Subscriber access provided by University of South Dakota
Remediation and Control Technologies
Genome-centered metagenomics analysis reveals the symbiotic organisms possessing ability to cross-feed with anammox bacteria in anammox consortia Yunpeng Zhao, Shufeng Liu, Bo Jiang, Ying Feng, Tingting Zhu, Hu-Chun Tao, Xi Tang, and Sitong Liu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b02599 • Publication Date (Web): 10 Sep 2018 Downloaded from http://pubs.acs.org on September 11, 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
Genome-centered metagenomics analysis reveals the
2
symbiotic organisms possessing ability to cross-feed with
3
anammox bacteria in anammox consortia
4
Yunpeng Zhao 1, Shufeng Liu 1, Bo Jiang 1, Ying Feng 1, Tingting Zhu 2, Huchun Tao
5
3
, Xi Tang 1, Sitong Liu 1, 3 *
6
1
Department of Environmental Engineering, Peking University, Beijing 100871, China
7
2
State Environmental Protection Key Laboratory of Drinking Water Source Management and
8
Technology, Shenzhen Key Laboratory of Emerging Contaminants Detection & Control in Water
9
Environment, Shenzhen Academy of Environmental Sciences, Shenzhen 518001, China
10
3
11
518055, China
12
*Corresponding author: Sitong Liu
13
Address: College of Environmental Science and Engineering, Peking University, Yiheyuan Road,
14
No.5, Haidian District, Beijing 100871, China.
15
E-mail:
[email protected] 16
Tel/Fax: 0086-10-62754290.
School of Environment and Energy, Peking University Shenzhen Graduate School, Shenzhen
17
1
ACS Paragon Plus Environment
Environmental Science & Technology
Page 2 of 43
18
Abstract
19
Although using anammox communities for efficient wastewater treatment has
20
attracted much attention, the pure anammox bacteria are difficult to obtain, and the
21
potential roles of symbiotic bacteria in anammox performance are still elusive. Here,
22
we combined long-term reactor operation, genome-centered metagenomics,
23
community functional structure, and metabolic pathway reconstruction to reveal
24
multiple potential cross-feedings during anammox reactor start-up according to the 37
25
recovered metagenome-assembled genomes (MAGs). We found Armatimonadetes
26
and Proteobacteria could contribute the secondary metabolites molybdopterin cofactor
27
and folate for anammox bacteria to benefit their activity and growth.
28
Chloroflexi-affiliated bacteria encoded the function of biosynthesizing
29
exopolysaccharides for anammox consortium aggregation, based on the partial
30
nucleotide sugars produced by anammox bacteria. Chlorobi-affiliated bacteria had the
31
ability to degrade extracellular proteins produced by anammox bacteria to amino
32
acids to affect consortium aggregation. Additionally, the Chloroflexi-affiliated
33
bacteria harbored genes for a nitrite loop and could have a dual role in anammox
34
performance during reactor start-up. Cross-feeding in anammox community adds a
35
different dimension for understanding microbial interactions and emphasizes the
36
importance of symbiotic bacteria in the anammox process for wastewater treatment.
37
2
ACS Paragon Plus Environment
Page 3 of 43
Environmental Science & Technology
38
Introduction
39
Anaerobic ammonium oxidation (anammox) converts ammonium directly to
40
dinitrogen gas, with nitrite as an electron acceptor,1 and has been used for treating
41
ammonium-containing wastewater for its energy-efficient advantages.2,3 This process
42
has not been widely applied at full scale, and even then, it has been almost exclusively
43
for sidestream wastewater.3,4 It is mediated by a group of Planctomycetes bacteria
44
called anammox bacteria.1 Five genera of anammox bacteria have been discovered,
45
Kuenenia, Brocadia, Anammoxoglobus, Jettenia, and Scalindua,2,5 yet none of these
46
anammox bacteria have been purified so far.2,5 The microorganisms belonging to the
47
phylum Chloroflexi, Chlorobi, Proteobacteria, Acidobacteria, and Bacteroidetes
48
comprise 30–70% of the anammox consortia.2,4,6,7 The members of the anammox
49
consortia may share symbiotic relationships, which might be useful for anammox
50
metabolism, including growth factor supply or anammox metabolite scavenging.
51
Although this phenomenon has been observed for a long time, the potential roles of
52
these symbiotic bacteria in the consortia are still elusive.
53
Life in the microbial world is structured on the basis of complex interactions
54
among microbiomes.8,9 Cross-feeding, the production of metabolites that include
55
primary and secondary metabolites and can be used both by the producer and other
56
microbes in the community,10 is a type of microbial interaction that is beneficial for
57
both the producer and the receiver.11,12 This ubiquitous phenomenon exists in
58
microbial communities, especially in those growing under anoxic conditions and
59
nutrient and energy constraints.13 It is useful for extending the biochemical and 3
ACS Paragon Plus Environment
Environmental Science & Technology
Page 4 of 43
60
physiological capabilities of each participant.11 There are two types of cross-feeding
61
interactions:14 i) substrate cross-feeding, in which energy-rich organic substances are
62
degraded by one and then used by another bacterium; and ii) metabolic cross-feeding,
63
where the metabolites utilized by one type of microorganism are produced from
64
another type. Recently, metabolic cross-feeding has received increasing attention
65
because it involves metabolite exchange in natural and particularly in artificial
66
ecosystems, such as carbon source cross-feeding between autotrophic and
67
heterotrophic bacteria in the autotrophic denitrifying community in WWTPs.15
68
Microbial cross-feedings have been reported in the anammox community,
69
especially for carbon and energy sources. Initially, heterotrophs were detected in
70
anammox sludge, which could grow on the soluble microbial products of the
71
biofilm.16 Subsequent reports showed that the extracellular polysaccharides produced
72
by the anammox bacteria could be used as carbon source by certain heterotrophs of
73
the Phycisphaerae classes and Anaerolineaceae families.2,7 Recent studies showed
74
that less costly extracellular amino acids are used as carbon and energy sources,
75
especially by Chlorobi microorganisms.2 Specifically, these species can degrade the
76
extracellular proteins and peptides secreted by the anammox bacteria, and the
77
resultant amino acids are used either by themselves or by other microorganisms in the
78
community that cannot biosynthesize amino acids. Meanwhile, anammox bacteria can
79
also supply vitamins to other bacteria in the community.2 Nitrogen exchange between
80
anammox and denitrifiers is also considered as a cross-feeding. Nitrate, a product of
81
anammox reactions, can be used for the growth of species affiliated to the phyla 4
ACS Paragon Plus Environment
Page 5 of 43
Environmental Science & Technology
82
Chlorobi, Acidobacteria, and Omnitrophica.3 Simultaneously, the nitrite produced in
83
this process can be reused by anammox, which improves effluent quality.3 Although
84
metabolic cross-feeding has been identified in the anammox community, whether
85
more cross-feedings exist and how the abundance of species containing genes
86
potentially involved in cross-feeding remain unclear.
87
In this study, a sequencing batch reactor (SBR) was operated for 280 days. Sludge
88
samples were taken at different operational phases and analyzed using metagenomic
89
sequencing combined with binning. The adopted genome-based microbial ecology
90
method was similar to that used in a recent study.3 The Chloroflexi, Armatimonadetes,
91
Proteobacteria, Chlorobi bacteria possessing the functions to cross-feed with
92
anammox bacteria with respect to secondary metabolite syntrophism,
93
exopolysaccharide biosynthesis, or amino acid and nitrite exchange, as well as the
94
abundance variations of the genomes containing the genes for the above functions
95
during the start-up were investigated. Our results provide deeper insights into
96
microbial metabolic cross-feeding in anammox consortia during reactor start-up,
97
which will aid in understanding the roles of symbiotic bacteria in such consortia and
98
developing operational strategies of anammox processes.
99
Material and Methods
100
Start-up of the anammox reactor. A lab-scale SBR with a working volume of 3 L
101
anammox consortia was inoculated with an initial volatile suspended solids (VSS)
102
concentration of 0.1 g/L. It was taken from an SBR which was enriched with
103
Candidatus Brocadia sinica (abundance of 50 ± 5% in the whole community) and had 5
ACS Paragon Plus Environment
Environmental Science & Technology
Page 6 of 43
104
been preserved at 4°C for 90 days before inoculation.17 The SBR was fed with the
105
synthetic inorganic wastewater18 (presented in Supporting Information (SI) Text S1)
106
and continuously run for 280 days. Each reactor operational cycle included
107
continuous feeding (200 min), an anaerobic reaction period (20 h stepwise reduced to
108
2 h), settling (35 min), and withdrawal (5 min). The operational strategy was stepwise
109
increasing the nitrogen loading rate (NLR), which included increasing the nitrogen
110
concentration in the influent from 100 to 550 mg/L and shortening the hydraulic
111
retention time (HRT) from 4 days to 18 h. According to the NLR and reactor
112
performance, the reactor could be divided into five phases, and the detailed NLR of
113
each phase is presented in SI Table S1. The SBR was stirred at 150 rpm with a turbine
114
stirrer. Anaerobic conditions inside the reactor were maintained by continuously
115
purging N2/CO2 (95/5%) gas. The pH in the influent was kept in the range of 7.5 ± 0.3,
116
and the temperature was maintained at 37 ± 1 °C with a thermostatic water jacket.
