ArticlePDF Available

TbasCO: trait-based comparative ‘omics identifies ecosystem-level and niche-differentiating adaptations of an engineered microbiome

Authors:

Abstract and Figures

A grand challenge in microbial ecology is disentangling the traits of individual populations within complex communities. Various cultivation-independent approaches have been used to infer traits based on the presence of marker genes. However, marker genes are not linked to traits with complete fidelity, nor do they capture important attributes, such as the timing of gene expression or coordination among traits. To address this, we present an approach for assessing the trait landscape of microbial communities by statistically defining a trait attribute as a shared transcriptional pattern across multiple organisms. Leveraging the KEGG pathway database as a trait library and the Enhanced Biological Phosphorus Removal (EBPR) model microbial ecosystem, we demonstrate that a majority (65%) of traits present in 10 or more genomes have niche-differentiating expression attributes. For example, while many genomes containing high-affinity phosphorus transporter pstABCS display a canonical attribute (e.g. up-regulation under phosphorus starvation), we identified another attribute shared by many genomes where transcription was highest under high phosphorus conditions. Taken together, we provide a novel framework for unravelling the functional dynamics of uncultivated microorganisms by assigning trait-attributes through genome-resolved time-series metatranscriptomics.
Content may be subject to copyright.
ARTICLE OPEN
TbasCO: trait-based comparative omics identies
ecosystem-level and niche-differentiating adaptations
of an engineered microbiome
E. A. McDaniel
1,2,8
, J. J. M. van Steenbrugge
3,4,5,8
, D. R. Noguera
6
, K. D. McMahon
1,6
, J. M. Raaijmakers
4,7
, M. H. Medema
3,7
and
B. O. Oyserman
3,4
© The Author(s) 2022
A grand challenge in microbial ecology is disentangling the traits of individual populations within complex communities. Various
cultivation-independent approaches have been used to infer traits based on the presence of marker genes. However, marker genes
are not linked to traits with complete delity, nor do they capture important attributes, such as the timing of gene expression or
coordination among traits. To address this, we present an approach for assessing the trait landscape of microbial communities by
statistically dening a trait attribute as a shared transcriptional pattern across multiple organisms. Leveraging the KEGG pathway
database as a trait library and the Enhanced Biological Phosphorus Removal (EBPR) model microbial ecosystem, we demonstrate
that a majority (65%) of traits present in 10 or more genomes have niche-differentiating expression attributes. For example, while
many genomes containing high-afnity phosphorus transporter pstABCS display a canonical attribute (e.g. up-regulation under
phosphorus starvation), we identied another attribute shared by many genomes where transcription was highest under high
phosphorus conditions. Taken together, we provide a novel framework for unravelling the functional dynamics of uncultivated
microorganisms by assigning trait-attributes through genome-resolved time-series metatranscriptomics.
ISME Communications; https://doi.org/10.1038/s43705-022-00189-2
INTRODUCTION
A longstanding cornerstone of deterministic ecological theory is
that the environment selects for traits. Traits may be dened as
any physiological, morphological, or genomic signature that
affects the tness or function of an individual [1]. Trait-based
approaches have become indispensable in macroecological
systems to describe tness trade-offs and the effects of
biodiversity on ecosystem functioning [25]. Recently, trait-based
frameworks have been proposed as an alternative to taxonomy-
based methods for describing microbial ecosystem processes
[6,7]. Connecting microbial traits and their phylogenetic distribu-
tions to ecosystem-level functions can provide powerful insights
into the ecological and evolutionary dynamics underpinning
community assembly, microbial biogeography, and organismal
responses to changes in the environment [810]. Additionally,
pinpointing the organismal distribution of traits and the ecological
selective pressures that enrich them may be leveraged to
reproducibly and rationally engineer stable, functionally redun-
dant ecosystems [1115]. However, applying trait-based
approaches to microbial communities is challenging due to the
difculty in identifying and measuring relevant ecological traits for
a given ecosystem [16].
High-throughput sequencing technologies and multi-omics
techniques are now routinely used to describe the diversity,
activity, and functional potential of uncultivated microbial
lineages [1721]. Improvements in bioinformatics algorithms,
and in particular metagenomic binning methods, have allowed
for genome-resolved investigations of microbial communities
rather than gene-based analyses of assembled contigs [22]. These
(meta) genomes are subsequently leveraged to detect the
presence of key genes or pathways and predict specic traits of
the whole community [19,23]. Integrating metatranscriptomics
data addresses a key limitation, as expression patterns better
reect the actual functional dynamics of a trait compared to gene
presence alone. Here, we present TbasCO, a software package and
statistical framework for Trait-based Comparative Omics to
identify expression attributes. We adopt the terminology attribute
as a hierarchically structured feature of a trait and assert that
statistically similar transcriptional patterns of traits across multiple
organisms be treated as attributes (Fig. 1). This new terminology
addresses two key semantic challenges. First, by extending upon
the current usage of the term traitfor the presence and absence
of pathways to the corresponding transcriptional patterns.
Second, it addresses a limitation of the terminology of co-
Received: 20 May 2022 Revised: 29 September 2022 Accepted: 10 October 2022
1
Department of Bacteriology, University of WisconsinMadison, Madison, WI, USA.
2
Microbiology Doctoral Training Program, University of WisconsinMadison, Madison, WI,
USA.
3
Bioinformatics Group, Wageningen University and Research, Wageningen, The Netherlands.
4
Microbial Ecology, Netherlands Institute of Ecological Research, Wageningen,
The Netherlands.
5
Laboratory of Nematology, Wageningen University, Wageningen, The Netherlands.
6
Department of Civil and Environmental Engineering, University of
WisconsinMadison, Madison, WI, USA.
7
Institute of Biology, Leiden University, Leiden, Netherlands.
8
These authors contributed equally: E. A. McDaniel, J. J. M. van Steenbrugge.
email: elizabethmcd93@gmail.com; jorisvansteebrugge@gmail.com; BenOyserman@gmail.com
www.nature.com/ismecomms
1234567890();,:
expression, which becomes biologically inaccurate when compar-
ing across independent populations of organisms within a
community. In this manner, the identication of expression-
based attributes provides a high-throughput and intuitive frame-
work for extending trait-based methods to time-series expression
patterns in microbial communities. We implement this trait-based
approach to classify transcriptional attributes in a microbial
community performing Enhanced Biological Phosphorus Removal
(EBPR), a globally important biotechnological process implemen-
ted in numerous wastewater treatment plants (WWTPs).
The fundamental feature of the engineered EBPR ecosystem is
the decoupled and cyclic availability of an external carbon source
and terminal electron acceptor. This cycling is often referred to as
feast-famineconditions and provides a strong selective pressure
for traits such as polymer cycling. Accumulation of intracellular
polyphosphate through cyclic anaerobic-aerobic conditions ulti-
mately results in net phosphorus removal and accomplishes the
EBPR process [24,25]. One of the most well-studied polypho-
sphate accumulating organisms (PAOs) belongs to the unculti-
vated bacterial lineage Candidatus Accumulibacter phosphatis
(hereby referred to as Accumulibacter) [24,26]. Numerous
genome-resolved omics methods have been used to investigate
the physiology and regulation of this model PAO enriched in
engineered lab-scale enrichment bioreactor systems [2735].
However, novel and putative PAOs have been discovered that
remove phosphorus without exhibiting the hallmark traits of
Accumulibacter [3641]. Additionally, although these lab-scale
systems are designed to specically enrich for Accumulibacter, a
diverse bacterial community persists in these environments [27],
and their ecological roles have largely remained unexplored. As a
result, the general adaptations of microbial lineages inhabiting the
EBPR community are not well understood. Using genome-resolved
metagenomics and metatranscriptomics, we assembled 66
species-representative genomes spanning several signicant EBPR
lineages and identied the distribution of expression-based
attributes. We show that while some expression attributes are
distributed in few genomes, many are redundant and shared
across many lineages. Furthermore, we nd that a majority of core
traits (as dened by the presence of marker genes) have multiple
attributes, suggesting that identifying niche-differentiating
expression attributes may be used to reveal a large hidden
metabolic versatility when investigating genomic data alone.
MATERIALS AND METHODS
Metagenomic assembly, annotation, and metatranscriptomic
mapping
Three metagenomes sampled from an EBPR bioreactor in May of 2013 with
linked time-series metatranscriptomics data were sequenced [42]. Samples
were collected and DNA extracted according to the Supplemental
Methods. Metagenomic samples were processed and assembled into 66
species-representative bins as described in detail in the Supplemental
Methods. All bins are greater than 75% complete and contain less than
10% contamination, with a large majority (44/66) >95% complete and <5%
redundant as calculated by CheckM [43] and are all described in Table 1.
Each bin was functionally annotated using the KEGG database through
an HMM-based approach under KEGG release 93.0 using the command-
Fig. 1 Overview of trait-based comparative transcriptomics approach In genome-resolved metagenomics approaches, representative MAGs
are assembled from a microbial community of interest, and the presence and/or absence of key metabolic pathways are used to make
inferences of metabolic potential and ecosystem processes. However, metagenomic data alone can only assess the metabolic potential of a
given pathway, and do not provide other biologically relevant information such as the timing or induction of these traits. Using time-series
metatranscriptomics, we developed a trait-based comparative omics (TbasCO) pipeline that statistically assesses the inter-organismal
differences in gene expression pattern of a given trait to cluster into trait attributes. As expression patterns are determined by the time-points
assessed in an experiment, it is important to design the sampling regime to capture relevant ecophysiological changes within the ecosystem.
E.A. McDaniel et al.
2
ISME Communications
Table 1. Genome quality statistics and relative abundance calculations for all 66 EBPR SBR MAGs.
Code Genbank
accession
Classication Completeness Contamination Size
(Mbp)
Contigs GC Abundance
2013-5-13
Abundance
2013-5-23
Abundance
2013-5-28
Total
Transcriptional
Reads Mapped
Total
rRNAs
Total
tRNAs
AUS1 GCA_020161845.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Actinomycetales;f__
Dermatophilaceae;g__
Austwickia;s__
99.45 5.01 4.39 82 71.2 0.261 0.720 0.124 255331 3 61
PHYC1 GCA_020161815.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Actinomycetales;f__
Dermatophilaceae;g__
Phycicoccus;s__
98.02 0.54 3.06 34 71 1.355 3.007 0.341 332509 1 49
PHYC2 GCA_020161155.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Actinomycetales;f__
Dermatophilaceae;g__
Phycicoccus;s__
Phycicoccus
95.82 1.89 3.20 111 69.2 0.047 0.174 0.112 152031 1 52
TET1 GCA_020160805.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Actinomycetales;f__
Dermatophilaceae;g__
Tetrasphaera_A;s__
98.42 0.54 3.75 57 67.9 0.446 0.436 0.507 1378316 2 47
TET2 GCA_020160795.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Actinomycetales;f__
Dermatophilaceae;g__
Tetrasphaera_A;s__
Tetrasphaera_A
98.92 0.05 3.96 66 69.3 0.803 0.236 1.244 2538782 1 76
LEU1 GCA_020161315.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Actinomycetales;f__
Microbacteriaceae;g__
Leucobacter;s__
96.06 2.05 3.01 74 63.5 0.272 0.083 0.093 99061 3 47
LEU2 GCA_020161175.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Actinomycetales;f__
Microbacteriaceae;g__
Leucobacter;s__
Leucobacter
83.22 1.48 2.31 140 64.8 0.065 0.101 0.092 22050 2 44
SAL1 GCA_020160915.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Actinomycetales;f__
Microbacteriaceae;g__
Salinibacterium;s__
97.81 0 2.93 8 67.2 0.335 0.142 0.559 178111 2 45
NANO1 GCA_020161245.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Nanopelagicales;f__;g__;s__
99.14 3.68 4.29 95 72.7 0.106 0.047 0.172 64510 1 52
PROP1 GCA_020161795.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Propionibacteriales;f__
Propionibacteriaceae;g__;s__
91.04 0.91 3.47 67 69.3 0.063 0.108 0.206 100351 0 60
PROP2 GCA_020161755.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Propionibacteriales;f__
Propionibacteriaceae;g__
Propionicimonas;s__
93.63 3.02 4.08 61 70.7 0.094 0.046 0.413 130384 3 52
E.A. McDaniel et al.
