_eu/Apart, R2 = 0.84). Given the high

26 Environment ACS Paragon Plus

Page 26 of 40

Page 27 of 40

Environmental Science & Technology

387

correlation between AAOB_eu/Apart and N2O emission observed by multiple linear regression and

388

Pearson correlation analysis, the AOB_eu-driven denitrification, i.e., nitrifier denitrification,

389

appears to be the main pathway of N2O emission. Therefore, during the entire aerobic

390

composting process, one variety of AOB participated in aerobic nitrification while another

391

variety performed nitrifier denitrification. Besides, as shown in Table S3, a positive correlation

392

was found between N2O emission and nosZ, although the correlation was expected to be negative

393

because of anaerobic denitrification. Some previous studies of wetland soils and agricultural

394

soils also found this discrepancy.53,54 The reason for this apparent anomaly is that nitrification by

395

AOB_Beta and nitrifier denitrification by AOB_eu contributed more to N2O emission than did

396

anaerobic denitrification, and since the nosZ gene has not yet been found to be in AOB_Beta and

397

AOB_eu;55-57 the N2O generated during nitrification and nitrifier denitrification must have

398

stimulated growth of DEN_Tb, thereby increasing the concentration of nosZ and generating a

399

positive relationship between N2O emission and nosZ.

400

This study presented the multiple linear regression of N2O emission and demonstrated the

401

important role of AOB_eu to N2O emission. More advanced technology, such as isotope labeling

402

and microelectrodes,14,58,59 could be employed in future studies to measure the contribution of

403

nitrification, nitrifier denitrification, and anaerobic denitrification to N2O emission and to better

404

identify the source of N2O emission during composting.

405

N2O emission mechanisms of pig manure–wheat straw aerobic composting based on

406

multivariate characterization at multi-scale. Based on the multivariate characterization results

407

at both the pile scale and the particle scale (Figures 1b, 2, 3a, and 5a), the nitrification and N2O

408

emission mechanisms of pig manure–wheat straw aerobic composting are summarized in Table 2.

409

Dividing the observation period into the usual three phases (mesophilic, thermophilic, and

27 Environment ACS Paragon Plus

Environmental Science & Technology

410

cooling), we observed growth of N2O emission in all three periods. Comparing the N2O growth

411

with the variation in the concentrations and abundance of ions and microbes across these three

412

periods enables us to draw some conclusions about the nitrification and N2O emission

413

mechanisms taking place during the three phases. The evolution of NO3– and AOB_Beta could be

414

used to reveal the status of nitrification during composting, because NO3– is the final product of

415

nitrification and AOB_Beta dominate the first step of nitrification (i.e., ammonia oxidation).

416

During the mesophilic phase, the concentration of NO3– was almost unchanged while AOB_Beta

417

kept increasing, suggesting that nitrification was not yet completed. During the thermophilic and

418

cooling phases, NO3– increased while AOB_Beta stayed constant at a high level, demonstrating a

419

period of active nitrification. According to previous studies,60,61 the increase in nirS + nirK could

420

be used to identify denitrification due to AOB_eu and/or DEN, because nirS + nirK is the nitrite

421

reductase derived from AOB_eu and/or DEN; since we observed increases in these enzymes in

422

all three phases, we could conclude that denitrification by one or both of these types of bacteria

423

was occurring. In contrast, an increase in nosZ seen in the third phase suggested denitrification

424

due to DEN, because nosZ is the nitrous oxide reductase generated by DEN. The evolution of

425

AAOB_eu/Apart indicated the status of nitrifier denitrification. As shown in Table 2, during the

426

mesophilic and thermophilic phases (days 0–6), the increasing nirS + nirK and AAOB_eu/Apart and

427

the constant nosZ implied that the denitrification of AOB_eu was more active than that of DEN

428

and, therefore, the N2O emission resulted mainly from the denitrification of AOB_eu, i.e.,

429

nitrifier denitrification. During the cooling phase (days 8–12), the rise in nosZ and the reduction

430

in AAOB_eu/Apart suggested a greater degree of anaerobic denitrification due to DEN, and so, over

431

the entire composting cycle, N2O emission was derived from a collaboration of AOB_eu and

432

DEN.