117
Influent and effluent samples were collected every 2 days to measure the
118
concentrations of NH4+–N, NO2-–N, and NO3-–N according to standard methods.19
119
Five samples were harvested on the initial day (day 0) and at the end of each
120
operational phase (day 106, day 166, day 218, and day 280) for DNA extraction, and
121
these samples were labeled S1–S5, respectively.
122
DNA extraction and meta-genome sequencing. Total genomic DNA was
123
extracted using the FastDNA® spin kit for soil (MOBIO Laboratories Inc.) according
124
to the manufacturer’s protocol. DNA purity and concentration were determined using
125
a NanoDrop 2000 and the TBS-380 PicoGreen assay. DNA quality was checked by 6
ACS Paragon Plus Environment
Page 7 of 43
Environmental Science & Technology
126
1% agarose gel electrophoresis with ethidium bromide staining. The DNA samples
127
were then used for shotgun library construction and Illumina high-throughput
128
sequencing on the HiSeq 2500 platform at Majorbio Co. Ltd. (Shanghai, China) to
129
generate 150 bp paired-end reads (420 bp mean insert size) with seven runs. The raw
130
Illumina reads could be found on the National Center for Biotechnology Information
131
(NCBI) website under BioProject PRJNA450161.
132
Metagenome assembly, binning, and analysis. The raw metagenomic reads were
133
initially trimmed by stripping the adaptor sequences and ambiguous nucleotides using
134
SeqPrep version 1.1 based on default parameters,20 and the trimmed sequences were
135
quality filtered using Sickle version 1.33 based on a minimum quality score of 20 and
136
a minimum sequence length of 50 bp.21 Contigs and scaffolds were then assembled
137
individually for each sample using IDBA–UD with default parameters.22 The
138
scaffolds generated were binned into the draft genomes based on abundance and the
139
tetranucleotide frequency using MetaBAT version 0.32.5 with the sensitive model.23
140
CheckM version 1.0.7 was used to assess the completeness and contamination of the
141
recovered draft genomes based on the marker gene sets of 111 essential single-copy
142
genes.24 Draft genomes have been deposited to GenBank under the accession numbers
143
presented in SI Table S2.
144
The relative abundance of each MAG was calculated using the method previously
145
reported.2 First, the raw metagenome reads from each sample were mapped to all
146
contigs assembled to the draft genomes and contigs after assembly using bbmap
147
version 37.75 (https://sourceforge.net/projects/bbmap/) with the parameters ‘minid = 7
ACS Paragon Plus Environment
Environmental Science & Technology
148
0.95’ and ‘ambig = random’. Then, the number of reads that were mapped to each
149
MAG was normalized by the genome size of each MAG and the total reads that were
150
mapped to all contigs after assembly, which was expressed as reads per kilobase per
151
million mapped reads (RPKM). After that, the relative abundance was calculated
152
using the ratio between the RPKM value of each genomes and the sum of the RPKM
153
values of all genomes in each sample. Additionally, the average nucleotide identity
154
and comparison of draft and reference genomes were calculated using OrthoANI
155
version 0.93.125 and BRIG version 0.95.26
Page 8 of 43
156
The open reading frames (ORFs) of each recovered draft genome were annotated
157
using the metagenome implementation of Prodigal version 2.6.3 based on a minimum
158
nucleotide length of 60.27 ORFs were then queried against eggNOG database
159
(accessed on December 2016) and Kyoto Encyclopedia of Genes and Genomes
160
pathway database (KEGG, accessed on August 2017) with the blast e value of 1e-5.
161
The subcellular locations of ORFs were predicted using CELLO version 2.5.28
162
Construction of phylogenetic tree for the recovered draft genomes. Phylosift
163
version 1.0.1 was used to construct the phylogenetic tree.29 First, the taxonomic
164
affiliation of each MAG was determined using the Phylosift “all” command. At the
165
same time, the marker genes of 58 reference genomes that were related to the
166
recovered MAGs and deposited in GenBank were also identified and used to build a
167
phylogenetic tree. Then, the concatenated protein alignments of the recovered draft
168
genomes and the reference genomes were aligned using MAFFT version 7.31030 and
169
reconcatenated. Finally, a maximum likelihood phylogenetic tree was conducted using 8
ACS Paragon Plus Environment
Page 9 of 43
Environmental Science & Technology
170
RAxML version 8.2.11 with 100 bootstraps and the automatic protein model
171
assignment algorithm (PROTGAMMAAUTO).31
172
Functional structure analysis of the anammox community. The community
173
functional structures of different operational phases were established using the
174
functional gene abundance at the level of COG classes with principal component
175
analysis (PCA), as previously reported.32 In detail, the functional gene abundance of
176
COG functional class was calculated by adding the gene abundance of a certain COG
177
class in each MAG, in which the gene abundance was calculated by normalizing the
178
number of reads that were mapped to the gene sequence by the number of the 16S
179
rRNA gene sequence identified for each metagenomic sequencing data and gene
180
sequence length to prevent deviations caused by the slight difference in sequencing
181
depth, as reported.33 Then, the functional gene abundance matrix of COG functional
182
class was used for PCA in the R environment using the ggbiplot package.
183
Results
184
Varied nitrogen removal performance during reactor start-up. According to
185
the nitrogen loading rate (NLR) and nitrogen removal rate (NRR), the reactor
186
operational period could be divided into five operational phases during the reactor
187
start-up. In phase I (lag phase, days 1–77), the reactor showed extremely low nitrogen
188
removal performance, with a NRR of only 0.005 kg N/(m3 d) (Figure 1c) and nitrogen
189
removal efficiency (NRE) of only 10% (Figure 1d). In phase II (early propagation
190
phase, days 78–106), the nitrogen removal performance increased sharply, with NRR
191
and NRE increasing from 0.005 to 0.076 kg N/(m3 d) and from 9% to 80%, 9
ACS Paragon Plus Environment
Environmental Science & Technology
Page 10 of 43
192
respectively. The NRR increased further to 0.24 kg N/(m3 d) with a stable NRE of
193
80% in phase III (later propagation phase, days 107–167). To avoid nitrite
194
accumulation leading to the inhibition of bacterial activity in the reactor, the influent
195
nitrogen concentration decreased to 400 mg/L from day 142, and it increased again
196
from day 162. However, once NLR increased to 0.5 kg N/(m3 d) on days 173 and 183
197
as the HRT shortened and nitrogen concentration increased (Figure 1d), the reactor
198
performance deteriorated, and the high NRR and NRE could not be maintained in
199
phase IV (inhibition phase). At the same time, the pH difference between the effluent
200
and influent also decreased during this period (SI Figure S1). After culturing with
201
relatively low influent nitrogen concentrations, the nitrogen removal performance
202
increased constantly in phase V (stationary phase, days 221–280), with a NRR of 0.57
203
kg N/(m3 d) and a NRE of 80%. The NRR was higher than that obtained in previous
204
SBRs34, membrane bioreactors (MBR)7,35 and continuous stirred-tank reactors
205
(CSTR).36
206
The nitrogen removal ratios were also different between the five phases. Although
207
both the nitrite-to-ammonium consumption ratio (△NO2-–N/△NH4+–N) and the
208
nitrate-to-ammonium conversion ratio (△NO3-–N/△NH4+–N) varied considerably (SI
209
Figure S2 and Table S3) in phase I and the early stage of phase II, their values were
210
close to the stoichiometry of the anammox reaction towards the end of phase II, with
211
the average △NO2-–N/△NH4+–N and △NO3-–N/△NH4+–N of 1.20 ± 0.10 and 0.48 ±
212
0.23. In phase III, the average △NO2-–N/△NH4+–N and △NO3-–N/△NH4+–N were
213
1.20 ± 0.08 and 0.28 ± 0.06, respectively, which were in agreement with the 10
ACS Paragon Plus Environment
Page 11 of 43
Environmental Science & Technology
214
stoichiometry of anammox reaction during the reactor start-up. Afterwards, both
215
values decreased while the reactor performance deteriorated in phase IV, with average
216
△NO2-–N/△NH4+–N and △NO3-–N/△NH4+–N of 1.18 ± 0.08 and 0.26 ± 0.07,
217
respectively. In phase V, the average values of △NO2-–N/△NH4+–N and
218
△NO3-–N/△NH4+–N were 1.16 ± 0.08 and 0.23 ± 0.03, respectively, which agreed
219
with the previously reported values of 1.1–1.32 of △NO2-–N/△NH4+–N and 0.18–0.26
220
of △NO3-–N/△NH4+–N for the anammox reaction.37
221
Metagenomic sequencing, binning, and microbial succession. In total,
222
382,017,580 reads were generated, and 373,141,944 reads were obtained after read
223
quality control for five sludge samples harvested at the start of phase I and the end of
224
four phases (SI Table S4). The clean reads were used for assembly, and approximately
225
167,980–461,389 contigs with an N50 of 2,533–3,345 bp were generated. Contigs
226
were binned to the draft genomes, and 37 high-quality genomes were recovered with
227
high completeness and low contamination. Importantly, these 37 draft genomes
228
accounted for 55.11% of the original sequencing data from five samples on average
229
(SI Table S5), which was higher than that obtained (47.72%) in a previous study about
230
anammox granular binning.3 It indicated that these MAGs represented a major
231
fraction of the anammox community during reactor start-up.3
232
These draft genomes were affiliated to the phyla Planctomycetes, Armatimonadetes,
233
Chloroflexi, Proteobacteria, Chlorobi, Acidobacteria, Actinobacteria, Nitrospirae,
234
Microgenomates, and some candidate phyla, including Candidate division KSB1 and
235
Uhrbacteria (Table 1 and Figure 2). Specifically, AMX1, ATM1, CFX4, and CHB1, 11
ACS Paragon Plus Environment
Environmental Science & Technology
Page 12 of 43
236
which had high abundance, were similar to Candidatus Brocadia sinica JPN1,
237
Armatimonadetes bacterium OLB18, Anaerolineae bacterium UTCFX3, and
238
Ignavibacteriales bacterium UTCHB1 (SI Figure S3 and Table S6).2,3,38 Additionally,
239
GWF1, KSB1, UCB1, UCB2, and UCB3 had low identities and were genetically
240
distant from previously sequenced bacteria (Figure 2a and SI Table S6).