3
ISME Communications
Table 1. continued
Code Genbank
accession
Classication Completeness Contamination Size
(Mbp)
Contigs GC Abundance
2013-5-13
Abundance
2013-5-23
Abundance
2013-5-28
Total
Transcriptional
Reads Mapped
Total
rRNAs
Total
tRNAs
PROP3 GCA_020161015.1 d__Bacteria;p__
Actinobacteriota;c__
Actinobacteria;o__
Propionibacteriales;f__
Propionibacteriaceae;g__
Propionicimonas;s__
94.14 3.15 3.67 65 71.6 0.074 0.176 0.249 96105 0 51
FIMBRI1 GCA_020161505.1 d__Bacteria;p__
Armatimonadota;c__
Fimbriimonadia;o__
Fimbriimonadales;f__
Fimbriimonadaceae;g__
Uphvl-Ar1;s__
96.55 0 3.14 38 58.8 0.068 0.234 0.009 27830 3 48
BAC1 GCA_020161835.1 d__Bacteria;p__
Bacteroidota;c__
;o__;f__;g__;s__
94.52 0 4.40 36 41.6 0.345 0.024 0.003 32140 4 42
BAC2 GCA_020162035.1 d__Bacteria;p__
Bacteroidota;c__
Bacteroidia;o__
AKYH767;f__
b-17BO;g__;s__
99.05 0.48 3.17 31 29.6 0.757 0.010 0.015 46346 3 32
CHIT1 GCA_020161435.1 d__Bacteria;p__
Bacteroidota;c__
Bacteroidia;o__
Chitinophagales;f__
Chitinophagaceae;g__;s__
99.01 0 4.19 10 46.3 0.183 0.174 3.613 3141341 0 34
CHIT2 GCA_020161535.1 d__Bacteria;p__
Bacteroidota;c__
Bacteroidia;o__
Chitinophagales;f__
Chitinophagaceae;g__
Flavihumibacter;s__
100 1.23 4.03 23 48.2 0.195 0.383 0.033 24003 3 40
SAP1 GCA_020160935.1 d__Bacteria;p__
Bacteroidota;c__
Bacteroidia;o__
Chitinophagales;f__
Saprospiraceae;g__;s__
96.53 0.99 5.84 51 50.3 0.226 0.007 0.128 702648 3 36
SAP2 GCA_020160855.1 d__Bacteria;p__
Bacteroidota;c__
Bacteroidia;o__
Chitinophagales;f__
Saprospiraceae;g__
OLB8;s__
97.52 0.5 3.73 65 37.2 0.290 0.167 0.016 10636 3 34
LEAD1 GCA_020161355.1 d__Bacteria;p__
Bacteroidota;c__
Bacteroidia;o__
Cytophagales;f__
Spirosomaceae;g__
Leadbetterella;s__
99.11 0.6 4.81 17 37.7 0.136 0.002 0.858 1017458 2 36
RUN1 GCA_020161055.1 d__Bacteria;p__
Bacteroidota;c__
Bacteroidia;o__
Cytophagales;f__
Spirosomaceae;g__
Runella;s__Runella
100 0 7.44 60 44.4 0.124 1.088 1.749 10725342 2 40
FLAVO1 GCA_020161455.1 d__Bacteria;p__
Bacteroidota;c__
Bacteroidia;o__
Flavobacteriales;f__
Flavobacteriaceae;g__
Flavobacterium;s__
99.29 0.35 3.08 18 32.5 0.030 0.002 0.742 3002991 3 36
CHRYS1 GCA_020161485.1 d__Bacteria;p__
Bacteroidota;c__
Bacteroidia;o__
Flavobacteriales;f__
Weeksellaceae;g__
Chryseobacterium_A;s__
Chryseobacterium_A
100 0.25 2.57 11 36.7 0.107 3.917 0.358 209940 2 35
E.A. McDaniel et al.
4
ISME Communications
Table 1. continued
Code Genbank
accession
Classication Completeness Contamination Size
(Mbp)
Contigs GC Abundance
2013-5-13
Abundance
2013-5-23
Abundance
2013-5-28
Total
Transcriptional
Reads Mapped
Total
rRNAs
Total
tRNAs
BAC3 GCA_020162015.1 d__Bacteria;p__
Bacteroidota;c__
Bacteroidia;o__
NS1112g;f__
UKL13-3;g__B1;s__
100 0 3.74 45 41.1 0.445 0.892 0.693 9991372 0 34
IGNAVI1 GCA_020161395.1 d__Bacteria;p__
Bacteroidota;c__
Ignavibacteria;o__
Ignavibacteriales;f__
Ignavibacteriaceae_A;g__
UTCHB3;s__
97.27 0.55 4.07 21 42.2 0.163 0.635 0.025 58496 3 44
RTHERM1 GCA_020160835.1 d__Bacteria;p__
Bacteroidota;c__
Rhodothermia;o__
Rhodothermales;f__;g__;s__
98.36 1.38 3.25 36 67 0.328 0.050 0.060 116472 3 52
ANAER1 GCA_020161935.1 d__Bacteria;p__
Chloroexota;c__
Anaerolineae;o__
SBR1031;f__A4b;g__;s__
98.17 0 7.64 32 54.2 0.375 0.190 0.153 910673 4 48
HERP1 GCA_020161265.1 d__Bacteria;p__
Chloroexota;c__
Chloroexia;o__
Chloroexales;f__
Herpetosiphonaceae;g__
Herpetosiphon;s__
99.09 0.91 6.04 13 50.2 0.774 0.025 0.008 7917 0 55
OBS1 GCA_020161235.1 d__Bacteria;p__
Cyanobacteria;c__
Vampirovibrionia;o__
Obscuribacterales;f__
Obscuribacteraceae;g__
Obscuribacter;s__
Obscuribacter
98.28 0.94 5.09 17 49.2 6.272 0.681 0.197 1713299 6 42
FUSI1 GCA_020161295.1 d__Bacteria;p__
Firmicutes_A;c__
Clostridia;o__
Peptostreptococcales;f__
Fusibacteraceae;g__
UBA5201;s__
96.5 1.75 3.08 41 42.8 0.001 0.580 0.001 11649 3 57
GEMMA1 GCA_020161135.1 d__Bacteria;p__
Gemmatimonadota;c__
Gemmatimonadetes;o__
Gemmatimonadales;f__
Gemmatimonadaceae;g__
;s__
98.35 3.3 4.55 8 70.1 0.004 0.031 0.494 2624259 3 55
SACCH1 GCA_020160975.1 d__Bacteria;p__
Patescibacteria;c__
Saccharimonadia;o__
Saccharimonadales;f__
Saccharimonadaceae;g__
Saccharimonas;s__
84.48 0 0.97 1 49.6 0.637 1.437 0.035 65079 3 43
ALPHA1 GCA_020161965.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
;f__;g__;s__
82.43 2.65 3.94 581 64.6 0.015 0.165 0.007 1283274 3 39
CAED1 GCA_020161545.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Caedimonadales;f__
UBA1908;g__;s__
86.36 1.1 1.88 96 52.8 0.034 0.201 0.002 41264 3 35
BREV1 GCA_020161595.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Caulobacterales;f__
Caulobacteraceae;g__
Brevundimonas;s__
Brevundimonas
97.51 3.41 3.07 155 67.2 0.011 0.254 0.004 27852 2 45
E.A. McDaniel et al.
5
ISME Communications
Table 1. continued
Code Genbank
accession
Classication Completeness Contamination Size
(Mbp)
Contigs GC Abundance
2013-5-13
Abundance
2013-5-23
Abundance
2013-5-28
Total
Transcriptional
Reads Mapped
Total
rRNAs
Total
tRNAs
CAULO1 GCA_020161365.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Caulobacterales;f__
Caulobacteraceae;g__
Caulobacter;s__
100 0 4.43 25 66.9 0.048 0.093 0.589 4627825 3 55
HYPHO1 GCA_020161405.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Caulobacterales;f__
Hyphomonadaceae;g__
UBA1942;s__
98.43 0.32 2.98 6 39.4 0.844 0.006 2.208 4138107 0 33
REYR1 GCA_020160955.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Reyranellales;f__
Reyranellaceae;g__
Reyranella;s__
89.96 7.34 5.08 210 70 0.057 0.090 0.238 224063 3 53
REYR2 GCA_020160995.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Reyranellales;f__
Reyranellaceae;g__
Reyranella;s__Reyranella
91.04 6.01 5.71 258 65.3 0.074 0.102 0.134 62018 1 53
ANDERS1 GCA_020161855.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhizobiales;f__
Anderseniellaceae;g__
PALSA-927;s__
97.64 0.4 3.36 19 61.6 0.187 0.175 0.029 25238 2 46
BEIJ1 GCA_020161915.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhizobiales;f__
Beijerinckiaceae;g__
Bosea;s__
81.6 8.48 4.44 777 66.3 0.156 0.319 0.423 338238 0 43
BEIJ2 GCA_020161975.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhizobiales;f__
Beijerinkiaceae_A;g__
;s__
81.18 5.25 3.99 465 62.5 0.042 0.157 0.018 28432 0 42
BEIJ3 GCA_020161475.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhizobiales;f__
Beijerinkiaceae_A;g__
PAR1;s__
76.21 1.72 3.08 320 63.3 0.017 1.744 0.099 77102 0 41
BEIJ4 GCA_020161575.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhizobiales;f__
Beijerinkiaceae_A;g__
PAR1;s__
97.89 0 3.19 17 63.2 0.176 0.538 0.014 26820 0 47
PHREA1 GCA_020161695.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhizobiales;f__
Phreatobacteraceae;g__
Phreatobacter;s__
98.35 3.96 4.69 38 67.7 0.022 0.273 0.103 133243 1 50
RHIZO1 GCA_020161035.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhizobiales;f__
Rhizobiaceae;g__
Aminobacter;s__
Aminobacter
94.26 5.5 5.50 80 63.8 0.136 0.095 0.095 219213 3 48
E.A. McDaniel et al.