28 Environment ACS Paragon Plus

Page 28 of 40

Page 29 of 40

Environmental Science & Technology

433

Table 2. Nitrification and nitrous oxide (N2O) emission mechanisms of pig manure–wheat straw

434

aerobic composting based on multivariate characterization at multi-scale. Time

Phase

N2O

NO3–

AOB_Beta

nirS + nirK nosZ

AAOB_eu/Apart

Days 0–4

Mesophilic



-





-



Days 4–6

Thermophilic





-



-







-







Days 8–12 Cooling 435

Note: NO3–, nitrate; AOB_Beta, Betaproteobacteria; nirS + nirK, nitrite reductase; nosZ, nitrous

436

oxide reductase; AAOB_eu, the distribution area of Nitrosomonas europea/Nitrosomonas eutropha;

437

Apart, the particle area.

438

To summarize, during pig manure–wheat straw aerobic composting, AOB_Beta

439

predominated on the outer area of the composting particles, which was exposed to O2, during the

440

first phase of the composting period, which suggested that AOB_Beta conducted ammonia

441

oxidation and generated NO2– aerobically during this phase. In the next phase, NOB_Nba grew

442

and attached internally to AOB_Beta, utilizing NO2– to produce NO3–. In the interior of the

443

composting particles, the anaerobic conditions favored AOB_eu and DEN_Tb to perform

444

denitrification and result in N2O emission. During composting overall, nitrification was

445

consistently carried out; during the mesophilic and thermophilic phases, N2O emission was

446

derived from nitrifier denitrification due to AOB_eu; during the cooling phase, N2O emission

447

was from the denitrification due to both AOB_eu and DEN.

448

ENVIRONMENTAL RELEVANCE

449

According to the results of our multivariate characterization, more attention should be

450

focused on N2O emission during the late phase of composting, which originated from the

451

denitrification of both nitrifiers and denitrifiers. Perhaps more AOB_Beta and NOB_Nba could 29 Environment ACS Paragon Plus

Environmental Science & Technology

452

be added into composting mixtures to maximize the efficiency of nitrification and to retard

453

denitrification. Improving O2 permeation through the composting particles may also shrink the

454

denitrification area and therefore reduce N2O emission. The statistical significance of the multi-

455

scale quantitative analysis of the variables measured here raises hope that further development of

456

multi-scale models of N2O emission of pig manure–wheat straw aerobic composting may be of

457

some value. For example, the quantitative results for AAOB_Beta/Apart, AAOB_eu/Apart, ANOB_Nba/Apart,

458

and ADEN_Tb/Apart could help to build particle-scale N2O emission models that might be

459

appropriate to laboratory-scale composting systems. And the correlation observed between the

460

particle-scale and pile-scale variables might be useful for scaling the models up and

461

approximating the reality of factory-scale composting. These findings improve our

462

understanding of the mechanisms of N2O emission during manure-based aerobic composting and

463

may provide insights to facilitate decision making regarding responsible strategies for mitigating

464

climate change.

465

The applicability of the multivariate and multi-scale approaches to other waste composting

466

practices, e.g., composting of cattle and cow manures, that are known to lead to production of

467

significant N2O emissions, should be verified and modified in further studies. Furthermore, these

468

systematic methods could be useful for investigating the mechanisms of other greenhouse gas

469

emissions and odors, such as methane (CH4) and hydrogen sulfide (H2S), although the ions and

470

microbes involved may differ and may have to be specifically characterized. The potential

471

competition between and among these microbes should also be addressed. For instance, acetate,

472

sulfate, sulfite, and sulfide are the key ions involved in CH4 and H2S emissions; there is synergy

473

and competition that needs to be investigated in future work on nitrifiers, denitrifiers,

474

methanogens, methanotrophs, sulfate-reducing bacteria, and sulfur-oxidizing bacteria.7,62-64

30 Environment ACS Paragon Plus

Page 30 of 40

Page 31 of 40

Environmental Science & Technology

475 476

CORRESPONDING AUTHOR

477

Lujia Han

478

China Agricultural University (East Campus), Box 191, Beijing 100083, China

479

Fax: 86-10-6273-6778

480

Tel: 86-10-6273-6313

481

E-mail: [email protected]

482 483

ACKNOWLEDGMENTS

484

This work was financially supported by the China Agriculture Research System (CARS-36),

485

the Program for Changjiang Scholars and Innovative Research Team in University (IRT1293),

486

and the National Natural Science Foundation of China (31771684). We thank Christopher

487

Monroe for editing support.