241
Although the relative abundances of CFX1, CFX2, and PRO1 were higher than
242
that of AMX1 for S1 (Figure 2b and SI Table S5), AMX1 became the predominant
243
MAG for S2, with an abundance increasing to 20.95% and then to 56.15% at the end
244
of the reactor start-up. Importantly, the second-most abundant MAG was CFX3 for S2,
245
whereas it was ATM1 for S3, S4, and S5. MAGs affiliated to Armatimonadetes,
246
Chloroflexi, Proteobacteria, and Chlorobi accounted for 54.8 ± 15.9% during S2–S5
247
(SI Table S7).
248
Variation of community functional structure. The PCA analyses using the
249
functional gene abundance of COG class showed the functional profile in the
250
anammox consortia (Figure 3), as the MAGs that had similar gene abundances at the
251
COG class level would colonize into a cluster. Intriguingly, there were marked
252
differences in the colonization-specific clusters between the five samples. In detail,
253
there were six groups for S1 within the complex MAG composition in every group
254
(Figure 3b). From S2 to S5, the number of clustering groups decreased, with relatively
255
smaller parts of MAGs clustering into a group. For example, partial MAGs of the
256
phyla Armatimonadetes, Chloroflexi, Proteobacteria, and Chlorobi were always in one
257
group that separated from AMX1 (Group II) and groups consisting of most other 12
ACS Paragon Plus Environment
Page 13 of 43
Environmental Science & Technology
258
MAGs (Figure 3c–f). However, the MAG composition for every group changed
259
constantly, while AMX1 (Group II) was separated from other MAGs alone in the last
260
four samples. These results suggest that the functional gene abundances of these
261
MAGs varied constantly and the symbiotic members had different functional gene
262
abundances compared with AMX1. For example, in addition to AMX1, CFX3 (Group
263
IV) was also separated from Group I (other MAGs) in S2, whereas ATM1 (Group IV)
264
was separated from Group I (other MAGs) in S3 and S4. In S5, groups of ATM1,
265
ATM2, PRO1, CFX4, and CHB1 (Group III and Group IV) were also separated from
266
Group I (other MAGs) and Group II (AMX1), which also indicated that these MAGs
267
had different functional gene abundances of COG class compared to AMX1 and other
268
MAGs.
269
To identify gene functions that were enriched in the above MAGs, we compared
270
the functional difference of each MAG at the level of COG class. Importantly, there
271
were distinctive metabolic functions for these MAGs (Data S1 and Figure S4). In
272
detail, CFX3 was enriched in COG class G (‘carbohydrate transport and metabolism’),
273
whereas CFX4 had more metabolic functions in COG class C (‘energy production and
274
conversion’). ATM1, ATM2, and PRO1 were enriched in COG class Q (‘secondary
275
metabolites biosynthesis, transport and catabolism’). However, AMX1 contained
276
fewer genes associated with such metabolic functions. These results suggest that the
277
different metabolic functions and different abundances in the five samples of these
278
MAGs made them colonize different clusters in different samples, which implied that
279
they might play different roles in different operational phases. 13
ACS Paragon Plus Environment
Environmental Science & Technology
Page 14 of 43
280
Metabolic pathway restructuration. To further identify the specific functional
281
potentials and metabolic pathways of the above MAGs, the ORFs were also blasted
282
against the KEGG database. The partial metabolic pathways of these MAGs with
283
more than 5% relative abundance in at least one sample are shown in Figure 4. They
284
included AMX1 (abundance of 6.72–56.15), ATM1 (abundance of 5.63–10.40%),
285
ATM2 (abundance of 5.23–25.01%), PRO1 (abundance of 0.25–6.89%), CFX3
286
(abundance of 0.25–10.42%), CFX4 (abundance of 1.58–12.52%), and CHB1
287
(abundance of 0.90–6.49%), which accounted for 83.12 ± 14.55% abundance during
288
the reactor start-up (Table S7). Interestingly, there were marked differences in
289
catabolic potential for different microbial genera: i) The MAGs affiliated to the phyla
290
Armatimonadetes and Proteobacteria were enriched in genes encoding components of
291
some secondary metabolite biosynthetic pathways, which are important for microbial
292
growth and activity. For example, ATM1 and ATM2 harbored genes (mogA and moeA)
293
for the biosynthesis of molybdopterin cofactor (MOCO). PRO1 encoded the complete
294
pathway for folate production. However, AMX1 lacked these metabolite functions. ii)
295
Although CFX3 harbored the complete pathway for biosynthesis of most
296
phosphorylated carbohydrates, including UDP-N-acetyl-D-glucosaminuronate
297
(UDP-GlcNAcA), GDP-mannose (GDP-Man), and GDP-D-rhamnose (GDP-Rha), it
298
lacked functions for other sugar nucleotide biosynthesis, such as for
299
UDP-N-acetyl-D-mannosamine (UDP-ManNAc) and CDP-glucose (CDP-Glc),
300
whereas AMX1 harbored the most abundant genes for biosynthesis of these molecules.
301
At the same time, CFX3 harbored more genes to facilitate macromolecule 14
ACS Paragon Plus Environment
Page 15 of 43
Environmental Science & Technology
302
polysaccharide biosynthesis, including protein-tyrosine phosphatase,
303
polyisoprenyl-phosphate glycosyltransferase, and polymerase, than other MAGs,
304
whereas AMX1 lacked them. Importantly, CFX3 also harbored an acidification
305
regulation gene for the macro-molecule polysaccharide biosynthesis (ldh), which were
306
missing in other bacteria, and contained more polysaccharide transporters in the inner
307
and outer membranes. iii) Nitrate reduction was redundant in CFX4, which had two
308
copies of nitrate reductase subunits genes (narGHIJ), whereas it harbored only one
309
copy of the nitrite reductase (NO-forming) nirS, which could cause nitrite
310
accumulation during denitrification. These observations were in agreement with the
311
comparative results of COG functional analysis. Furthermore, CHB1 had more
312
extracellular peptidases for protein hydrolysis, such as serine-type D-Ala-D-Ala
313
carboxypeptidase/endopeptidase, and D-alanyl-D-alanine carboxypeptidase among
314
specific peptidases, as well as other unspecific peptidases. At the same time, CHB1
315
possessed less complete pathways for amino acid biosynthesis, especially for many
316
amino acids with high biosynthetic cost, including tryptophan, phenylalanine, tyrosine,
317
methionine, and isoleucine, in line with a previous study.2
318
Based on gene abundance calculations in the whole metagenome, significant
319
discrepancies for the relative abundance of the above-mentioned genes in different
320
operational phases were observed (Figure 4). It was higher for moeA (for MOCO
321
production) from ATM1 and ATM2, and folA (for folate biosynthesis) from PRO1 in
322
the later propagation and stationary phases. The genes for polysaccharide
323
glycosyltransferase, polymerase, and transporters in CFX3 were abundant in the early 15
ACS Paragon Plus Environment
Environmental Science & Technology
Page 16 of 43
324
propagation phase, which decreased constantly till the end of the start-up. The
325
abundance of peptidase-encoding gene from CHB1 was relatively similar for S2, S3,
326
and S5, while it was lowest in the inhibition phase (S4). In contrast, narGHIJ of
327
CFX4 was highest in the inhibition phase.