6
ISME Communications
Table 1. continued
Code Genbank
accession
Classication Completeness Contamination Size
(Mbp)
Contigs GC Abundance
2013-5-13
Abundance
2013-5-23
Abundance
2013-5-28
Total
Transcriptional
Reads Mapped
Total
rRNAs
Total
tRNAs
RHIZO2 GCA_020161665.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhizobiales;f__
Rhizobiaceae;g__
QFOR01;s__
88.41 2.12 3.39 43 60.6 0.035 0.335 0.003 24536 0 47
RHIZO3 GCA_020161625.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhizobiales;f__
Rhizobiaceae;g__
Shinella;s__Shinella
78.53 6.03 6.98 935 63.6 0.010 0.169 0.037 149921 0 48
RHODO1 GCA_020161655.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhodobacterales;f__
Rhodobacteraceae;g__
Deuviimonas;s__
100 0.35 4.08 24 65.5 0.321 0.141 0.848 3645270 0 44
RHODO2 GCA_020161615.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhodobacterales;f__
Rhodobacteraceae;g__
Pararhodobacter;s__
99.09 1.19 4.87 26 67.9 0.084 0.534 0.009 25807 1 49
RHODO3 GCA_020160875.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rhodospirillales_C;f__
Rhodospirillaceae_A;g__
;s__
91.2 2.27 3.76 236 62.2 0.153 0.046 0.185 153017 1 39
RICK1 GCA_020160775.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Rickettsiales;f__
Rickettsiaceae;g__
GCA-2402195;s__
75.59 1.58 1.18 82 34.5 0.085 0.075 0.052 17671 2 26
SPHING1 GCA_020160755.1 d__Bacteria;p__
Proteobacteria;c__
Alphaproteobacteria;o__
Sphingomonadales;f__
Sphingomonadaceae;g__
Sphingopyxis;s__
99.98 1.56 4.31 20 65.1 0.026 0.014 0.607 600695 3 47
ALIC1 GCA_020161945.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Burkholderiales;f__
Burkholderiaceae;g__
Alicycliphilus;s__
99.64 1.04 3.83 33 66.3 0.166 2.959 0.738 770970 1 48
OTTO1 GCA_020161215.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Burkholderiales;f__
Burkholderiaceae;g__
Ottowia;s__
93.66 5.56 4.52 250 67.1 0.011 0.276 0.004 26717 1 46
OTTO2 GCA_020161715.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Burkholderiales;f__
Burkholderiaceae;g__
Ottowia;s__Ottowia
99.26 0.62 3.40 35 69.1 0.372 4.140 0.424 121379 1 50
RAM1 GCA_020161775.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Burkholderiales;f__
Burkholderiaceae;g__
Ramlibacter;s__
99.84 0.06 4.36 32 66.1 0.778 0.536 1.814 1832037 1 45
E.A. McDaniel et al.
7
ISME Communications
Table 1. continued
Code Genbank
accession
Classication Completeness Contamination Size
(Mbp)
Contigs GC Abundance
2013-5-13
Abundance
2013-5-23
Abundance
2013-5-28
Total
Transcriptional
Reads Mapped
Total
rRNAs
Total
tRNAs
RUBRI1 GCA_020161065.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Burkholderiales;f__
Burkholderiaceae;g__
Rubrivivax;s__
99.52 0.05 6.29 41 71.2 0.236 0.347 0.306 1259737 1 73
VITREO1 GCA_020161145.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Burkholderiales;f__
Burkholderiaceae;g__
Vitreoscilla_A;s__
100 0.7 3.51 13 68.9 0.397 4.498 0.530 382529 1 46
CAPIA NA d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Burkholderiales;f__
Rhodocyclaceae;g__
Accumulibacter;s__
Accumulibacter
100 0.03 4.59 61 63.8 18.797 10.533 0.106 2411395 0 46
CAPIIA NA d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Burkholderiales;f__
Rhodocyclaceae;g__
Accumulibacter;s__
Accumulibacter
99.84 0.24 4.64 81 64.3 33.479 26.824 49.334 102762132 0 44
ZOO1 GCA_020161115.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Burkholderiales;f__
Rhodocyclaceae;g__
Zoogloea;s__
91.62 3.51 4.99 501 65.7 0.090 0.026 0.106 913411 4 59
LEG1 GCA_020161725.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Legionellales;f__
Legionellaceae;g__;s__
92.74 1.07 2.58 182 36.1 0.094 0.126 0.006 19591 1 27
LUTEI1 GCA_020161335.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Xanthomonadales;f__
Xanthomonadaceae;g__
Luteimonas;s__
96.89 0.71 3.56 252 69.9 0.002 0.309 0.011 49418 1 39
PSEUDO1 GCA_020160895.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Xanthomonadales;f__
Xanthomonadaceae;g__
Pseudoxanthomonas_A;s__
99.95 0.89 3.67 28 67.8 0.416 0.730 3.125 3964795 2 50
PSEUDO2 GCA_020161075.1 d__Bacteria;p__
Proteobacteria;c__
Gammaproteobacteria;o__
Xanthomonadales;f__
Xanthomonadaceae;g__
Pseudoxanthomonas;s__
99.02 0 2.99 6 69.6 1.750 6.111 1.228 515369 3 52
Genome code names match names used in all gures and within the text. Classications were assigned using the GTDB-tk [87] and conrmed by comparing against select publicly available references and a
subset of HQ MAGs from Singleton et al. 2021 [40]. Completeness and redundancy estimates and GC content were calculated by CheckM [43]. tRNA and rRNA predictions were performed with Barrnap as
part of the Prokka software [88]. Relative abundance estimates reect the proportion of reads mapped to the genome in that sample divided by the total number of reads mapped to all genomes as
performed with SingleM. Table available at https://gshare.com/articles/dataset/EBPR_SBR_MAGs_Metadata/13063874.
E.A. McDaniel et al.
8
ISME Communications
line KofamKOALA pipeline [44,45], selecting annotations that were
signicant hits above the specic HMM threshold. This resulted in 117,657
total annotations with 5,228 unique annotations. We used a metatran-
scriptomic dataset of six timepoints collected over a single EBPR cycle from
Oyserman et al. 2016 [42], with three timepoints from the anaerobic phase
and three from the aerobic phase. Raw reads were quality ltered using
BBtools suite v38.07 [46] and ribosomal rRNA was removed from each
sample using SortMeRNA [47]. Reads from each sample were mapped
against the concatenated set of open reading frames from all 66 bins using
kallisto v0.44.0 and parsed using the R package tximport [48,49].
TbasCO method implementation
The TbasCO package identies expression-based attributes of predened
traits using time-series (meta)transcriptomics data (Fig. 1). As expression
patterns are determined by the time-points assessed in an experiment, it is
important to design the sampling regime to capture relevant ecophysio-
logical changes within the ecosystem. In general, traits are dened by the
presence of a pathway or other collection of genes from an externally
provided database. A weighted distance metric between expression
patterns for all genes that dene a trait is calculated, and statistically
signicant similarity is determined based on the background distribution
of a trait of equal size. Thereby, two or more organisms with a statistically
similar expression pattern for a trait share an attribute. As the expression
proles of genes within a trait are compared across genomes indepen-
dently, co-expression of genes within a genome is not a pre-requisite for
identifying an attribute.
Input and preprocessing. The input that is accepted by TbasCO is a table
of RNAseq counts in csv format. Each row is treated as gene that has
columns for the gene/locus name, counts per sample, the genome the
gene belongs to, and the KEGG Orthology (KO) identier. The RNAseq
counts table may be provided pre-normalized or can be normalized by the
program. The default normalization method is designed to minimize
compositional bias in the differential abundance and activity of constituent
populations in metatranscriptomics studies. RNA expression counts are
therefore normalized relative to each genomic bin separately for each
sample [42]. After normalization, a pruning step is introduced to lter
genes that have zero counts or a mean absolute deviation of less than one
across all time points. To make inter-organismal comparisons of the
relative contribution of a gene to total measured organismal RNA, an
additional statistic is calculated ranking the expression counts from each
sample from highest to lowest. The ranks for each sample are then
normalized by dividing them by the maximum rank value in that sample.
This normalization is applied to make ranks comparable between
organisms with different genome sizes.
To assess the statistical signicance of the calculated distances between
the expression patterns of all genes within a trait, random background
distributions are created for (1) individual genes and (2) traits of N genes.
For individual genes, three different distributions were calculated, based
on the distances between randomly sampled open reading frames,
randomly sampled genes with an annotation (but not necessarily the same
annotation), and randomly sampled genes with the same annotation. The
background distribution for a trait of N genes is based on the distances
between randomly composed sets of genes. For each gene pair, two
distances metrics are calculated, the Pearson Correlation (PC) and the
Normalized Rank Euclidean Distance (NRED). In practice, it is often found
that a certain annotation is assigned to multiple genes in the same
genome. If this occurs, there is an option to use either a random selection,
or the highest scoring pair. In the latter case, a correction for multiple
testing is implemented. This process is repeated N-times, where N equals
the number of genes in any given trait. The background distribution for
traits is determined by rst randomly sampling two genomes, identifying
the overlap in annotations, and nally articially dening a trait containing
N annotations. For each annotation in the trait, the distances are calculated
between genome A and genome B, as described in the previous section.
As modules vary in size, this process is repeated for traits of different sizes.
Identifying attributes. TbasCO provides both a cluster-based and pair-wise
approach to identify attributes. In both methods, the distance between
expression patterns of a trait between two genomes is rst calculated
based on a composite Z score of the PC and NRED for each gene
composing the trait. In the cluster-based analysis, the distances are
subsequently clustered using the Louvain clustering algorithm to identify
trait attributes. To determine if an attribute is signicantly similar or not, a
one-sided T-test between the attribute and the random background
distribution of traits is conducted. This is done for both cluster-based and
model-based comparisons. Many traits are complex and represented in
databases such as KEGG by numerous alternative routes. To deal with this
complexity, each pathway is expanded into all possible alternative routes.
Due to the extremely high number of alternative routes for some traits,
attributes are pruned based on a strict requirement of 100% completion.
Distance calculations. To determine the similarity in expression patterns
between genes, two dissimilarity metrics are calculated: the PC between
RNAseq counts across samples, and the NRED, where ranks are a measure
of relative abundance of RNA in each sample, normalized the abundance
of RNA in the corresponding genome. These distance scores are converted
to Z scores using a background distribution of distances between
randomly sampled genes as previously described. To determine statisti-
cally signicant similarities in expression patterns of a trait, a composite
score is calculated. For each of these genes the PC and NRED are calculated
and transformed to Z scores and combined as (1*PC +NRED). The
distance of the trait between two genomes is dened as the average of
these composite distance scores. If traits being compared do not have
100% overlap in gene content, then the dissimilarity score is normalized by
the Jaccard distance between gene content of the trait.
PC þNREDðÞ1dJðÞ
Statistical assessment of trait attributes. In both model-based and pair-
wise approaches, the distance is rst calculated between expression
patterns of a trait between two genomes based on the composite Z score
of the PC and NRED for each gene composing the trait. In the clustering-
based analysis, the distances are subsequently clustered using the Louvain
clustering algorithm to identify trait-attributes. To determine if attributes
are signicantly similar, a one-sided T-test is conducted between the
attribute and a background distribution of randomly sampled traits with
the same number of genes. To derive the random background
distributions, multiple distributions are calculated ranging in gene
numbers from the smallest trait to the largest trait in the dataset as
described previously. For each background distribution, N (default: 10,000)
traits are randomly composed. The distances between these articial traits
are calculated in the same way as for the actual traits. In addition to a
statistical pruning step, the attributes are pruned based on a strict
requirement of 100% completion of each module. A benchmarking
analysis to examine the effects of different parameters, including the
presence of zero counts, was conducted to determine their inuence on
the number of attributes identied and may be found in the
supplementary materials (Supplementary Table 1, Supplementary
Figs. 24).