488 489

SUPPORTING INFORMATION AVAILABLE

490

Section S1. Details of real-time quantitative polymerase chain reaction (qPCR).

491

Table S1. Primers used in qPCR analysis.

492

Section S2. Preparation of particle samples for fluorescent in situ hybridization–confocal laser

493

scanning microscopy (FISH–CLSM) analysis.

494

Figure S1. Micro-CT images and thickness of the composting particles.

495

Table S2. Fluorescent probes and hybridization conditions.

496

Section S3. Hybridization details.

497

Table S3. Pearson correlation analysis between nitrous oxide (N2O) emission and other variables.

31 Environment ACS Paragon Plus

Environmental Science & Technology

498 499

REFERENCES

500

1.

World Meteorological Organization. Greenhouse Gas Bulletin: The State of Greenhouse

501

Gases in the Atmosphere Using Global Observations through 2016; World Meteorological

502

Organization: Geneva, Switzerland, 2017.

503

2.

Szanto, G. L.; Hamelers, H. V. M.; Rulkens, W. H.; Veeken, A. H. M. NH3, N2O and CH4

504

emissions during passively aerated composting of straw-rich pig manure. Bioresour.

505

Technol. 2007, 98 (14), 2659−2670.

506

3.

Fukumoto, Y.; Inubushi, K. Effect of nitrite accumulation on nitrous oxide emission and

507

total nitrogen loss during swine manure composting. Soil Sci. Plant Nutr. 2009, 55 (3), 428–

508

434.

509

4.

Bong, C. P. C.; Li, Y. L.; Ho, W. S.; Lim, J. S.; Klemeš, J. J.; Towprayoon, S.; Ho, C. S.;

510

Lee, C. T. A review on the global warming potential of cleaner composting and mitigation

511

strategies. J. Clean. Prod. 2016, 146, 149–157.

512

5.

Li, S.; Song, L.; Gao, X.; Jin, Y.; Liu, S.; Shen, Q.; Zou, J. Microbial abundances predict

513

methane and nitrous oxide fluxes from a windrow composting system. Front. Microbiol.

514

2017, 8 (409), 1–15.

515

6.

516 517

Haug, R. T. The Practical Handbook of Compost Engineering; CRC Press: Boca Raton, FL, U.S.A., 1993.

7.

Sánchez, A.; Artola, A.; Font, X.; Gea, T.; Barrena, R.; Gabriel, D.; Sánchez-Monedero, M.;

518

Roig, A.; Cayuela, M.; Mondini, C. Greenhouse gas emissions from organic waste

519

composting. Environ. Chem. Lett. 2015, 1–16.

32 Environment ACS Paragon Plus

Page 32 of 40

Page 33 of 40

520

Environmental Science & Technology

8.

Maeda, K.; Toyoda, S.; Shimojima, R.; Osada, T.; Hanajima, D.; Morioka, R.; Yoshida, N.

521

Source of nitrous oxide emissions during the cow manure composting process as revealed by

522

isotopomer analysis of and amoA abundance in betaproteobacterial ammonia-oxidizing

523

bacteria. Appl. Environ. Microb. 2010, 76 (5), 1555–1562.

524 525

9.

Maeda, K.; Hanajima, D.; Toyoda, S.; Yoshida, N.; Morioka, R.; Osada, T. Microbiology of nitrogen cycle in animal manure compost. Microb. Biotechnol. 2011, 4 (6), 700–709.