328
Discussion
329
Metabolic cross-feedings in anammox consortia. Although the existence of
330
symbiotic bacteria in anammox consortia has been observed and the potential roles of
331
these bacteria in amino acid and vitamin exchange in anammox communities have
332
been reported recently,2,3 this study attempted to identify additional roles of these
333
bacteria in anammox performance with respect to cross-feedings within the anammox
334
consortia in terms of secondary metabolite symbiosis and polysaccharide biosynthesis.
335
Our study not only highlights cross-feedings in the anammox consortia but also
336
implies their potential roles in anammox activity, growth, and aggregation.
337
Here, we presented that multiple cross-feedings could exist in the anammox
338
consortia, and the abundances of recovered genomes containing genes potentially
339
involved in cross-feeding varied during the reactor start-up. Although the abundance
340
of AMX1 increased constantly from S1 to S5, microorganisms belonging to
341
Armatimonadetes, Chloroflexi, Proteobacteria, and Chlorobi showed relatively higher
342
percentages all along, which was in agreement with the results of previous studies.7,39
343
These metagenome-assembled genomes (MAGs) play an important ecological role in
344
the anammox community.2,3,7 Reports indicate functional complementarity of nitrogen
345
and amino acid metabolism between organisms in the microbial consortia, which 16
ACS Paragon Plus Environment
Page 17 of 43
Environmental Science & Technology
346
results in metabolite cross-feedings.2,3 We observed that these microorganisms had
347
distinct metabolic potentials with respect to secondary metabolite biosynthesis,
348
carbohydrate metabolism, extracellular protein and peptide degradation, and partial
349
denitrification, which were complementary to the metabolic potential of anammox
350
bacteria. Importantly, most of these functions were vital for anammox metabolism and
351
activity.2,3,7,40,41 Therefore, multiple metabolic cross-feedings among these bacteria
352
could exist in the anammox community (detailed discussion below). Simultaneously,
353
the relative abundances of the core genes encoding the above functions and the
354
microbial functional structures in the different operational phases during reactor
355
start-up were discrepant, which might lead to the varied cross-feeding strength. It has
356
been reported that microbial cross-feeding plays an important role in driving genetic
357
diversity and community stability and even affects the biochemical capacities of
358
ecological niches.11,12 Consequently, these cross-feedings could affect the metabolic
359
activity of anammox bacteria and the reactor performance during reactor start-up.
360
Cross-feedings of the secondary metabolites MOCO and folate. We propose
361
cross-feedings of secondary metabolites in the anammox consortia. Unlike primary
362
metabolites (amino acids, carbohydrates, and vitamins), secondary metabolites are not
363
directly required for microorganism growth.42 However, they are important for many
364
biological processes, particularly those involved in microbial interactions.42,43
365
Cross-feeding of secondary metabolites, including folate and biotin cofactor, has been
366
reported,10,44,45 as synthesizing them is energetically costly and requires complex
367
pathways. We observed that the anammox bacteria lacked the MOCO and folate 17
ACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 43
368
biosynthetic pathways, whereas the abundant Armatimonadetes and Proteobacteria
369
microorganisms in the consortia were enriched with these metabolic functions. At the
370
same time, the deposited GenBank genomes of anammox bacteria are all missing
371
complete pathways for MOCO and folate biosynthesis, including Candidatus Jettenia
372
caeni (BAFH00000000), Candidatus Brocadia sinica (BAFN00000000), and
373
Candidatus Brocadia fulgida (LAQJ00000000). However, MOCO is a cofactor for
374
many oxidoreductases, including the formate dehydrogenases (FDHs),40 which are
375
involved in the Wood–Ljungdahl pathway of anammox CO2 fixation.2 Folate is also a
376
necessary subunit for most compounds for anammox CO2 fixation, including
377
10-formyltetrahydrofolate, 5,10-methenyltetrahydrofolate, and
378
5,10-methylenetetrahydrofolate.41 These secondary metabolites are vital for anammox
379
growth and activity, as they affect carbon fixation or acetyl-coA production.41 The
380
anammox bacterium Candidatus K. stuttgartiensis also lacks some genes required for
381
folate production and has a mutualistic relationship with the symbiotic organisms.41
382
This is the same as the amino acid or nitrogen metabolism complementarity based on
383
microbial genomes.2,3 Therefore, MAGs belonging to the phyla Armatimonadetes and
384
Proteobacteria can produce these metabolites, which could be subsequently
385
transported to the anammox cell instead of direct production by the anammox for
386
energy saving. This process is a cross-feeding, considering the feedback process of
387
amino acid biosynthesis and secretion by anammox bacteria for Armatimonadetes and
388
Proteobacteria (detailed discussion below). As the cross-feedings of secondary
389
metabolites could lead to high bacterial activity,10 this process also could contribute to 18
ACS Paragon Plus Environment
Page 19 of 43
Environmental Science & Technology
390 391
the anammox activity and growth. Furthermore, the secondary-metabolite biosynthesis genes had higher abundance in
392
the propagation (S3) and stationary phases (S5), which was in line with the general
393
observation that secondary metabolites are often secreted during the stationary
394
phase.46 Their abundance was also positively correlated with the increasing NRR in
395
these two phases. The growth and activity of folate-deficient bacteria sharply
396
increases after co-culture with folate-producing bacteria.47 Therefore, the
397
cross-feedings of secondary metabolites between anammox and Armatimonadetes or
398
Proteobacteria bacteria could benefit for improving NRR via affecting anammox
399
growth and activity. This also emphasizes the importance of these bacteria for
400
anammox performance and the difficulty of anammox isolation.48 The exchange of
401
secondary metabolites possibly occurs via cell fusion,49 cell-cell connections,50 or
402
membrane vesicles.51 The detailed functional mechanism of secondary metabolite
403
transportation between bacteria is currently being investigated.52
404
Cross-feeding on macromolecular exopolysaccharide biosynthesis. We firstly
405
propose that CFX3 plays important roles in sticky macromolecular exopolysaccharide
406
biosynthesis owing to the cross-feeding of AMX1 and CFX3 in the anammox
407
consortia. Extracellular polysaccharides are usually assembled using hexose and/or its
408
derivative (sugar nucleotide) as repeating units.53,54 We observed that anammox
409
bacteria showed functional complementarity with another high-abundance MAG,
410
CFX3, in sugar nucleotide production, as AMX1 could biosynthesize UDP-ManNAc
411
and CDP-Glc, whereas CFX3 could produce UDP-GlcNAcA, GDP-Man, and 19
ACS Paragon Plus Environment
Environmental Science & Technology
Page 20 of 43
412
GDP-Rha. Meanwhile, compared to AMX1 and other microorganisms, CFX3
413
possessed more enzymes for modification, polymerization, and transportation of
414
carbohydrates, and some of them only existed in CFX3, such as protein-tyrosine
415
phosphatases and polyisoprenyl-phosphate glycosyltransferase. These enzymes are
416
used for sticky polysaccharide biosynthesis.55 Importantly, CFX3 encoded an enzyme
417
(ldh) for the biosynthesis of lactate, which is secreted into the extracellular media to
418
maintain acidity, a prerequisite for exopolysaccharide production.55 Hence, partial
419
nucleotide sugars could be biosynthesized by AMX1 and then transported inside
420
CFX3 for exopolysaccharide production. As the exopolysaccharides form a gel-like
421
cross-network, which is important for the formation of microbial aggregates,53–56 the
422
nucleotide sugar cross-feeding and secretion of macromolecule polysaccharides by
423
CFX3 could benefit for anammox bacteria aggregation. It also feeds back to AMX1
424
for the supply of amino acids (detailed discussion below). Thus, most macromolecule
425
exopolysaccharides could be biosynthesized and secreted by the cross-feedings of
426
anammox and Chloroflexi bacteria in the anammox consortia.
427
The anammox aggregation also could contribute to improvements in NRR and
428
NRE.7,57 Here, the relative abundances of core genes for exopolysaccharide
429
production in CFX3were highest in the early propagation phase (S2), which was
430
coincident with the increasing NRE in this phase, as it benefited for the recovery of
431
the anammox aggregation that had been hydrolyzed during storage.58 Accordingly, the
432
biosynthesis of macromolecular exopolysaccharides through cross-feeding between
433
anammox and Chloroflexi bacteria could be useful for the recovery of aggregation 20
ACS Paragon Plus Environment
Page 21 of 43
Environmental Science & Technology
434 435
and activity of the anammox consortia. Cross-feeding of amino acids and nitrite. Cross-feeding of amino acids has been
436
reported recently.2 The abundance of genomes containing the genes associated with
437
amino acid cross-feeding varied constantly during the reactor start-up. The Chlorobi
438
microorganisms harbored the maximum number of extracellular peptidases for
439
hydrolyzing proteins and peptides, while they also had the least amino acid
440
biosynthesis-related genes. This indicated that the Chlorobi microorganisms were the
441
dominant degraders of extracellular proteins, which were mainly produced by the
442
anammox bacteria.2 Importantly, this process released the amino acids sequestered in
443
the extracellular matrix for self-use and as carbon and energy sources for the other
444
organisms in the anammox consortia.2 This can also be a feedback mechanism for the
445
contributions of secondary metabolites and exopolysaccharides of Armatimonadetes,
446
Proteobacteria, and Chloroflexi organisms to anammox bacteria. Importantly, the
447
peptidase gene abundance of CHB1 changed constantly throughout S1-S5, and it was
448
lowest in the inhibition phase (S4). The anammox consortia can enhance its
449
aggregation capacity to cope with environmental stress.6 Hence, the low abundance of
450
extracellular peptidases potentially leads to less protein degradation, which improves
451
anammox aggregation by increasing the ratio of extracellular proteins and resisting
452
the high-nitrite phase.7 Therefore, we conclude that amino acid cross-feeding is a
453
self-protective strategy of the anammox consortia for coping with external stress.