RESULTS AND DISCUSSION
Reconstructing a diverse EBPR SBR community
To explore trait-based transcriptional dynamics of a semi-complex
microbial community, we applied genome-resolved metage-
nomics and metatranscriptomics to an EBPR sequencing-batch
reactor (SBR) ecosystem (Fig. 2). We previously performed a
metatranscriptomics time-series experiment over the course of a
normally operating EBPR cycle to investigate the regulatory
controls of Accumulibacter gene expression [42]. In this experi-
ment, six samples were collected for RNA sequencing: three from
the anaerobic phase and three from the aerobic phase (Fig. 2A).
Additionally, three metagenomes were collected from the same
month of the metatranscriptomic experiment, including a sample
from the same date of the experiment. We reassembled
contemporary Accumulibacter clade IIA and IA genomes that
were previously assembled from the same bioreactor system
[27,28]. The genomes of Accumulibacter clades IA and IIA are
similar by approximately 85% average-nucleotide identity [28,31],
which is well below the common species-resolved cutoff of 95%,
and these groups have recently been designated as separate
species (Candidatus Accumulibacter regalis and Candidatus
Accumulibacter phosphatis, respectively) [35]. However, we
maintain references to the Accumulibacter clade nomenclature
E.A. McDaniel et al.
9
ISME Communications
based on polyphosphate kinase (ppk1) sequence identity through-
out the manuscript (CAPIA and CAPIIA) [31,50,51]. During the
experiment, the bioreactor was highly enriched in Accumulibacter
clade IIA, accounting for approximately 50% of the mapped
metagenomic reads and the highest transcriptional counts (Fig. 2B,
C) [42]. Whereas Accumulibacter clade IA exhibited low abun-
dance patterns but was within the top 10 genomes with the
highest total transcriptional counts (Fig. 2C).
Although this bioreactor system was highly enriched in
Accumulibacter, a diverse bacterial community persisted and
was active in this ecosystem (Fig. 2B, C). We reconstructed
representative population genomes of the microbial community
of the SBR system, resulting in 64 metagenome-assembled
genomes (MAGs) of the (non-Accumulibacter) bacterial commu-
nity. Interestingly, we recovered genomes of experimentally
veried and putative PAOs previously not detected in these
bioreactors, including two Tetrasphaera spp. (TET1 and TET2)
Candidatus Obscuribacter phosphatis(OBS1), and Gemmatimona-
detes (GEMMA1). Pure cultures of Tetrasphaera have been
experimentally shown to cycle polyphosphate without incorpor-
ating PHA [37], deviating from the hallmark Accumulibacter PAO
model. The rst cultured representative of the Gemmatimonadetes
phylum Gemmatimonas aurantiaca was isolated from an SBR
simulating EBPR and was shown to accumulate polyphosphate
through Neisser and DAPI staining [52]. Additionally, Ca.
Obscuribacter phosphatis has been hypothesized to cycle phos-
phorus based on the presence of genes for phosphorus transport,
polyphosphate incorporation, and potential for both anaerobic
and aerobic respiration [38], and was enriched in a photobior-
eactor EBPR system [53]. Both Tetrasphaera spp. TET1 and TET2,
OBS1, and GEMMA1 groups exhibit higher relative abundance
patterns than CAPIA but have similar relative transcriptional levels
(Fig. 2B, C, Table 1).
Numerous SBR MAGs among the Actinobacteria and Proteobac-
teria contain the high-afnity phosphorus transporter pstABCS
system, polyphosphate kinase ppk1, and the low-afnity pit
phosphorus transporter (Supplementary Fig. 5). Additionally,
select MAGs within the Alphaproteobacteria,Betaproteobacteria,
and Gammaproteobacteria contain all required subunits for
polyhydroxyalkanoate synthesis (Supplementary Fig. 5). Other
abundant and transcriptionally active groups in the SBR ecosys-
tem that are not predicted to be PAOs are members of the
Fig. 2 Genome-resolved metatranscriptomics approach of an EBPR system. Application of a genome-resolved metatranscriptomics
approach to a lab-scale sequencing batch reactor (SBR) designed to enrich for Accumulibacter. ASchematic of the main cycle parameters and
analyte dynamics of an SBR simulating EBPR. Six samples were taken for RNA sequencing within the cycle at time-points denoted by arrows.
BPhylogenetic identity and abundance patterns of 66 assembled MAGs from the EBPR system. The phylogenetic tree was constructed from
concatenated markers contained in the GTDB-tk with muscle, calculated with RAxML, and visualized in iTOL. A phylogenetic tree of all 66
MAGs with reference genomes and high-quality genomes from Singleton et al. constructed with concatenated markers from GTDB-tk are
provided in Supplementary Fig. 1. Sizes of circles represent relative abundance patterns calculated from metagenomic reads obtained from a
sample collected the same day as the metatranscriptomic experiment was performed, and are not to scale. CTranscriptional patterns of each
MAG in the anaerobic and aerobic phases of the EBPR cycle. RNA-seq reads from each time-point were competitively mapped to all 66
assembled MAGs and counts normalized by transcripts per million (TPM). Total counts in the anaerobic and aerobic phases for each genome
were averaged separately and plotted on a log scale. Order of MAGs from left to right mirrors the order of MAGs in the phylogenetic tree in
Bfrom the top of the circle going clockwise.
E.A. McDaniel et al.
10
ISME Communications
Bacteroidetes such as CHIT1 within the Chitinophagaceae, and
Cytophagales members Runella sp. RUN1 and Leadbetterella sp.
LEAD1 (Fig. 2B, C, Table 1). Interestingly, an uncharacterized group
within the Bacteroidetes, represented by BAC1, contributed the
third most to the pool of transcripts (Fig. 2C), and did not show
phylogenetic similarity to MAGs assembled from Danish full-scale
wastewater treatment systems [40] (Supplementary Fig. 1). Other
groups from which we assembled MAGs for that do not exhibit
clear roles in EBPR systems were Chloroexi ANAER1 and HERP1
MAGs, Armatimonadetes FIMBRI1, Firmicutes FUSI1, and Patesci-
bacteria SACCH1. Members of the Chloroexi are lamentous
bacteria that have been associated with bulking and foaming
events in full-scale WWTPs [5456], but also aid in forming the
scaffolding around oc aggregates and degrade complex poly-
mers [5658]. The Patescibacteria (formerly TM7) are widespread
but low abundant members of natural and engineered ecosys-
tems, have reduced genome sizes, and may contribute to
lamentous bulking in activated sludge [22,59]. To summarize,
lab-scale SBRs designed to enrich for Accumulibacter contain
diverse bacterial microorganisms [27,32], but their ecological
functions and putative interactions remain to be fully understood
in the context of the EBPR ecosystem.
Identifying expression-based trait attributes among the EBPR
SBR community with TbasCO
Current metatranscriptomics analyses often employ either a gene-
centric [31,6062] or genome-centric approach [42,6365]. In
both approaches, highly, differentially, or co-expressed genes are
identied and tested for enrichment of specic functions.
Enrichment- or annotation-based approaches are employed in
numerous metatranscriptomics tools such as MG-RAST, MetaTrans,
SAMSA2, COMAN, IMP, and Anvio[6671]. Here, we expand on
the use of molecular markers as traits by dening expression
attributes by leveraging a priori knowledge from predened trait
libraries, such as the KEGG database [72], to statistically assess
inter-species expression patterns of genes that together form a
trait (Fig. 1). First, our results showed that there is statistically
signicant transcriptional conservation of genes at the community
level; genes that share an annotation were signicantly more
similar than expected using two different distance metrics (NRED:
pvalue <2.2e16, PC: pvalue <2.2e16). Extending this statistical
analysis to the trait level, we identied 1674 attributes distributed
across the 66 genomes. On average, we identied 9.12 genomes
per attribute (SD -5.22), with a minimum of 3 genomes and a
maximum of 35 (Fig. 3B). Based on these statistics, we dened
redundant attributes as those two standard deviations above the
mean (19 genomes). With this cutoff applied, we identied 79
redundant trait attributes mostly belonging to pathways among
carbohydrate metabolism, purine metabolism, and fatty acid
metabolism categories (Table 2). Of 290 traits, we identied 97
traits with two or more attributes identied (33%). Of these, traits
in 10 or more genomes were twice as likely to have two or more
attributes (65%), suggesting that divergent expression patterns for
a trait are common, and may represent a niche-differentiating
feature (Fig. 3A). Henceforth, when multiple attributes are
identied for a trait, we refer to these as niche-differentiating
attributes.
From the ecosystem perspective, a clear phylogenetic signal is
observed in the distribution of attributes, as genomes cluster
together by shared trait attributes by phylum with some
exceptions, such as genomes belonging to the Bacteroidetes,
Actinobacteria, and Proteobacteria clustering together, respectively
(Fig. 3C). For simplicity, we ltered the network to only include
nodes with more than 5 connections. Highly redundant trait
attributes belonged to modules in the lipid metabolism, energy
metabolism, and nucleotide metabolism KEGG functional cate-
gories. In contrast, more specialized trait attributes on the
periphery of the network or amongst group-specic clusters such
as within the Actinobacteria or subsets of the Proteobacteria
belonged to amino acid metabolism, biosynthesis of terpenoids
and polyketides, metabolism of cofactors and vitamins, and
carbohydrate metabolism KEGG modules. Pathways of note that
showed a high level of redundancy include the TCA cycle,
isoleucine biosynthesis, acyl-CoA synthesis, threonine biosynth-
esis, and cytochrome c oxidase activity (Table 2). Large pathways
with hundreds of possible routes such as glycolysis, the TCA cycle,
gluconeogenesis, and the pentose phosphate pathway are not
included in the main network and are displayed as individual
networks (Supplementary Fig. 6).
We next explored the distribution of non-redundant attributes
(e.g. 318 genomes) (Fig. 3B). A total of 796 trait attributes with
low redundancy were identied belonging to pathways involved
in carbohydrate cofactor and vitamin metabolism including
glycolysis, gluconeogenesis, parts of the TCA cycle, tetrahydrofo-
late biosynthesis, tryptophan biosynthesis, and the pentose
phosphate pathway (Table 3). Different sets of low redundancy
trait attributes were identied within respective phyla (Supple-
mentary Fig. 7). Between genomes belonging to the Actinobac-
teria,Alphaproteobacteria, Bacteroidetes, Betaproteobacteria, and
Gammaproteobacteria, low redundancy attributes (belonging to
less than half of the total genomes within the phylum) include
carbohydrate metabolism, amino acid metabolism and metabo-
lism of cofactors and vitamins (Supplementary Fig. 7). Redundant
trait attributes within individual phyla belong to core energy
metabolism pathways, fatty acid biosynthesis, and carbohydrate
metabolism. However, even within individual phyla, non-
redundant attributes include different amino acids and cofactors
(Extended Table 1 - available on Figshare https://gshare.com/
articles/dataset/Lineage-Specic_Core_and_Niche_Differentiating
_Traits/15001200).