526

10. Posmanik, R.; Gross, A.; Nejidat, A. Effect of high ammonia loads emitted from poultry-

527

manure digestion on nitrification activity and nitrifier-community structure in a compost

528

biofilter. Ecol. Eng. 2014, 62, 140–147.

529

11. Reyes, M.; Borrás, L.; Seco, A.; Ferrer, J. Identification and quantification of microbial

530

populations in activated sludge and anaerobic digestion processes. Environ. Technol. 2015,

531

36 (1), 45–53.

532

12. Bock, E.; Schmidt, I.; Stüven, R.; Zart, D. Nitrogen loss caused by denitrifying

533

Nitrosomonas cells using ammonium or hydrogen as electron donors and nitrite as electron

534

acceptor. Arch. Microbiol. 1995, 163 (1), 16–20.

535

13. Gieseke, A.; Bjerrum, L.; Wagner, M.; Amann, R. Structure and activity of multiple

536

nitrifying bacterial populations co-existing in a biofilm. Environ. Microbiol. 2003, 5 (5),

537

355–369.

538

14. Wunderlin, P.; Mohn, J.; Joss, A.; Emmenegger, L.; Siegrist, H. Mechanisms of N2O

539

production in biological wastewater treatment under nitrifying and denitrifying conditions.

540

Water Res. 2012, 46 (4), 1027–1037.

33 Environment ACS Paragon Plus

Environmental Science & Technology

541

15. Inoue, D.; Sawada, K.; Tsutsui, H.; Fujiwara, T. Identification of microbial populations

542

contributing to nitrification-associated nitrous oxide emission during cattle manure

543

composting process with forced aeration. J Mater. Cycles Waste. 2017, 1–8.

544

16. Zeman, C.; Depken, D.; Rich, M. Research on how the composting process impacts

545

greenhouse gas emissions and global warming. Compost Sci. Util. 2002, 10 (1), 72−86.

546

17. He, Y.; Inamori, Y.; Mizuochi, M.; Kong, H.; Iwami, N.; Sun, T. Measurements of N2O and

547 548 549 550 551 552 553

CH4 from the aerated composting of food waste. Sci. Total Environ. 2000, 254 (1), 65–74. 18. Czepiel, P.; Douglas, E.; Harriss, R.; Crill, P. Measurements of N2O from composted organic wastes. Environ. Sci. Technol. 1996, 30 (8), 2519–2525. 19. Hamelers, H. V. M. A mathematical model for composting kinetics. Ph.D. Dissertation, Wageningen University, Wageningen, Netherlands, 2001. 20. Hamelers, H. V. M. Modeling composting kinetics: A review of approaches. Rev. Environ. Sci. Biotechnol. 2004, 3 (4), 331–342.

554

21. D'Imporzano, G.; Crivelli, F.; Adani, F. Biological compost stability influences odor

555

molecules production measured by electronic nose during food-waste high-rate composting.

556

Sci. Total Environ. 2008, 402 (2–3), 278–284.

557

22. Maulini-Duran, C.; Puyuelo, B.; Artola, A.; Font, X.; Sánchez, A.; Gea, T. VOC emissions

558

from the composting of the organic fraction of municipal solid waste using standard and

559

advanced aeration strategies. J. Chem. Technol. Biotechnol. 2014, 89 (4), 579–586.

560

23. Ge, J.; Huang, G.; Yang, Z.; Huang, J.; Han, L. Characterization of the dynamic thickness of

561

the aerobic layer during pig manure aerobic composting by Fourier transform infrared

562

microspectroscopy. Environ. Sci. Technol. 2014, 48 (9), 5043–5050.

34 Environment ACS Paragon Plus

Page 34 of 40

Page 35 of 40

563 564 565 566

Environmental Science & Technology

24. Wang, K.; Li, W.; Li, X.; Ren, N. Spatial nitrifications of microbial processes during composting of swine, cow and chicken manure. Sci. Rep. 2015, 5, 14932. 25. Sun, Y.; Zou, J.; Li, J.; Lu, Y. Analysis of microorganism population in anaerobic granule with molecular bio-techniques. China Environl. Sci. 2006, 26 (2), 183–187.