454 455
Additionally, the Chloroflexi organisms harbored the genes for the nitrite loop, and they could play dual roles in nitrogen removal. The nitrate produced from the 21
ACS Paragon Plus Environment
Environmental Science & Technology
Page 22 of 43
456
anammox reaction can be reduced to nitrite by partial denitrifiers in the anammox
457
consortia. Subsequently, the nitrite can be reused by the anammox bacteria, and the
458
nitrite loop increases the nitrogen removal efficiency.2,3 This process can accelerate
459
the reactor start-up in the early period when the influent nitrite concentration is
460
limited. In contrast, nitrite cross-feeding could intensify the inhibition of anammox in
461
the later period when the influent nitrite concentration was high. Specifically, in the
462
propagation phases (phases II and III), the production and consumption of nitrite was
463
in dynamic balance, and the nitrogen removal rates were relatively higher. In the
464
inhibition phase (Phase IV), the balance was disrupted with the increase in abundance
465
of narGHIJ, from 0.01 in S1 to 0.15 in S4. The increasing amounts of nitrate could be
466
reduced to nitrite, whereas nitrite reductase (nirS) accounted for only 0.07 in S4,
467
which could lead to further inhibition of anammox when the influent nitrite
468
concentration was relatively higher. These data agreed with the low NRR in this phase.
469
Importantly, the average values of △NO3-–N/△NH4+–N and △NO2-–N/△NH4+–N of
470
phase IV were both lower than those of phases II and III, respectively, indicating
471
lower accumulation and removal rates of nitrate and nitrite, respectively.
472
Consequently, nitrite cross-feeding played a double role in nitrogen removal during
473
the reactor start-up.
474
Significance of this study. Microbial activity and aggregation capacity are
475
important for wastewater treatment.6,7,57 Here, we found that MAGs belonging to the
476
phyla Armatimonadetes and Proteobacteria harboring the secondary-metabolite
477
(MOCO and folate) biosynthesis ability to cross-feed with anammox and Chloroflexi 22
ACS Paragon Plus Environment
Page 23 of 43
Environmental Science & Technology
478
microorganisms encoding the function for a nitrite loop, which could affect anammox
479
activity. Meanwhile, exchanges of nucleotide sugars between anammox and
480
Chloroflexi bacteria and of amino acids between anammox and Chlorobi bacteria
481
might affect the aggregation capacity of the anammox consortia by producing
482
exopolysaccharides or degrading extracellular proteins. Thus, we emphasize the
483
important roles of MAGs belonging to the phyla Chloroflexi, Armatimonadetes,
484
Proteobacteria, and Chlorobi in the anammox consortia for nitrogen removal. The
485
microbial cross-feeding networks could affect reactor performance by regulating
486
microbial activity and aggregation, which indicates a potential strategy for improving
487
reactor performance by maintaining the balance and diversity of community structure
488
in different operational phases of the anammox reactor.
489
Furthermore, this is the first attempt at studying the microbial succession and
490
start-up process of an anammox reactor using a genome-centered metagenomic
491
analysis. Our proposed ecological interaction offers a deeper understanding of
492
anammox nitrogen removal from the viewpoint of bacterial cross-feeding during the
493
reactor start-up. The adopted genome-based microbial ecology method was similar to
494
that used in a recent study of the PN-anammox reactor.3 It should be noted that our
495
model was based on genomes and their varied abundances and should hence be
496
further validated using metatranscriptomics and metaproteomics2 or isolation of these
497
microorganisms.59 We expect that our results will provide a valuable reference for
498
such studies.
499 23
ACS Paragon Plus Environment
Environmental Science & Technology
Page 24 of 43
500
Acknowledgments
501
The financial support from Shenzhen Science and Technology Innovation Committee
502
(No. JSGG20160429162015597) should be highly appreciated. The authors are also
503
grateful to the National Natural Science Foundations of China (No. 51478006) for
504
financial support.
505
Supporting Information
506
Tables showing operational strategies, GenBank accession numbers, nitrogen
507
conversion ratios, sequencing and mapping statistics, average nucleotide identities,
508
partial MAGs abundance, genes definitions, and metabolites abbreviations; Details of
509
mineral medium composition; Figures illustrating the pH difference, nitrogen
510
conversion ratios, comparative genome analysis and gene number comparative
511
analysis of COG class (PDF).
512
File about the details of genes’ functional annotations (XLSX).
513
24
ACS Paragon Plus Environment
Page 25 of 43
Environmental Science & Technology
514
References
515
(1) Van Teeseling, M. C. F.; Mesman, R. J.; Kuru, E.; Espaillat, A.; Cava, F.; Brun,
516
Y. V.; Vannieuwenhze, M. S.; Kartal, B.; Van Niftrik, L. Anammox Planctomycetes
517
have a peptidoglycan cell wall. Nat. Commun. 2015, 6, 6878.
518
(2) Lawson, C. E.; Wu, S.; Bhattacharjee, A. S.; Hamilton, J. J.; McMahon, K. D.;
519
Goel, R.; Noguera, D. R. Metabolic network analysis reveals microbial community
520
interactions in anammox granules. Nat. Commun. 2017, 8, 15416.
521
(3) Speth, D. R.; Guerrero-Cruz, S.; Dutilh, B. E.; Jetten, M. S. M. Genome-based
522
microbial ecology of anammox granules in a full-scale wastewater treatment system.
523
Nat. Commun. 2016, 7, 11172.
524
(4) Bhattacharjee, A. S.; Wu, S.; Lawson, C. E.; Jetten, M. S. M.; Kapoor, V.;
525
Domingo, J. W. S.; McMahon, K. D.; Noguera, D. R.; Goel, R. Whole-community
526
metagenomics in two different anammox configurations: process performance and
527
community structure. Environm. Sci. Technol. 2017, 51 (8), 4317–4327.
528
(5) Kartal, B.; de Almeida, N. M.; Maalcke, W. J.; Op den Camp, H. J. M.; Jetten, M.
529
S. M.; Keltjens, J. T. How to make a living from anaerobic ammonium oxidation.
530
FEMS Microbiol. Rev. 2013, 37 (3), 428–461.
531
(6) Feng, Y.; Zhao, Y.; Guo, Y.; Liu, S. Microbial transcript and metabolome
532
analysis uncover discrepant metabolic pathways in autotrophic and mixotrophic
533
anammox consortia. Water Res. 2018, 128, 402–411.
534
(7) Zhao, Y.; Feng, Y.; Li, J.; Guo, Y.; Chen, L.; Liu, S. Insight into the aggregation
535
capacity of anammox consortia during reactor start-up. Environ. Sci. Technol. 2018, 25
ACS Paragon Plus Environment
Environmental Science & Technology
Page 26 of 43
536
52 (6), 402–411.
537
(8) Faust, K.; Raes, J. Microbial interactions: from networks to models. Nat. Rev.
538
Microbiol. 2012, 10 (8), 538–550.
539
(9) Alivisatos, A. P.; Blaser, M. J.; Brodie, E. L.; Chun, M.; Dangl, J. L.; Donohue, T.
540
J.; Dorrestein, P. C.; Gilbert, J. A.; Green, J. L.; Jansson, J. K.; Knight, R.; Maxon, M.
541
E.; McFall-Ngai, M. J.; Miller, J. F.; Pollard, K. S.; Ruby, E. G.; Taha, S. A. A
542
unified initiative to harness earth’s microbiomes. Science 2015, 350 (6260), 507–508.
543
(10) Seth, E. C.; Taga, M. E. Nutrient cross-feeding in the microbial world. Front.
544
Microbiol. 2014, 5, 350.
545
(11) Pande, S.; Merker, H.; Bohl, K.; Reichelt, M.; Schuster, S.; de Figueiredo, L. F.;
546
Kaleta, C.; Kost, C. Fitness and stability of obligate cross-feeding interactions that
547
emerge upon gene loss in bacteria. ISME J. 2014, 8, 953–962.
548
(12) Hoek, T. A.; Axelrod, K.; Biancalani, T.; Yurtsev, E. A.; Liu, J.; Gore, J.
549
Resource availability modulates the cooperative and competitive nature of a microbial
550
cross-feeding mutualism. PLoS Biol. 2016, 14 (8), e1002540.