As noted previously, one of the most striking ndings is that a
majority, 65% of traits present in 10 or more genomes have multiple
expression attributes. Thus, it seems that while the presence of
marker genes suggests many organisms share a particular trait, the
presence of niche-differentiating expression proles suggest an
alternative story, that there is a level of hidden metabolic diversity.
For example, central carbon metabolism and energy pathways such
as the TCA cycle, glycolysis, gluconeogenesis, and the pentose
phosphate pathway are oftentimes considered core traits when
only analyzing the presence and/or absence of individual markers
belonging to these pathways. Among over 1000 high-quality MAGs
assembled from full-scale Danish WWTPs, the TCA cycle and
pentose phosphate pathway are highly represented among the
abundant microorganisms, with glycolysis less so [40]. Whereas the
TCA cycle and pentose phosphate pathway are present among a
high number of genomes in the EBPR SBR community, different
routes or parts of these pathways have niche-differentiating
distributions (Supplementary Fig. 6, Tables 2and 3). These ner-
scale differences in expression of coretraits may explain the
persistence of a diverse community when solely fed acetate, as
different lineages could employ similar carbon utilization pathways
differently or in more versatile ways. Another salient aspect of this
analysis is the astonishingly high number of possible routes within
individual pathways here represented by their Disjunctive Normal
Forms. For example, accounting for all alternative routes and
enzymes, the glycolysis pathway has 100 s of possible routes.
Layering upon this many expression attributes reveals a large
hidden metabolic versatility.
Dimensionality of the high-afnity phosphorus transporter
system PstABCS
The EBPR ecosystem is characterized by its highly dynamic
phosphorus cycles. To explore how different lineages respond to
uctuating phosphorus concentrations, we examined the
expression-based attributes for the KEGG module of the high-
afnity phosphorus transporter pstABCS (Fig. 4). The pstABCS
E.A. McDaniel et al.
11
ISME Communications
system is an ABC-type transporter that strongly binds phosphate
with high afnity under phosphorus-limiting conditions, and
therefore we expected that the highest expression levels would be
at the end of the aerobic cycle [73]. In contrast, we found that
pstABCS expression was characterized by two different trait
attributes. In the rst attribute shared by 14 community members,
all pstABCS components displayed the highest activity towards the
end of the aerobic cycle, when phosphorus concentrations were
depleted (Fig. 4, Attribute 1). Conversely, 11 community members
displayed an alternate attribute where the highest activity of
pstABCS was at the transition from anaerobic to aerobic phases
when phosphorus concentrations are highest (Fig. 4, Attribute 2).
Interestingly, the two Accumulibacter clades IA and IIA are split
amongst these separate pstABCS attributes. These results are in
agreement with previous results showing that Accumulibacter
clade IIC has a canonical pstABCS expression pattern (as in Fig. 4,
Attribute 1), whereas the Accumulibacter clade IA has a non-
canonical expression (as in Fig. 4, Attribute 2) [31]. By assigning
trait attributes, we can extend these ndings beyond Accumuli-
bacter to other community members in the SBR ecosystem
suggesting that there are conserved ecological pressures driving
niche differentiating expression patterns in pstABCS within the
EBPR community.
Distribution and expression of truncated denitrication steps
among EPBR community members
Denitrication gene induction is an important ecosystem property
linked to the redox status of an environment. In EBPR
Fig. 3 Clustering and distribution of trait attributes across EBPR SBR community members. Using the TbasCO method, we identied
expression-based trait attributes from predened trait modules in the KEGG library and explored the distribution of these trait attributes
across community members. ADistribution of trait-attributes among sets of genomes. Bars represent the number of trait-attributes present in
a set number of genomes and colored by KEGG module category. Among a total of 35 genomes, trait attributes present between 3 and 18
genomes are designated as niche differentiating, whereas trait attributes present in 19 or greater genomes are designated as core trait
attributes. Inset gure demonstrates the maximum number of attributes for the maximum number of genomes. BCytoscape network
showing the connectedness of genomes to trait attributes. The network was ltered to only include nodes with more than 5 connections,
therefore ltering out both genomes with few trait attributes and trait attributes connected to less than 5 genomes. Genomes are represented
as squares colored by phylum, and trait attributes are represented as circles colored by KEGG category. The size of both the squares and circles
represents the number of connections to that genome or trait attribute, respectively.
E.A. McDaniel et al.
12
ISME Communications
Table 2. KEGG pathways for core trait-attributes present in greater than 19 genomes.
Module description Number of attributes
Citrate cycle, second carbon oxidation, 2-oxoglutarate => oxaloacetate [PATH:map00020 map01200
map01100]
13
Citrate cycle (TCA cycle, Krebs cycle) [PATH:map00020 map01200 map01100] 10
Shikimate pathway, phosphoenolpyruvate +erythrose-4P => chorismate [PATH:map00400 map01230
map01100 map01110]
8
Fatty acid biosynthesis, initiation [PATH:map00061 map01212 map01100] 7
Glycolysis, core module involving three-carbon compounds [PATH:map00010 map01200 map01230
map01100]
7
Adenine ribonucleotide biosynthesis, IMP => ADP,ATP [PATH:map00230 map01100] 4
Guanine ribonucleotide biosynthesis IMP => GDP,GTP [PATH:map00230 map01100] 4
Inosine monophosphate biosynthesis, PRPP +glutamine => IMP [PATH:map00230 map01100] 4
Isoleucine biosynthesis, threonine => 2-oxobutanoate => isoleucine [PATH:map00290 map01230
map01100]
3
NADH:quinone oxidoreductase, prokaryotes [PATH:map00190] 3
beta-Oxidation, acyl-CoA synthesis [PATH:map00061 map00071 map01212 map01100] 2
F-type ATPase, prokaryotes and chloroplasts [PATH:map00190 map00195] 2
Valine/isoleucine biosynthesis, pyruvate => valine / 2-oxobutanoate => isoleucine [PATH:map00290
map00770 map01210 map01230 map01100 map01110]
2
CAM (Crassulacean acid metabolism), dark [PATH:map00620 map00710 map01200 map01100
map01120]
1
Cytochrome c oxidase, cbb3-type [PATH:map00190] 1
Cytochrome c oxidase, prokaryotes [PATH:map00190] 1
dTDP-L-rhamnose biosynthesis [PATH:map00521 map00523 map01100 map01130] 1
Leucine biosynthesis, 2-oxoisovalerate => 2-oxoisocaproate [PATH:map00290 map01210 map01230
map01100 map01110]
1
Phosphatidylethanolamine (PE) biosynthesis, PA =>PS=> PE [PATH:map00564 map01100] 1
PRPP biosynthesis, ribose 5 P => PRPP [PATH:map00030 map00230 map01200 map01230 map01100] 1
Pyruvate oxidation, pyruvate => acetyl-CoA [PATH:map00010 map00020 map00620 map01200
map01100]
1
Semi-phosphorylative Entner-Doudoroff pathway, gluconate => glycerate-3P [PATH:map00030
map01200 map01100 map01120]
1
Threonine biosynthesis, aspartate => homoserine => threonine [PATH:map00260 map01230
map01100 map01110]
1
Table 3. KEGG Pathways for differentiating trait-attributes present between 3 and 18 genomes.
Module_description Number of attributes
Glycolysis (Embden-Meyerhof pathway), glucose => pyruvate [PATH:map00010 map01200
map01100]
279
Citrate cycle (TCA cycle, Krebs cycle) [PATH:map00020 map01200 map01100] 208
Gluconeogenesis, oxaloacetate => fructose-6P [PATH:map00010 map00020 map01100] 76
Inosine monophosphate biosynthesis, PRPP +glutamine => IMP [PATH:map00230 map01100] 45
Citrate cycle, second carbon oxidation, 2-oxoglutarate => oxaloacetate [PATH:map00020
map01200 map01100]
31
Heme biosynthesis, plants and bacteria, glutamate => heme [PATH:map00860 map01100
map01110]
27
Tetrahydrofolate biosynthesis, GTP => THF [PATH:map00790 map00670 map01100] 25
Tryptophan biosynthesis, chorismate => tryptophan [PATH:map00400 map01230 map01100
map01110]
25
Ornithine biosynthesis, glutamate => ornithine [PATH:map00220 map01210 map01230 map01100] 24
Histidine biosynthesis, PRPP => histidine [PATH:map00340 map01230 map01100 map01110] 17
Pentose phosphate pathway (Pentose phosphate cycle) [PATH:map00030 map01200 map01100
map01120]
16
Lysine biosynthesis, succinyl-DAP pathway, aspar tate => lysine [PATH:map00300 map01230
map01100]
12
Uridine monophosphate biosynthesis, glutamine (+PRPP) => UMP [PATH:map00240 map01100] 11
E.A. McDaniel et al.
13
ISME Communications
communities, we nd many genomes with diverse and incom-
plete denitrication pathways, distributed across many lineages
denitrication steps expected in denitrifying systems (Fig. 5).
Among all 66 MAGs, we did not identify any single MAG with a
complete denitrication pathway consisting of the genetic
repertoire necessary to fully reduce nitrate to nitrogen gas
(Supplementary Fig. 5). Instead, we identied multiple groups of
organisms with truncated denitrication pathways, with steps
distributed among cohorts of community members (Fig. 5).
For the rst steps of reducing nitrate to nitrite, we examined
expression attributes of the napAB and narGH pathways (Fig. 5B,
C). For the narGH pathway, two attributes were identied (Fig. 5B).
The rst narGH attribute was characterized by high expression in
the anaerobic phase, with decreasing transcript levels by the
second time point of the anaerobic phase. Genomes containing
this attribute included the experimentally veried and putative
PAOs Tetrasphaera (TET1 and TET2) and Ca. Obscuribacter (OBS1),
respectively. The second attribute was exhibited among members
of the Actinobacteria (PROP2, PHYC2, PROP3, and NANO1),
Proteobacteria (BEIJ4), and Bacteroidetes (BAC1). The attribute
identied for napAB was also more highly expressed anaerobically
and included CAPIA, CAPIIA, ALIC1, REYR2, RUBRI1, and BEIJ3.
Interestingly, this napAB attribute had expression patterns that
quickly decreased in the rst aerobic time point, suggesting a
tighter regulation than Attribute 1 for narGH. Together, this
suggests that the regulation of denitrication within the EBPR
ecosystem is a niche-differentiating feature whereby the induction
of denitrication pathways occurs either anaerobically or only
after anaerobic carbon contact.
A smaller cohort contained the genetic repertoire to reduce
nitrite to nitrogen gas and exhibited hallmark anaerobic-aerobic
expression patterns (Fig. 5E) These members within the Proteo-
bacteria (OTTO2, BEIJ3, VITREO1, and ZOO1) contained the nirS
nitrite reductase, the norBC nitric oxide reductase, and nosZ, and
showed highest expression of these subunits towards the
beginning of the anaerobic cycle, slowly decreasing over the
aerobic period to their lowest in the end of the aerobic cycle.