567

26. Kim, T. G.; Yi, T.; Lee, E.-H.; Ryu, H. W.; Cho, K.-S. Characterization of a methane-

568

oxidizing biofilm using microarray, and confocal microscopy with image and geostatic

569

analyses. Appl. Microbiol. Biot. 2012, 95 (4), 1051–1059.

570

27. Ge, J.; Huang, G.; Huang, J.; Zeng, J.; Han, L. Modeling of oxygen uptake rate evolution in

571

pig manure–wheat straw aerobic composting process. Chem. Eng. J. 2015, 276, 29–36.

572

28. Ge, J.; Huang, G.; Huang, J.; Zeng, J.; Han, L. Mechanism and kinetics of organic matter

573

degradation based on particle structure variation during pig manure aerobic composting. J.

574

Hazard. Mater. 2015, 292, 19–26.

575

29. Ge, J.; Huang, G.; Huang, J.; Zeng, J.; Han, L. Particle-scale modeling of oxygen uptake rate

576

during pig manure–wheat straw composting: A new approach that considers surface

577

convection. Int. J. Heat Mass Trans. 2016, 97, 735–741.

578 579 580 581

30. U.S. Composting Council. Test methods for the examination of composting and compost; U.S. Composting Council: Bethesda, MD, U.S.A., 2002. 31. Frear, D.; Burrell, R. Spectrophotometric method for determining hydroxylamine reductase activity in higher plants. Anal. Chem. 1955, 27 (10), 1664–1665.

582

32. Féray, C.; Volat, B.; Degrange, V.; Clays-Josserand, A.; Montuelle, B. Assessment of three

583

methods for detection and quantification of nitrite-oxidizing bacteria and Nitrobacter in

584

freshwater sediments (MPN-PCR, MPN-Griess, immunofluorescence). Microb. Ecol. 1999,

585

37 (3), 208–217.

35 Environment ACS Paragon Plus

Environmental Science & Technology

586

33. Winkler, M. K.; Bassin, J. P.; Kleerebezem, R.; Sorokin, D. Y.; van Loosdrecht, M. C.

587

Unravelling the reasons for disproportion in the ratio of AOB and NOB in aerobic granular

588

sludge. Appl. Microbiol. Biot. 2012, 94 (6), 1657–1666.

589

34. Yoshida, N. 15N-depleted N2O as a product of nitrification. Nature 1988, 528–529.

590

35. Zou, G.; Papirio, S.; Lakaniemi, A. M.; Ahoranta, S. H.; Puhakka, J. A. High rate

591

autotrophic denitrification in fluidized-bed biofilm reactors. Chem. Eng. J. 2016, 284, 1287–

592

1294.

593

36. Szilveszter, S.; Ráduly, B.; Bucs, S.; Ábrahám, B.; Lányi, S.; Robescu, D. N. Activated

594

sludge floc characterization by confocal laser scanning microscopy. Environ. Eng. Manag. J.

595

2012, 11 (3), 669–674.

596

37. Szilveszter, S.; Ráduly, B.; Abraham, B.; Lányi, S. In situ imaging of biopolymers and

597

extracellular enzymes in activated sludge flocs of a municipal wastewater treatment plant. J.

598

Chem. Technol. Biotechnol. 2013, 88 (7), 1295–1304.

599

38. Kumar, T.; Majumdar, A.; Das, P.; Sarafis, V.; Ghose, M. Trypan blue as a fluorochrome for

600

confocal laser scanning microscopy of arbuscular mycorrhizae in three mangroves. Biotech.

601

Histochem. 2008, 83 (3–4), 153–159.

602

39. Jarvis, Å.; Sundberg, C.; Milenkovski, S.; Pell, M.; Smårs, S.; Lindgren, P. E.; Hallin, S.

603

Activity and composition of ammonia oxidizing bacterial communities and emission

604

dynamics of NH3 and N2O in a compost reactor treating organic household waste. J. Appl.

605

Microbiol. 2009, 106 (5), 1502–1511.