551
(13) Morris, B. E. L.; Henneberger, R.; Huber, H.; Moissl-Eichinger, C. Microbial
552
syntrophy: interaction for the common good. FEMS Microbiol. Rev. 2013, 37 (3),
553
384–406.
554
(14) Ríos-Covián, D.; Ruas-Madiedo, P.; Margolles, A.; Gueimonde, M.; de los
555
Reyes-Gavilán, C. G.; Salazar, N. Intestinal short chain fatty acids and their link with
556
diet and human health. Front. Microbiol. 2016, 7, 185.
557
(15) Xing, W.; Li, J.; Cong, Y.; Gao, W.; Jia, Z.; Li, D. Identification of the 26
ACS Paragon Plus Environment
Page 27 of 43
Environmental Science & Technology
558
autotrophic denitrifying community in nitrate removal reactors by DNA-stable isotope
559
probing. Bioresour. Technol. 2017, 229, 134–142.
560
(16) Ni, B.; Ruscalleda, M.; Smets, B. F. Evaluation on the microbial interactions of
561
anaerobic ammonium oxidizers and heterotrophs in anammox biofilm. Water Res.
562
2012, 46 (15), 4645–4652.
563
(17) Guo, Y.; Liu, S.; Tang, X.; Wang, C.; Niu, Z.; Feng, Y. Insight into c-di-GMP
564
regulation in anammox aggregation in response to alternating feed loadings.
565
Environm. Sci. Technol. 2017, 51 (16), 9155–9164.
566
(18) van de Graaf, A. A.; de bruijn, P.; Robertson, L. A.; Jetten, M. S. M.; Kuenen, J.
567
G. Autotrophic growth of anaerobic ammonium-oxidizing micro-organisms in a
568
fluidized bed reactor. Microbiology 1996, 142 (8), 2187–2196.
569
(19) APHA. Standard Methods for the Examination of Water and Wastewater, 20th
570
ed.; American Public Health Association: Washington, DC, 1999.
571
(20) John, J. St. SeqPrep: tool for stripping adaptors and/or merging paired reads with
572
overlap into single reads. Available at https://github.com/jstjohn/SeqPrep. 2011.
573
(21) Joshi, N. A; Fass, J. N. A sliding-window, adaptive, quality-based trimming tool
574
for FastQ files. Available at https://github.com/najoshi/sickle. 2011.
575
(22) Peng, Y.; Leung, H. C. M.; Yiu, S. M.; Chin, F. Y. L. IDBA-UD: a de novo
576
assembler for single-cell and metagenomic sequencing data with highly uneven depth.
577
Bioinformatics 2012, 28 (11), 1420–1428.
578
(23) Kang, D. D.; Froula, J.; Egan, R.; Wang, Z. MetaBAT, an efficient tool for
579
accurately reconstructing single genomes from complex microbial communities. 27
ACS Paragon Plus Environment
Environmental Science & Technology
Page 28 of 43
580
PeerJ. 2015, 3, e1165.
581
(24) Parks, D. H.; Imelfort, M.; Skennerton, C. T.; Hugenholtz, P.; Tyson, G. W.
582
CheckM: assessing the quality of microbial genomes recovered from isolates, single
583
cells, and metagenomes. Genome Res. 2015, 25, 1043–1055.
584
(25) Lee, I.; Kim, Y. O.; Park, S. C.; Chun, J. OrthoANI: an improved algorithm and
585
software for calculating average nucleotide identity. Int. J. Syst. Evol. Microbiol. 2016,
586
66 (2), 1100–1103.
587
(26) Alikhan, N. F.; Petty, N. K.; Ben Zakour, N. L.; Beatson, S. A. BLAST ring
588
image generator (BRIG): simple prokaryote genome comparisons. BMC Genomics
589
2011, 12, 402.
590
(27) Hyatt, D.; Chen, G.; LoCascio, P. F.; Land, M. L.; Larimer, F. W.; Hauser, L. J.
591
Prodigal: prokaryotic gene recognition and translation initiation site identification.
592
BMC Bioinformatics 2010, 11 (1), 119.
593
(28) Yu, C.; Lin, C.; Hwang, J. Predicting subcellular localization of proteins for
594
Gram-negative bacteria by support vector machines based on n-peptide compositions.
595
Protein Sci. 2004, 13 (5), 1402–1406.
596
(29) Darling, A. E.; Jospin, G.; Lowe, E.; Matsen, F. A.; Bik, H. M.; Eisen, J. A.
597
PhyloSift: phylogenetic analysis of genomes and metagenomes. PeerJ 2014, 2, e243.
598
(30) Katoh, K.; Misawa, K.; Kuma, K.; Miyata, T. MAFFT: a novel method for rapid
599
multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 2002,
600
30 (14), 3059–3066.
601
(31) Stamatakis, A.; Hoover, P.; Rougemont, J. A rapid bootstrap algorithm for the 28
ACS Paragon Plus Environment
Page 29 of 43
Environmental Science & Technology
602
RAxML web servers. Syst. Biol. 2008, 57 (5), 758–771.
603
(32) Slaby, B. M.; Hackl, T.; Horn, H.; Bayer, K.; Hentschel, U. Metagenomic binning
604
of a marine sponge microbiome reveals unity in defense but metabolic specialization.
605
ISME J. 2017, 11 (11), 2465–2478.
606
(33) Ma, L.; Li, B.; Jiang, X.; Wang, Y.; Xia, Y.; Li, A.; Zhang, T. Catalogue of
607
antibiotic resistome and host-tracking in drinking water deciphered by a large scale
608
survey. Microbiome 2017, 5 (1), 154.
609
(34) Zhu, G. L.; Yan, J.; Hu, Y. Y. Anaerobic ammonium oxidation in polyvinyl
610
alcohol and sodium alginate immobilized biomass system: A potential tool to
611
maintain anammox biomass in application. Wat. Sci. Technol. 2014, 69 (4), 718–726.
612
(35) Xie, G. J.; Liu, T.; Cai, C.; Hu, S.; Yuan, Z. Achieving high-level nitrogen
613
removal in mainstream by coupling anammox with denitrifying anaerobic methane
614
oxidation in a membrane biofilm reactor. Water Res. 2018, 131, 196–204.
615
(36) Magrí, A.; Vanotti, M. B.; Szögi, A. A. Anammox sludge immobilized in
616
polyvinyl alcohol (PVA) cryogel carriers. Bioresour. Technol. 2012, 114, 231–240.
617
(37)van der Star, W. R. L.; Abma, W. R.; Blommers, D.; Mulder, J. W.; Tokutomi, T.;
618
Strous, M.; Picioreanu, C.; van Loosdrecht, M. C. M. Startup of reactors for anoxic
619
ammonium oxidation: Experiences from the first full-scale anammox reactor in
620
Rotterdam. Water Res. 2007, 41 (18), 4149–4163.
621
(38) Oshiki, M.; Shinyako-Hata, K.; Satoh, H.; Okabe, S. Draft genome sequence of
622
an anaerobic ammonium-oxidizing bacterium,“Candidatus Brocadia sinica.” Genome
623
Announc. 2015, 3 (2), 3–4. 29
ACS Paragon Plus Environment
Environmental Science & Technology
Page 30 of 43
624
(39) Chen, H.; Hu, H.; Chen, Q.; Shi, M.; Jin, R. Successful start-up of the anammox
625
process: influence of the seeding strategy on performance and granule properties.
626
Bioresour. Technol. 2016, 211, 594–602.
627
(40) Cotton, C. A.; Edlich-Muth, C.; Bar-Even, A. Reinforcing carbon fixation : CO2
628
reduction replacing and supporting carboxylation. Curr. Opin. Biotechnol. 2018, 49,
629
49–56.
630
(41) Wang, Y.; Hu, X.; Jiang, B.; Song, Z.; Ma, Y. Symbiotic relationship analysis of
631
predominant bacteria in a lab-scale anammox UASB bioreactor. Environ. Sci. Pollut.
632
Res. 2016, 23 (8), 7615–7626.
633
(42)Brakhage, A. A. Regulation of fungal secondary metabolism. Nat. Rev. Microbiol.
634
2013, 11 (1), 21–32.
635
(43) Brader. G.; Compant. S.; Mitter, B.; Mitter, B.; Trognitz, F.; Sessitsch, A.
636
Metabolic potential of endophytic bacteria. Curr. Opin. Biotechnol. 2014, 27, 30–37.
637
(44) Watrous, J.; Hendricks, N.; Meehan, M.; Dorrestein, P. C. Capturing bacterial
638
metabolic exchange using thin film desorption electrospray ionization-imaging mass
639
spectrometry. Anal. Chem. 2010, 82 (5), 1598–1600.
640
(45) Phelan, V. V.; Liu, W. T.; Pogliano, K.; Dorrestein, P. C. Microbial metabolic
641
exchangeg-the chemotype-to-phenotype link. Nat. Chem. Biol. 2012, 8 (1), 26–35.