Although BEIJ2 was lacking the norBC system, it contained the nirS
nitrite reductase and nosZ subunit, and exhibited similar expres-
sion patterns to others in this cohort. Other Proteobacteria
Fig. 4 Trait attributes of the high-afnity phosphorus transporter system pstABCS.Using the TbasCO method, two trait attributes of the
high-afnity phosphorus transporter system pstABCS were identied. The pstABCS system consists of a phosphate-binding protein and ABC-
type transporter, and the corresponding KEGG orthologs for each subunit are shown. Timepoints 13 refer to the three anaerobic phase
timepoints, and timepoints 46 refer to the three anaerobic phase timepoints (Fig. 1). Expression values are log-transformed based on setting
the lowest expression value within each genome across the time-series to 0 for each subunit. Specic subunits for some genomes in both
attributes are missing to the high cutoff thresholds for annotations. However we kept genomes with 2/4 subunits to show similarities in
expression proles. The rst pstABCS trait-attribute includes microbial lineages that exhibited the highest expression of all subunits towards
the end of the aerobic cycle, when phosphate concentrations are expected to be lowest. This includes microbial lineages within the
Actinobacteria, Proteobacteria, Gemmatimonadetes, and Chloroexi. The second pstABCS trait-attribute includes lineages that exhibited highest
expression of all subunits upon the switch from anaerobic to aerobic phases, or when phosphate concentrations are expected to be the
highest. This includes lineages within the Actinobacteria and Proteobacteria.
E.A. McDaniel et al.
14
ISME Communications
Fig. 5 Expression dynamics of distributed denitrication routes. Expression of denitrication traits distributed among community members
in the EBPR SBR ecosystem. Timepoints 13 correspond to the anaerobic phase and timepoints 46 correspond to the aerobic phase as
referenced in Fig. 1.AComplete denitrication pathway and associated genetic repertoire with each sequential step. BTrait attributes of
expression dynamics for community members with the narGH nitrate reductase system. This trait was the only denitrication trait identied
with more than one attribute. CExpression dynamics of the napAB nitrate reductase system. DExpression dynamics of the norBC nitrous oxide
reductase system. EExpression of all steps of denitrication starting at nitrite reduction. FExpression of the most complete denitrication
route among three community members, with the norC subunit for nitrous oxide reduction missing. Note that OTTO1 only contains nirS but is
included in this trait attribute because the expression dynamics are similar to that of the other three genomes for this subunit.
E.A. McDaniel et al.
15
ISME Communications
lineages only contained the norBC subunits but were expressed in
similar fashions (RHODO2, FLAVO1, RHIZO1, and LEAD1) (Fig. 5D).
Accumulibacter clades IA and IIA as well as ALIC1 were the only
lineages with near-complete denitrication pathways. These
lineages contained the napAB nitrate reductase system as
mentioned above, the nirS nitrite reductase, norB (missing a
condent hit for the norC subunit), and nosZ. These three lineages
also exhibited hallmark upregulation of all steps in the anaerobic
phase, with decreased activity after aerobic contact (Fig. 5F).
Interestingly, Accumulibacter clade IA exhibited a higher level
of transcripts associated denitrication steps when expression
levels were normalized relative to clade IIA, supporting the
hypothesis that denitrication is a niche-differentiating feature
among clades [28,31,74], and possibly a strain-specic trait since
denitrication traits cannot be predicted based on ppk1 clade
designations [32]. For example, independent observations in
differences among denitrication activities among strains within
Accumulibacter clade IC are inconsistent [34,75]. Within the same
bioreactor environment, coexisting Accumulibacter clades differ
between denitrication abilities and expression proles [3133].
Truncated denitrication pathways have also been previously
shown to be distributed among community members, with the
complete denitrication genetic repertoire only present in few
members [32,33], which could be due to extensive horizontal
Fig. 6 Biosynthetic potential compared to expression of amino acid and vitamin synthesis pathways for top 15 expressed MAGs.
Biosynthetic potential and expression patterns of amino acid and vitamin pathways were analyzed for the top 15 genomes with the highest
transcriptional counts (Table 1). AFor a pathway to be considered present for downstream analysis in the TbasCO pipeline, 80% of the
pathway had to be present in a genome. Thus, we used this cutoff criterion to discern whether a specic pathway was present or absent in a
genome (with the expectation of methionine, as all genomes did not contain at least 80% of the subunits in the KEGG methionine synthase
pathway, we inferred the presence of the methionine synthase as presence of this pathway). Orange colored boxes for cofactor biosynthesis
pathways represents the presence of that pathway, whereas grey infers absence. For amino acid biosynthetic pathways, amino acids are listed
by their side chain groups charged, polar, hydrophobic, and other. Blue colored boxes for amino acid biosynthesis pathways represents the
presence of that pathway, whereas grey infers absence. BMini-networks of vitamin co-factors. Squares are genomes with the colors matching
the color bar in A. Nodes are attributes, where the colored nodes for the tetrahydrofolate attributes represent the different routes. CMini-
networks of amino acid biosynthesis pathways split by type. Colors of nodes for each amino acid represent the different routes for that
pathway. Squares represent genomes with colors matching the color bar in A.
E.A. McDaniel et al.
16
ISME Communications
gene transfer of genes comprising denitrication steps [32,76].
Although this experiment was not conducted under denitrifying
conditions, our approach could be applied to denitrifying EBPR
systems to further understand the distribution of denitrication
traits among community members and how to selectively enrich
for diverse DPAOs.
Biosynthetic potential and expression dynamics of amino acid
and vitamin synthesis pathways
Although SBRs are designed to enrich for Accumulibacter by
providing acetate as the sole carbon source, a diverse bacterial
community persists in these setups [27,32]. One hypothesis for
the persistence of these bacterial community members may be
cooperative interactions due to underlying auxotrophies of amino
acid and vitamin biosynthetic pathways in Accumulibacter. Amino
acids and vitamin cofactors are metabolically expensive to
synthesize, and widespread auxotrophies have been widely
documented among microbial communities [77,78]. Specically,
auxotrophies of vitamin cofactors have been shown to fuel
bacterial and cross-kingdom interactions with de novo synthesi-
zers [79,80]. To explore this hypothesis in the EPBR SBR
community, we analyzed the presence of amino acid and vitamin
biosynthetic pathways and their expression patterns among the
top 15 genomes based on transcript abundance (Fig. 6).
Within Accumulibacter, there are a few key vitamin cofactor and
amino acid auxotrophies that could fuel potential interactions
with other community members. Both Accumulibacter clade
genomes are missing the riboavin pathway for FAD cofactor
synthesis, as well as known pathways for serine and aspartic acid
(Fig. 6A). The biosynthetic pathway for aspartic acid is distributed
among members of the Bacteroidetes and Proteobacteria, whereas
only TET2 contains the pathway for serine synthesis (Fig. 5A). The
lack of serine biosynthesis pathways in Accumulibacter and other
genomes seems striking given that serine is one of the least
metabolically costly amino acids to synthesize [81]. Interestingly,
Accumulibacter clade IIA (strain CAPIIA) does not contain the
biosynthetic machinery for thiamine and pantothenate synthesis,
whereas clade IA (strain CAPIA) does (Fig. 6A). Only the CAULO1,
HYPHO1, and PSEUDO1 genomes within the Proteobacteria can
synthesize thiamine, whereas several other members can synthe-
size pantothenate (Fig. 6A). The absence of the pantothenate
biosynthetic pathway in Accumulibacter CAP IIA is particularly
interesting given that coenzyme A is essential for polyhydroxyalk-
anote biosynthesis, which fuels the rapid and extensive polymer
cycling PAO phenotype of Accumulibacter [24].
In addition to other community members potentially support-
ing the growth of Accumulibacter due to underlying auxotrophies,
the reciprocal logic may be possible as well. Both Accumulibacter
clades contain the pathways for synthesizing tyrosine and
phenylalanine, which are missing in a majority of the top 15
active non-Accumulibacter bacterial genomes (Fig. 6A). Only two
other members within the Proteobacteria can synthesize tyrosine
and phenylalanine, where RAM1 can synthesize both and
PSEUDO1 only phenylalanine. Interestingly, phenylalanine and
tyrosine are the second and third most metabolically expensive
amino acids to synthesize, respectively, with tryptophan being the
most costly [81]. Additionally, a few highly active non-
Accumulibacter bacterial community members lack the biosyn-
thetic machinery for several vitamin cofactors and amino acids,
such as FLAVO1 and BAC3 within the Bacteroidetes and the
putative PAO Ca. Obscuribacter phosphatis OBS1 (Fig. 6A).
Particularly, RAM1 within the Proteobacteria is missing the
biosynthetic machinery for all vitamin cofactors but can synthesize
most amino acids including the most metabolically expensive as
mentioned above.
We next analyzed the distribution of trait-attributes of vitamin
and amino acid pathways among these genomes to understand
how these biosynthetic pathways are expressed similarly or
differently in the EBPR SBR ecosystem (Fig. 6B, C). Members of
the Proteobacteria containing thiamine and cobalamin biosyn-
thetic pathways all express these traits similarly (Fig. 6B). However,
the pantothenate synthesis pathway contains two trait-attributes
and is expressed differently among two cohorts. In the rst
attribute, RUN1, TET1, CAULO1, CAPIA, and PSEUDO1 express the
pantothenate pathway similarly. However, OBS1 and TET2 express
the pantothenate pathway differently (Fig. 6B). Because tetrahy-
drofolate can be synthesized through different metabolic routes,
we analyzed the differences in trait attribute expression for all
routes in genomes that contained sufcient coverage of this trait.
Bacteroidetes and Proteobacteria members mostly cluster together
among tetrahydrofolate attributes, whereas the TET1 and TET2
genomes are differentiated (Fig. 6B).
Expression of various groups of amino acids show more
differentiated expression patterns for genomes with these path-
ways. Several amino acids also contain different metabolic routes
for biosynthesis, and we analyzed all trait attributes for each
amino acid for all routes grouped by type (Fig. 6C). For the
charged amino acids arginine, histidine, and lysine, Proteobacteria
and Bacteroidetes members cluster within their phylogenetic
groups, respectively, with lysine and histidine expressed differ-
ently among these groups (Fig. 6C). In contrast, arginine is
expressed similarly among all Proteobacteria genomes. Among the
polar charged amino acids, TET2 is the only genome among the
top 15 genomes that contains the pathway to synthesize serine
(Fig. 6A). Several groups contain the pathway for threonine
synthesis, and expression of different threonine routes are
differentiated among the Proteobacteria, Bacteroidetes, and Tetra-
sphaera spp., though they mostly cluster phylogenetically (Fig. 6C).
Notably, the expression patterns for the cysteine and proline
biosynthetic pathways do not cluster phylogenetically, such as
both Tetrasphaera genomes expressing the proline pathway more
similarly to other Proteobacteria and Bacteroidetes (Fig. 6C). The
few lineages that can synthesize tyrosine and phenylalanine
(CAPIA, CAPIIA, RAM1, PSEUDO1) show different expression
patterns. These results show that beyond the presence or absence
of key vitamin cofactor and amino acid biosynthetic pathways,
EBPR SBR organisms also display coherent and differentiated
expression patterns for these traits, of which the functional
consequences remain to be further understood.