606

40. Fukumoto, Y.; Osada, T.; Dai, H.; Haga, K. Patterns and quantities of NH3 , N2O and CH4

607

emissions during swine manure composting without forced aeration––effect of compost pile

608

scale. Bioresour. Technol. 2003, 89 (2), 109–114.

36 Environment ACS Paragon Plus

Page 36 of 40

Page 37 of 40

609 610

Environmental Science & Technology

41. Oudart, D.; Robin, P.; Paillat, J.; Paul, E. Modelling nitrogen and carbon interactions in composting of animal manure in naturally aerated piles. Waste Manage. 2015, 46, 588–598.

611

42. López-Cano, I.; Roig, A.; Cayuela, M. L.; Alburquerque, J. A.; Sánchez-Monedero, M. A.

612

Biochar improves N cycling during composting of olive mill wastes and sheep manure.

613

Waste Manage. 2016, 49, 553–559.

614

43. Yamada, T.; Araki, S.; Ikeda-Ohtsubo, W.; Okamura, K.; Hiraishi, A.; Ueda, H.; Ueda, Y.;

615

Miyauchi, K.; Endo, G. Community structure and population dynamics of ammonia

616

oxidizers in composting processes of ammonia-rich livestock waste. Syst. Appl. Microbiol.

617

2013, 36 (5), 359–367.

618

44. Qin, W.; Meinhardt, K. A.; Moffett, J. W.; Devol, A. H.; Virginia Armbrust, E.; Ingalls, A.

619

E.; Stahl, D. A. Influence of oxygen availability on the activities of ammonia-oxidizing

620

archaea. Env. Microbiol. Rep. 2017, 9 (3), 250–256.

621 622

45. Gauch, H. G. Multivariate analysis in community ecology; Cambridge University Press: Cambridge, England, 1982.

623

46. Ali, M.; Rathnayake, R. M. L. D.; Zhang, L.; Ishii, S.; Kindaichi, T.; Satoh, H.; Toyoda, S.;

624

Yoshida, N.; Okabe, S. Source identification of nitrous oxide emission pathways from a

625

single-stage nitritation-anammox granular reactor. Water Res. 2016, 102, 147–157.

626

47. Wells, G.; Shi, Y.; Laureni, M.; Rosenthal, A.; Szivák, I.; Weissbrodt, D.; Joss, A.;

627

Buergmann, H.; Johnson, D.; Morgenroth, E. Comparing resistance, resilience, and stability

628

of replicate moving bed biofilm and suspended growth combined nitritation-anammox

629

reactors. Environ. Sci. Technol. 2017, 51 (9), 5108–5117.

630

48. Matsumoto, S.; Katoku, M.; Saeki, G.; Terada, A.; Aoi, Y.; Tsuneda, S.; Picioreanu, C.; Van

631

Loosdrecht, M. Microbial community structure in autotrophic nitrifying granules

37 Environment ACS Paragon Plus

Environmental Science & Technology

632

characterized by experimental and simulation analyses. Environ. Microbiol. 2010, 12 (1),

633

192–206.

634

49. He, Q.; Zhu, Y.; Fan, L.; Ai, H.; Huangfu, X.; Chen, M. Effects of C/N ratio on nitrous

635

oxide production from nitrification in a laboratory-scale biological aerated filter reactor.

636

Water Sci. Technol. 2017, 75 (6), 1270–1280.

637

50. Persson, F.; Sultana, R.; Suarez, M.; Hermansson, M.; Plaza, E.; Wilén, B.-M. Structure and

638

composition of biofilm communities in a moving bed biofilm reactor for nitritation–

639

anammox at low temperatures. Bioresour. Technol. 2014, 154, 267–273.

640

51. Mota, C. R.; Head, M. A.; Williams, J. C.; Eland, L.; Cheng, J. J.; de los Reyes, F. L.

641

Structural integrity affects nitrogen removal activity of granules in semi-continuous reactors.

642

Biodegradation 2014, 25 (6), 923–934.

643

52. Laanbroek, H. J.; Gerards, S. Competition for limiting amounts of oxygen between

644

Nitrosomonas europaea and Nitrobacter winogradskyi grown in mixed continuous cultures.