642
(46)Majerczyk, C.; Brittnacher, M.; Jacobs, M.; Armour, C. D.; Radey, M.; Schneider,
643
E.; Phattarasokul, S.; Bunt, R.; Peter Greenberg, E. Global analysis of the
644
Burkholderia thailandensis quorum sensing-controlled regulon. J. Bacteriol. 2014,
645
196 (7), 1412–1424. 30
ACS Paragon Plus Environment
Page 31 of 43
Environmental Science & Technology
646
(47) Zhou, J.; Ma, Q.; Yi, H.; Wang, L.; Song, H.; Yuan, Y. J. Metabolome profiling
647
reveals metabolic cooperation between Bacillus megaterium and Ketogulonicigenium
648
vulgare during induced swarm motility. Appl. Environm. Microbiol. 2011, 77 (19),
649
7023–7030.
650
(48) Strous, M.; Fuerst, J. a.; Kramer, E. H.; Logemann, S.; Muyzer,
651
G.; van de Pas-Schoonen, K. T.; Webb, R.; Kuenen, J. G.; Jetten, M. S. M. Missing
652
lithotroph identified as new planctomycete. Nature 1999, 400 (6743), 446–449.
653
(49) Lovley, D. R. Happy together: Microbial communities that hook up to swap
654
electrons. ISME J. 2017, 11 (2), 327–336.
655
(50) Pande, S.; Shitut, S.; Freund, L.; Westermann, M.; Bertels, F.; Colesie, C.;
656
Bischofs, I. B.; Kost, C. Metabolic cross-feeding via intercellular nanotubes among
657
bacteria. Nat. Commun. 2015, 6, 6238.
658
(51) Hasegawa, Y.; Futamata, H.; Tashiro, Y. Complexities of cell-to-cell
659
communication through membrane vesicles: Implications for selective interaction of
660
membrane vesicles with microbial cells. Front. Microbiol. 2015, 6, 633..
661
(52) Wang, R.; Xu, S.; Wang, N.; Xia, B.; Jiang, Y.; Wang, R. Transcriptome analysis
662
of secondary metabolism pathway, transcription factors, and transporters in response
663
to methyl jasmonate in Lycoris aurea. Front. Plant Sci. 2017, 7 1971.
664
(53) Flemming, H. C.; Wingender, J. The biofilm matrix. Nat. Rev. Microbiol. 2010, 8
665
(9), 623–633.
666
(54) Jennings, L. K.; Storek, K. M.; Ledvina, H. E.; Coulon, C.; Marmont, L. S.;
667
Sadovskaya, I.; Secor, P. R.; Tseng, B. S.; Scian, M.; Filloux, A.; Wozniak, D. J.; 31
ACS Paragon Plus Environment
Environmental Science & Technology
Page 32 of 43
668
Howell, P. L.; Parsek, M. R. Pel is a cationic exopolysaccharide that cross-links
669
extracellular DNA in the Pseudomonas aeruginosa biofilm matrix. Proc. Natl. Acad.
670
Sci. U. S. A. 2015, 112 (36), 11353–11358.
671
(55) Silva, I. N.; Ramires, M. J.; Azevedo, L. A.; Guerreiro, A. R.; Tavares, A. C.;
672
Becker, J. D.; Moreira, L. M. The regulator LdhR and D-lactate dehydrogenase LdhA
673
of Burkholderia multivorans play a role in carbon overflow and in planktonic cellular
674
aggregate formation. Appl. Environ. Microbiol. 2017, 83 (19), 01343-17.
675
(56) Li, X.; Luo, J.; Guo, G.; Mackey, H. R.; Hao, T.; Chen, G. Seawater-based
676
wastewater accelerates development of aerobic granular sludge: A laboratory
677
proof-of-concept. Water Res. 2017, 115, 210–219.
678
(57) Tang, X.; Guo, Y.; Wu, S.; Chen, L.; Tao, H.; Liu, S. Metabolomics uncovers the
679
regulatory pathway of acyl-homoserine lactones-based quorum sensing in anammox
680
consortia. Environ. Sci. Technol. 2018, 52 (4), 2206–2216.
681
(58) Adav, S. S.; Lee, D.; Lai, J. Proteolytic activity in stored aerobic granular sludge
682
and structural integrity. Bioresour. technol. 2009, 100 (1), 68–73.
683
(59) Krause, S. M. B.; Johnson, T.; Samadhi Karunaratne, Y.; Fu, Y.; Beck, D. A. C.;
684
Chistoserdova, L.; Lidstrom, M. E. Lanthanide-dependent cross-feeding of
685
methane-derived carbon is linked by microbial community interactions. Proc. Natl.
686
Acad. Sci. U. S. A. 2017, 114 (2), 358–363.
687
32
ACS Paragon Plus Environment
Page 33 of 43
688
Bin ID AMX1 AMX2 ACD1 ATM1 ATM2 ATN1 ATN2 ATN3 ATN4 CFX1 CFX2 CFX3 CFX4 CFX5 CFX6 CFX7 CFX8
Environmental Science & Technology
Table 1. 37 genome statistics recovered in this SBR. Taxonomy Bacteria; Planctomycetes; Planctomycetia; Planctomycetales; Planctomycetaceae; Brocadia Bacteria; Planctomycetes; Planctomycetia; Planctomycetales; Planctomycetaceae; Brocadia Bacteria; Acidobacteria Bacteria; Armatimonadetes Bacteria; Armatimonadetes Bacteria; Actinobacteria; Actinobacteria; Corynebacteriales; Nocardiaceae; Rhodococcus Bacteria; Actinobacteria; Actinobacteria Bacteria; Actinobacteria Bacteria; Actinobacteria; Actinobacteria; Micrococcales; Intrasporangiaceae; Ornithinimicrobium Bacteria; Chloroflexi; Anaerolineae Bacteria; Chloroflexi Bacteria; Chloroflexi Bacteria; Chloroflexi; Anaerolineae Bacteria; Chloroflexi; Caldilineae; Caldilineales; Caldilineaceae; Caldilinea Bacteria; Chloroflexi; Anaerolineae Bacteria; Chloroflexi; Caldilineae Bacteria; Chloroflexi; Sphaerobacteridae; Sphaerobacterales;
N50 Length (bp)
Predicted genes
3,796,533 42.25
77,967
3,630
253
3,493,969 45.00
23,006
3,326
0.85 1.82 1.82
132 29 23
2,802,878 54.67 2,809,340 60.98 2,746,897 60.69
36,387 156,133 167,677
2,663 2,587 2,601
95.1
0.38
268
4,790,008 70.96
35,293
4,631
99.2 97.4
2.14 1.28
68 104
4,051,767 67.23 2,867,214 59.93
140,716 42,447
3,731 2,781
85.0
0.54
346
3,062,586 73.54
10,976
3,128
86.2 72.1 96.3 91.8
7.71 2.94 0 4.61
540 908 139 153
3,574,124 3,466,261 3,966,637 2,971,039
60.34 56.47 62.82 61.17
9,222 4,548 51,483 34,524
3,573 4,142 3,454 2,887
96.4
3.03
847
9,366,412 59.40
18,517
8,515
83.0 89.6 87.3
0.18 3.13 0.94
18 1075 535
3,467,924 53.68 5,419,247 64.66 4,430,357 63.96
578,594 6,426 11,330
3,296 5,297 4,345
Completeness (%)
Contamination (%)
# Scaffolds
95.6
0.55
82
98.9
6.04
78.5 95.0 94.2
Genome size (bp)
GC (%)
33
ACS Paragon Plus Environment
Environmental Science & Technology
CFX9 CHB1 NOB1 PLA1 PLA2 PLA3 PLA4 PLA5 PRO1 PRO2 PRO3 PRO4 PRO5 PRO6 PRO7 GWF1 KSB1 UCB1 UCB2 UCB3
Sphaerobacterineae Bacteria; Chloroflexi Bacteria; Ignavibacteriae; Ignavibacteria; Ignavibacteriales Bacteria; Nitrospirae; Nitrospirales; Nitrospiraceae; Nitrospira Bacteria; Planctomycetes Bacteria; Planctomycetes Bacteria; Planctomycetes Bacteria; Planctomycetes; Planctomycetia; Planctomycetales; Planctomycetaceae; Blastopirellula Bacteria; Planctomycetes; Planctomycetia; Planctomycetales; Planctomycetaceae; Pirellula Bacteria; Proteobacteria; Alphaproteobacteria; Rhizobiales; Phyllobacteriaceae; Aminobacter Bacteria; Proteobacteria; Betaproteobacteria; Burkholderiales; Burkholderiaceae; Lautropia Bacteria; Proteobacteria; Deltaproteobacteria Bacteria; Proteobacteria; Betaproteobacteria; Burkholderiales Bacteria; Proteobacteria; Alphaproteobacteria Bacteria; Proteobacteria; Alphaproteobacteria; Rhizobiales Bacteria; Proteobacteria; Deltaproteobacteria Bacteria; Microgenomates Bacteria; Candidate division KSB1 Bacteria; Candidatus Uhrbacteria Bacteria Bacteria
Page 34 of 43
90.4 94.8 96.8 94.3 97.7 89.8
1.82 0 5.23 3.41 3.41 0
357 30 82 59 65 44
3,726,020 2,332,291 3,961,664 3,885,146 4,711,417 3,355,315
56.77 37.60 60.33 64.60 63.76 63.47
14,876 129,973 86,454 125,775 95,457 100,689
3,597 2,082 3,787 3,207 3,764 2,747
89.4
0.59
417
5,112,838 66.05
17,011
4,309
97.5
0
289
7,048,853 63.88
41,231
5,340
92.3
1.33
46
3,123,305 69.36
101,651
2,980
94.5
1.65
36
3,224,835 69.45
127,685
3,017
90.9 75.9 89.5 77.8 77.4 76.3 92.2 86.2 71.9 76.9
1.94 0 0 0.79 0.84 0 3.3 0 0 0.92
422 209 14 49 643 33 776 4 15 489
5,539,874 2,591,218 3,248,505 2,438,700 2,693,467 825,322 7,248,468 1,064,618 2,159,727 4,407,992
21,504 15,555 654,661 74,166 4,964 38,480 14,792 432,179 196,713 13,236
5,265 2,536 3,044 2,279 2,993 913 6,734 1,023 2,184 3,359
73.14 70.12 60.98 67.02 52.28 44.44 52.87 54.64 66.56 73.19
34
ACS Paragon Plus Environment
Page 35 of 43
Environmental Science & Technology
689
Figure captions
690
Figure 1. Profile of nitrogen removal during SBR start-up: (a) Time courses of
691
influent ammonia (NH4+–N) concentration (yellow line), effluent ammonia
692
concentration (green line) and ammonia removal rate (blue line). S1, S2, S3, S4 and
693
S5 represent five samples that were obtained on days 0, 106, 166, 218 and 280,
694
respectively. (b) Time courses of influent nitrite (NO2-–N) concentration (yellow line),
695
effluent nitrite concentration (green line) and nitrite removal rate (blue line). (c) Time
696
courses of influent nitrate (NO3-–N) concentration (yellow line), effluent nitrate
697
concentration (green line) and total nitrogen (TN) removal rate (blue line). (d) Time
698
courses of nitrogen loading rate (NLR) (orange line), nitrogen removal efficiency
699
(NRE) (blue line) and hydraulic retention time (HRT) (black line).