CONCLUSIONS AND FUTURE PERSPECTIVES
In this work, we applied a novel trait-based omics pipeline to a
semi-complex, engineered bioreactor microbial community to
explore ecosystem-level and niche-differentiating traits. Through
recovering 66 MAGs from the EBPR SBR community and using a
time-series metatranscriptomics experiment, we were able to extend
functional predictions such as identifying multiple attributes of high-
afnity phosphate transporters beyond hypotheses made from traits
alone. We extended this framework to other signicant traits that
are distributed among community members such as denitrication
and amino acid metabolism. Specically, we demonstrate that traits
with similar expression proles may be clustered into attributes
providing a new layer to trait-based approaches.
We believe that identifying expression-based attributes will be a
powerful tool to explore microbial traits in natural, engineered, and
host-associated microbiomes. Outside of activated sludge systems,
trait-based approaches could illuminate how similar secondary
metabolite clusters are expressed among different species in a
community [82,83], how auxotrophies for amino acid and vitamin
cofactors govern interactions [84], how rhizosphere microorganisms
respond to day-night cycles, and identify putative traits that
universally exhibit ecosystem-level or niche-differentiating patterns
across ecosystems [19,23]. Importantly, our trait-based approach
can be used to screen for expected expression patterns of a key trait
compared to a model organism, and then prioritize specic
E.A. McDaniel et al.
17
ISME Communications
microbial lineages for downstream experimental verication with
techniques such as Raman-FISH [85,86].
DATA AVAILABILITY
All supplementary les including functional annotations and transcriptome count les
are available at https://gshare.com/projects/EBPR_Trait-Based_Comparative_Omics/
90437. All 64 genomes have been deposited in NCBI at Bioproject PRJNA714686.The
remainingtwo reassembled Accumulibacter genomes have not been deposited in NCBI
to not confuse between the original CAPIA and CAPIIA assemblies [27,28]. These
contemporary assemblies are available at the Figshare repository. The three
metagenomes and six metatranscriptomes used in this study are available on the
JGI/IMG at accession co des 3300026302, 3300026286, 3300009517, and 33000 02341-46 ,
respectively. All code for performing metagenomic assembly, binning, and annotation
can be found at https://github.com/elizabethmcd/EBPR-MAGs. The TbasCO method has
been implemented as a reproducible R package and can be accessed at https://
github.com/Jorisvansteenbrugge/TbasCO.
REFERENCES
1. Violle C, Navas M-L, Vile D, Kazakou E, Fortunel C, Hummel I, et al. Let the concept
of trait be functional! Oikos. 2007;116:88292.
2. Lavorel S, Garnier E. Predicting changes in community composition and eco-
system functioning from plant traits: revisiting the Holy Grail. Funct Ecol.
2002;16:54556.
3. Hooper DU, Chapin FS, Ewel JJ, Hector A, Inchausti P, Lavorel S, et al. Effects of
biodiversity on ecosystem functioning: A consensus of current knowledge. Ecol
Monogr. 2005;75:335.
4. Pianka ER. On r-and K-selection. The American Naturalist, Vol. 104 (Nov. - Dec.,
1970), pp. 592-597.
5. Wright IJ, Reich PB, Westoby M, Ackerly DD, Baruch Z, Bongers F, et al. The
worldwide leaf economics spectrum. Nature. 2004;428:8217.
6. Krause S, Le Roux X, Niklaus PA, Van Bodegom PM, Lennon JT, Bertilsson S, et al.
Trait-based approaches for understanding microbial biodiversity and ecosystem
functioning. Front Microbiol. 2014;5:251.
7. Malik, A.A., Martiny, J.B.H., Brodie, E.L. et al. Dening trait-based microbial stra-
tegies with consequences for soil carbon cycling under climate change. ISME J.
2020;14:19. https://doi.org/10.1038/s41396-019-0510-0
8. Guittar J, Shade A, Litchman E. Trait-based community assembly and succession
of the infant gut microbiome. Nat Commun. 2019;10:512.
9. Wolfe BE, Button JE, Santarelli M, Dutton RJ. Cheese rind communities provide
tractable systems for in situ and in vitro studies of microbial diversity. Cell.
2014;158:42233.
10. Enke TN, Datta MS, Schwartzman J, Barrere J, Pascual-García A, Cordero OX.
modular assembly of polysaccharide-degrading marine microbial communities.
Curr Biol. 2019;29:152835.
11. Herrera Paredes S, Gao T, Law TF, Finkel OM, Mucyn T, Teixeira PJPL, et al. Design
of synthetic bacterial communities for predictable plant phenotypes. PLOS Biol.
2018;16:e2003962.
12. Lindemann SR, Bernstein HC, Song H-S, Fredrickson JK, Fields MW, Shou W, et al.
Engineering microbial consortia for controllable outputs. ISME J. 2016;10:207784.
13. Oyserman BO, Medema MH, Raaijmakers JM. Road MAPs to engineer host
microbiomes. Curr Opin Microbiol. 2018;43:4654.
14. Lawson CE, Harcombe WR, Hatzenpichler R, Lindemann SR, Löfer FE, OMalley
MA, et al. Common principles and best practices for engineering microbiomes.
Nat Rev Microbiol. 2019;17:72541. Nature Publishing Group.
15. Gutierrez CF, Sanabria J, Raaijmakers JM, Oyserman BO Restoring degraded
microbiome function with self-assembled communities. FEMS Microbiol Ecol.
2020;96:aa225.
16. Allison SD. A trait-based approach for modelling microbial litter decomposition.
Ecol Lett. 2012;15:105870.
17. Tringe SG, von Mering C, Kobayashi A, Salamov AA, Chen K, Chang HW, et al.
Comparative metagenomics of microbial communities. Science. 2005;308:5547.
18. Tyson GW, Chapman J, Hugenholtz P, Allen EE, Ram RJ, Richardson PM, et al.
Community structure and metabolism through reconstruction of microbial
genomes from the environment. Nature. 2004;428:3743.
19. Anantharaman K, Brown CT, Hug LA, Sharon I, Castelle CJ, Probst AJ, et al.
Thousands of microbial genomes shed light on interconnected biogeochemical
processes in an aquifer system. Nat Commun. 2016;7:13219.
20. Woodcroft BJ, Singleton CM, Boyd JA, Evans PN, Emerson JB, Zayed AAF, et al.
Genome-centric view of carbon processing in thawing permafrost. Nature.
2018;560:4954.
21. McDaniel EA, Wahl SA, Ishii S, Pinto A, Ziels R, Nielsen PH, et al. Prospects for
multi-omics in the microbial ecology of water engineering. Water Res. 2021; 205.
22. Albertsen M, Hugenholtz P, Skarshewski A, Nielsen KL, Tyson GW, Nielsen PH.
Genome sequences of rare, uncultured bacteria obtained by differential coverage
binning of multiple metagenomes. Nat Biotechnol. 2013;31:5338.
23. Wrighton KC, Thomas BC, Sharon I, Miller CS, Castelle CJ, VerBerkmoes NC, et al.
Fermentation, hydrogen, and sulfur metabolism in multiple uncultivated bac-
terial phyla. Science. 2012;337:16615.
24. Hesselmann RPX, Werlen C, Hahn D, van der Meer JR, Zehnder AJB. Enrichment,
phylogenetic analysis and detection of a bacterium th atperforms enhanced biological
phosphate removal in activated sludge. Syst Appl Microbiol. 1999;22:45465.
25. Seviour RJ, Mino T, Onuki M. The microbiology of biological phosphorus removal
in activated sludge systems. FEMS Microbiol Rev. 2003;27:99127.
26. Crocetti GR, Hugenholtz P, Bond PL, Schuler A, Ju¨ J, Keller J, et al. Identication of
polyphosphate-accumulating organisms and design of 16S rRNA-directed probes
for their detection and quantitation. Appl Environ Microbiol. 2000;66:117582.
27. Martín HG, Ivanova N, Kunin V, Warnecke F, Barry KW, McHardy AC, et al.
Metagenomic analysis of two enhanced biological phosphorus removal (EBPR)
sludge communities. Nat Biotechnol. 2006;24:12639.
28. Flowers JJ, He S, Malfatti S, del Rio TG, Tringe SG, Hugenholtz P, et al. Com-
parative genomics of two Candidatus Accumulibacterclades performing bio-
logical phosphorus removal. ISME J. 2013;7:230114.
29. Oyserman BO, Moya F, Lawson CE, Garcia AL, Vogt M, Heffernen M, et al.
Ancestral genome reconstruction identies the evolutionary basis for trait
acquisition in polyphosphate accumulating bacteria. ISME J. 2016;10:293145.
30. Wilmes P, Andersson AF, Lefsrud MG, Wexler M, Shah M, Zhang B, et al. Community
proteogenomics highlights microbial strain-variantproteinexpressionwithinactivated
sludge performing enhanced biological phosphorus removal. ISME J. 2008;2:85364.
31. McDaniel EA, Moya-Flores F, Keene Beach N, Camejo PY, Oyserman BO, Kizaric M,
et al. Metabolic Differentiation of Co-occurring Accumulibacter Clades Revealed
through Genome-Resolved Metatranscriptomics. mSystems. 2021;6:47495.
32. Gao H, Mao Y, Zhao X, Liu WT, Zhang T, Wells G. Genome-centric metagenomics
resolves microbial diversity and prevalent truncated denitrication pathways in a
denitrifying PAO-enriched bioprocess. Water Res. 2019;155:27587.
33. Wang Y, Gao H, F Wells G. Integrated omics analyses reveal differential gene
expression and potential for cooperation between denitrifying polyphosphate
and glycogen accumulating organisms. Environ Microbiol. 2021;23:327493.
34. Camejo PY, Oyserman BO, McMahon KD, Noguera DR. Integrated omic analyses
provide evidence that a candidatus accumulibacter phosphatisstrain performs
denitrication under microaerobic conditions. mSystems. 2019;4:e0019318.
35. Petriglieri F, Singleton CM, Kondrotaite Z, Dueholm MKD, McDaniel EA, McMahon
KD, et al. Reevaluation of the Phylogenetic Diversity and Global Distribution of
the Genus Candidatus Accumulibacter. mSystems. 2022;7:e00016-22.
36. Kong Y, Nielsen JL, Nielsen PH. Identity and ecophysiology of uncultured acti-
nobacterial polyphosphate-accumulating organisms in full-scale enhanced bio-
logical phosphorus removal plants. Appl Environ Microbiol. 2005;71:407685.
37. Kristiansen R, Nguyen HTT, Saunders AM, Nielsen JL, Wimmer R, Le VQ, et al. A
metabolic model for members of the genus Tetrasphaera involved in enhanced
biological phosphorus removal. ISME J. 2013;7:54354.
38. Soo R, Skennerton CT, Sekiguchi Y, Imelfort M, Paech S, Dennis P, et al. An
expanded genomic representation of the phylum cyanobacteria. Genome Biol-
ogy Evolut. https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4040986/. Accessed
11 Jul 2020.
39. Petriglieri F, Singleton C, Peces M, Petersen JF, Nierychlo M, Nielsen PH. Can-
didatus Dechloromonas phosphoritrophaand Ca. D. phosphorivorans, novel
polyphosphate accumulating organisms abundant in wastewater treatment
systems. ISME J. 2021;15:360514. 2021 1512
40. Singleton CM, Petriglieri F, Kristensen JM, Kirkegaard RH, Michaelsen TY, Ander-
sen MH, et al. Connecting structure to function with the recovery of over 1000
high-quality metagenome-assembled genomes from activated sludge using
long-read sequencing. Nat Commun. 2021;12:2009.