645

Arch. Microbiol. 1993, 159 (5), 453–459.

646

53. Ma, W.; Bedard-Haughn, A.; Siciliano, S.; Farrell, R. Relationship between nitrifier and

647

denitrifier community composition and abundance in predicting nitrous oxide emissions

648

from ephemeral wetland soils. Soil. Bio. Biochem. 2008, 40 (5), 1114–1123.

649

54. Dandie, C. E.; Burton, D. L.; Zebarth, B. J.; Henderson, S. L.; Trevors, J. T.; Goyer, C.

650

Changes in bacterial denitrifier community abundance over time in an agricultural field and

651

their relationship with denitrification activity. Appl. Environ. Microb. 2008, 74 (19), 5997–

652

6005.

38 Environment ACS Paragon Plus

Page 38 of 40

Page 39 of 40

Environmental Science & Technology

653

55. Kim, S.-W.; Miyahara, M.; Fushinobu, S.; Wakagi, T.; Shoun, H. Nitrous oxide emission

654

from nitrifying activated sludge dependent on denitrification by ammonia-oxidizing bacteria.

655

Bioresour. Technol. 2010, 101 (11), 3958–3963.

656

56. Schreiber, F.; Wunderlin, P.; Udert, K. M.; Wells, G. F. Nitric oxide and nitrous oxide

657

turnover in natural and engineered microbial communities: biological pathways, chemical

658

reactions, and novel technologies. Front. Microbiol. 2012, 3, 372.

659

57. Schmidt, I.; van Spanning, R. J.; Jetten, M. S. Denitrification and ammonia oxidation by

660

Nitrosomonas europaea wild-type, and NirK-and NorB-deficient mutants. Microbiology

661

2004, 150 (12), 4107–4114.

662

58. Wunderlin, P.; Lehmann, M.; Siegrist, H.; Tuzson, B.; Joss, A.; Emmenegger, L.; Mohn, J.

663

Isotope signatures of N2O in a mixed microbial population system: Constraints on N2O

664

producing pathways in wastewater treatment. Environ. Sci. Technol. 2013, 47 (3), 1339–

665

1348.

666

59. Rathnayake, R.; Song, Y.; Tumendelger, A.; Oshiki, M.; Ishii, S.; Satoh, H.; Toyoda, S.;

667

Yoshida, N.; Okabe, S. Source identification of nitrous oxide on autotrophic partial

668

nitrification in a granular sludge reactor. Water Res. 2013, 47 (19), 7078–7086.

669

60. Gao, K.; Zhao, J.; Ge, G.; Ding, X.; Wang, S.; Li, X.; Yu, Y. Effect of ammonium

670

concentration on N2O emission during autotrophic nitritation under oxygen-limited

671

conditions. Environ. Eng. Sci. 2017, 34 (2), 96–102.

672 673

61. Ni, B. J.; Yuan, Z. Recent advances in mathematical modeling of nitrous oxides emissions from wastewater treatment processes. Water Res. 2015, 87, 336–346.

39 Environment ACS Paragon Plus

Environmental Science & Technology

674

62. Malhautier, L.; Gracian, C.; Roux, J.-C.; Fanlo, J.-L.; Le Cloirec, P. Biological treatment

675

process of air loaded with an ammonia and hydrogen sulfide mixture. Chemosphere 2003,

676

50 (1), 145–153.

677

63. Galera, M. M.; Cho, E.; Kim, Y.; Farnazo, D.; Park, S.-j.; Oh, Y.-S.; Park, J. K.; Chung, W.-

678

J. Two-step pilot-scale biofilter system for the abatement of food waste composting

679

emission. J. Environ. Sci. Heal. A 2008, 43 (4), 412–418.

680

64. Chen, J.; Chen, T.; Gao, D.; Lei, M.; Zheng, G.; Liu, H.; Guo, S.; Cai, L. Reducing H2S

681

production by O2 feedback control during large-scale sewage sludge composting. Waste

682

Manage. 2011, 31 (1), 65–70.

683

40 Environment ACS Paragon Plus

Page 40 of 40