700
Figure 2. The phylogenetic tree and relative abundance of 37 recovered genomes in
701
this SBR. (a) Phylogenetic tree of the recovered 37 draft genomes (red) and 56
702
reference genomes (black) deposited in GenBank. GenBank accession numbers of
703
each genome were provided in parentheses. Branch nodes numbers represent the
704
bootstrap values. The tree was constructed using the software RAxML based on the
705
37 Phylosift marker genes. (b) The relative abundance of 37 microorganisms in five
706
samples. Blue, yellow, grey, coffee and light blue columns represent S1, S2, S3, S4
707
and S5, respectively. Five samples were obtained on day 0, 106, 166, 218 and 280,
708
respectively.
709
Figure 3. (a) The total functional gene abundance on the level of COG classes in 37
710
MAGs for five samples. Blue, yellow, grey, coffee and light blue columns represent 35
ACS Paragon Plus Environment
Environmental Science & Technology
Page 36 of 43
711
S1, S2, S3, S4 and S5, respectively. The community functional structure analysis
712
using the PCA analysis of functional gene abundance on the COG classes level based
713
on the sum of gene abundance of a certain COG class in every MAG in in (b) S1, (c)
714
S2, (d) S3, (e) S4, and (f) S5, respectively. The gene abundance was calculated by
715
normalizing the number of reads that were mapped to the gene sequence by the
716
number of the 16S rRNA gene sequence identified for the metagenomic sequencing
717
data and gene sequence length.
718
Figure 4. Potential cross-feedings and variations gene abundances potentially
719
associated to the cross-feeding between Brocadia (AMX1), and Chloroflexi (CFX3
720
and CFX4), Armatimonadetes (ATM1 and ATM2), Proteobacteria (PRO1), Chlorobi
721
(CHB1). The Armatimonadetes and Proteobacteria bacterium encoded the pathways
722
for the production of secondary metabolites for the anammox metabolism. The
723
Chloroflexi bacterium encoded the pathways for the biosynthesis of sticky
724
macromolecular exopolysaccharides (CFX3) and denitrification of nitrate (CFX4).
725
The Chlorobi bacterium had the potential to hydrolyze extracellular proteins and
726
liberate amino acids. The heat map represents the relative abundance of key genes that
727
were complementary between symbiotic and anammox bacteria and were potentially
728
associated with microbial interactions in the studied genomes for five samples. S1-S5
729
represent the samples obtained on days 0, 106, 166, 218 and 280, respectively.
730
Definitions for genes and abbreviations for metabolites are listed in SI Table S8 and
731
S9, respectively.
732 36
ACS Paragon Plus Environment
Page 37 of 43
Environmental Science & Technology
733 734
Figure 1. Profile of nitrogen removal during SBR start-up: (a) Time courses of
735
influent ammonia (NH4+–N) concentration (yellow line), effluent ammonia
736
concentration (green line) and ammonia removal rate (blue line). S1, S2, S3, S4 and
737
S5 represent five samples that were obtained on days 0, 106, 166, 218 and 280,
738
respectively. (b) Time courses of influent nitrite (NO2-–N) concentration (yellow line),
739
effluent nitrite concentration (green line) and nitrite removal rate (blue line). (c) Time
740
courses of influent nitrate (NO3-–N) concentration (yellow line), effluent nitrate
741
concentration (green line) and total nitrogen (TN) removal rate (blue line). (d) Time
742
courses of nitrogen loading rate (NLR) (orange line), nitrogen removal efficiency
743
(NRE) (blue line) and hydraulic retention time (HRT) (black line).
744
37
ACS Paragon Plus Environment
Environmental Science & Technology
Page 38 of 43
745 746
Figure 2. The phylogenetic tree and relative abundance of 37 recovered genomes in
747
this SBR. (a) Phylogenetic tree of the recovered 37 draft genomes (red) and 56 38
ACS Paragon Plus Environment
Page 39 of 43
Environmental Science & Technology
748
reference genomes (black) deposited in GenBank. GenBank accession numbers of
749
each genome were provided in parentheses. Branch nodes numbers represent the
750
bootstrap values. The tree was constructed using the software RAxML based on the
751
37 Phylosift marker genes. (b) The relative abundance of 37 microorganisms in five
752
samples. Blue, yellow, grey, coffee and light blue columns represent S1, S2, S3, S4
753
and S5, respectively. Five samples were obtained on day 0, 106, 166, 218 and 280,
754
respectively.
39
ACS Paragon Plus Environment
Environmental Science & Technology
Page 40 of 43
755 756
Figure 3. (a) The total functional gene abundance on the level of COG classes in 37
757
MAGs for five samples. Blue, yellow, grey, coffee and light blue columns represent
758
S1, S2, S3, S4 and S5, respectively. The community functional structure analysis
759
using the PCA analysis of functional gene abundance on the COG classes level based
760
on the sum of gene abundance of a certain COG class in every MAG in in (b) S1, (c)
761
S2, (d) S3, (e) S4, and (f) S5, respectively. The gene abundance was calculated by
762
normalizing the number of reads that were mapped to the gene sequence by the
763
number of the 16S rRNA gene sequence identified for the metagenomic sequencing
764
data and gene sequence length.
765
40
ACS Paragon Plus Environment
Page 41 of 43
Environmental Science & Technology
766 767
Figure 4. Potential cross-feedings and variations gene abundances potentially
768
associated to the cross-feeding between Brocadia (AMX1), and Chloroflexi (CFX3
769
and CFX4), Armatimonadetes (ATM1 and ATM2), Proteobacteria (PRO1), Chlorobi
770
(CHB1). The Armatimonadetes and Proteobacteria bacterium encoded the pathways
771
for the production of secondary metabolites for the anammox metabolism. The
772
Chloroflexi bacterium encoded the pathways for the biosynthesis of sticky
773
macromolecular exopolysaccharides (CFX3) and denitrification of nitrate (CFX4).
774
The Chlorobi bacterium had the potential to hydrolyze extracellular proteins and
775
liberate amino acids. The heat map represents the relative abundance of key genes that 41
ACS Paragon Plus Environment
Environmental Science & Technology
Page 42 of 43
776
were complementary between symbiotic and anammox bacteria and were potentially
777
associated with microbial interactions in the studied genomes for five samples. S1-S5
778
represent the samples obtained on days 0, 106, 166, 218 and 280, respectively.
779
Definitions for genes and abbreviations for metabolites are listed in SI Table S8 and
780
S9, respectively.
781
42
ACS Paragon Plus Environment
Page 43 of 43
Environmental Science & Technology
782
TOC/Abstract graphic
783 784
For Table of Contents Only
43
ACS Paragon Plus Environment