41. Singleton CM, Petriglieri F, Wasmund K, Nierychlo M, Kondrotaite Z, Petersen JF,
et al. The novel genus, Candidatus Phosphoribacter, previously identied as
Tetrasphaera, is the dominant polyphosphate accumulating lineage in EBPR
wastewater treatment plants worldwide. ISME J. 2022 2022; 112.
42. Oyserman BO, Noguera DR, del Rio TG, Tringe SG, McMahon KD. Metatran-
scriptomic insights on gene expression and regulatory controls in Candidatus
Accumulibacter phosphatis. ISME J. 2016;10:81022.
43. Parks DH, Imelfort M, Skennerton CT, Hugenholtz P, Tyson GW. CheckM: assessing
the quality of microbial genomes recovered from isolates, single cells, and
metagenomes. Genome Res. 2015;25:104355.
44. Kanehisa M, Sato Y, Morishima K. BlastKOALA and GhostKOALA: KEGG tools for
functional characterization of genome and metagenome sequences. J Mol Biol.
2016;428:72631.
45. Aramaki T, Blanc-Mathieu R, Endo H, Ohkubo K, Kanehisa M, Goto S, et al.
KofamKOALA: KEGG ortholog assignment based on prole HMM and adaptive
score threshold. Bioinformatics. 2020;36:2251-2.
E.A. McDaniel et al.
18
ISME Communications
46. Bushnell B, Rood J, Singer E. BBMerge Accurate paired shotgun read merging
via overlap. PLoS One. 2017;12:e0185056.
47. Kopylova E, Noé L, Touzet H. SortMeRNA: fast and accurate ltering of ribosomal
RNAs in metatranscriptomic data. Bioinformatics. 2012;28:32117.
48. Bray NL, Pimentel H, Melsted P, Pachter L. Near-optimal probabilistic RNA-seq
quantication. Nat Biotechnol. 2016;34:5257.
49. Soneson C, Love MI, Robinson MD. Differential analyses for RNA-seq: transcript-
level estimates improve gene-level inferences. F1000Research. 2015;4:1521.
50. He S, Gall DL, McMahon KD. Candidatus accumulibacterpopulation structure in
enhanced biological phosphorus removal sludges as revealed by polyphosphate
kinase genes. Appl Environ Microbiol. 2007;73:586574.
51. Camejo PY, Owen BR, Martirano J, Ma J, Kapoor V, Santo Domingo J, et al.
Candidatus Accumulibacter phosphatis clades enriched under cyclic anaerobic
and microaerobic conditions simultaneously use different electron acceptors.
Water Res. 2016;102:12537.
52. Zhang H, Sekiguchi Y, Hanada S, Hugenholtz P, Kim H, Kamagata Y, et al. Gemmati-
monas aurantiaca gen. nov., sp. nov., a Gram-negative, aerobic, polyphosphate-
accumulating micro-organism, the rst cultured representative of the new bacterial
phylum Gemmatimonadetes phyl. nov. Int J Syst Evol Microbiol. 2003;53:115563.
53. McDaniel EA, Wever R, Oyserman BO, Noguera DR, McMahon KD Genome-
resolved metagenomics of a photosynthetic bioreactor performing biological
nutrient removal. Microbiol Resour Announc. 2021;10:e00244-21.
54. Speirs LBM, Rice DTF, Petrovski S, Seviour RJ. The phylogeny, biodiversity, and
ecology of the chloroexi in activated sludge. Front Microbiol. 2019;10:2015.
Frontiers Media S.A.
55. Andersen MH, McIlroy SJ, Nierychlo M, Nielsen PH, Albertsen M. Genomicinsights into
Candidatus Amarolinea aalborgensis gen. nov., sp. nov., associated with settleability
problems in wastewater treatment plants. Syst Appl Microbiol. 2019;42:7784.
56. Nierychlo M, Miłobȩdzka A, Petriglieri F, McIlroy B, Nielsen PH, McIlroy SJ. The
morphology and metabolic potential of the Chloroexi in full-scale activated
sludge wastewater treatment plants. FEMS Microbiol Ecol. 2019;95:y228.
57. McIlroy SJ, Karst SM, Nierychlo M, Dueholm MS, Albertsen M, Kirkegaard RH, et al.
Genomic and in situ investigations of the novel uncultured Chloroexi associated with
0092 morphotype lamentous bulking in activated sludge. ISME J. 2016;10:222334.
58. Kragelund C, Levantesi C, Borger A, Thelen K, Eikelboom D, Tandoi V, et al.
Identity, abundance and ecophysiology of lamentous Chloroexi species pre-
sent in activated sludge treatment plants. FEMS Microbiol Ecol. 2007;59:67182.
59. Kindaichi T, Yamaoka S, Uehara R, Ozaki N, Ohashi A, Albertsen M, et al. Phylo-
genetic diversity and ecophysiology of Candidate phylum Saccharibacteria in
activated sludge. FEMS Microbiol Ecol. 2016;92:111.
60. Mann E, Wetzels SU, Wagner M, Zebeli Q, Schmitz-Esser S. Metatranscriptome
sequencing reveals insights into the gene expression and functional potential of
rumen wall bacteria. Front Microbiol. 2018;9:43.
61. Jiang Y, Xiong X, Danska J, Parkinson J. Metatranscriptomic analysis of diverse
microbial communities reveals core metabolic pathways and microbiome-
specic functionality. Microbiome. 2016;4:2.
62. Linz AM, Aylward FO, Bertilsson S, McMahon KD Time-series metatranscriptomes
reveal conserved patterns between phototrophic and heterotrophic microbes in
diverse freshwater systems. Limnol Oceanogr. 2019;65:10112.
63. Lawson CE, Wu S, Bhattacharjee AS, Hamilton JJ, McMahon KD, Goel R, et al.
Metabolic network analysis reveals microbial community interactions in ana-
mmox granules. Nat Commun. 2017;8:15416.
64. Aylward FO, Eppley JM, Smith JM, Chavez FP, Scholin CA, DeLong EF. Microbial
community transcriptional networks are conserved in three domains at ocean
basin scales. Proc Natl Acad Sci. 2015;112:54438.
65. Hao L, Michaelsen TY, Singleton CM, Dottorini G, Kirkegaard RH, Albertsen M,
et al. Novel syntrophic bacteria in full-scale anaerobic digesters revealed by
genome-centric metatranscriptomics. ISME J. 2020;14:90618.
66. Glass EM, Meyer F. The metagenomics RAST server: a public resource for the
automatic phylogenetic and functional analysis of metagenomes. Handb Mol
Microb Ecol I Metagenomics Complement Approaches. 2011;8:32531.
67. Martinez X, Pozuelo M, Pascal V, Campos D, Gut I, Gut M, et al. MetaTrans: an
open-source pipeline for metatranscriptomics. Sci Rep. 2016;6:26447.
68. Westreich ST, Treiber ML, Mills DA, Korf I, Lemay DG. SAMSA2: a standalone
metatranscriptome analysis pipeline. BMC Bioinformatics. 2018;19:175.
69. Ni Y, Li J, Panagiotou G. COMAN: a web server for comprehensive metatran-
scriptomics analysis. BMC Genomics. 2016;17:622.
70. Narayanasamy S, Jarosz Y, Muller EEL, Heintz-Buschart A, Herold M, Kaysen A,
et al. IMP: a pipeline for reproducible reference-independent integrated meta-
genomic and metatranscriptomic analyses. Genome Biol. 2016;17:260.
71. Eren AM, Esen ÖC, Quince C, Vineis JH, Morrison HG, Sogin ML, et al. Anvio: an
advanced analysis and visualization platform for omics data. PeerJ. 2015;3:e1319.
72. Kanehisa M, Sato Y, Kawashima M, Furumichi M, Tanabe M. KEGG as a reference
resource for gene and protein annotation. Nucleic Acids Res. 2016;44:D457D462.
73. Wanner BL. Gene regulation by phosphate in enteric bacteria. J Cell Biochem.
1993;51:4754.
74. Flowers JJ, He S, Yilmaz S, Noguera DR, McMahon KD. Denitrication capabilities
of two biological phosphorus removal sludges dominated by different Candi-
datus Accumulibacterclades. Environ Microbiol Rep. 2009;1:5838.
75. Rubio-Rincón FJ, Weissbrodt DG, Lopez-Vazquez CM, Welles L, Abbas B, Albertsen
M, et al. Candidatus Accumulibacter delftensis: A clade IC novel polyphosphate-
accumulating organism without denitrifying activity on nitrate. Water Res.
2019;161:13651.
76. Parsons C, Stüeken EE, Rosen CJ, Mateos K, Anderson RE Radiation of nitrogen-
metabolizing enzymes across the tree of life tracks environmental transitions in
Earth history. Geobiology. 2021;1:1834.
77. Gómez-Consarnau L, Sachdeva R, Gifford SM, Cutter LS, Fuhrman JA, Sañudo-
Wilhelmy SA, et al. Mosaic patterns of B-vitamin synthesis and utilization in a
natural marine microbial community. Environ Microbiol. 2018;20:280923.
78. Hamilton JJ, Garcia SL, Brown BS, Oyserman BO, Moya-Flores F, Bertilsson S, et al.
Metabolic Network Analysis and Metatranscriptomics Reveal Auxotrophies and
Nutrient Sources of the Cosmopolitan Freshwater Microbial Lineage acI. mSys-
tems. 2017;2:e0009117.
79. McClure RS, Overall CC, Hill EA, Song H-S, Charania M, Bernstein HC, et al. Species-
specic transcriptomic network inference of interspecies interactions. ISME J.
2018;1:201123.
80. CroftMT,LawrenceAD,Raux-DeeryE,Warren MJ, Smith AG. Algae acquire vitamin
B12 through a symbiotic relationship with bacteria. Nature. 2005;438:9093.
81. Akashi H, Gojobori T. Metabolic efciency and amino acid composition in the proteomes
of Escherichia coli and Bacillus subtilis. Proc Natl Acad Sci USA. 2002;99:36953700.
82. Lozano GL, Bravo JI, Diago MFG, Park HB, Hurley A, Peterson SB, et al. Introducing
THOR, a model microbiome for genetic dissection of community behavior. MBio.
2019;10:e0284618.
83. Crits-Christoph A, Diamond S, Buttereld CN, Thomas BC, Baneld JF. Novel soil bacteria
possess diverse genes for secondary metabolite biosynthesis. Nature. 2018;558:4404.
84. Zengler K, Zaramela LS. The social network of microorganismshow auxo-
trophies shape complex communities. Nat Rev Microbiol. 2018;16:38390.
85. Fernando EY, McIlroy SJ, Nierychlo M, Herbst FA, Petriglieri F, Schmid MC, et al.
Resolving the individual contribution of key microbial populations to enhanced
biological phosphorus removal with RamanFISH. ISME J. 2019;13:193346.
86. Petriglieri F, Petersen JF, Peces M, Nierychlo M, Hansen K, Baastrand CE, et al. Quan-
tication of biologically and chemically bound phosphorus in activated sludge from
full-scale plants with biological P-removal. Environ Sci Technol. 2022;56:513240.
87. Chaumeil P-A, Mussig AJ, Hugenholtz P, Parks DH GTDB-Tk: a toolkit to classify
genomes with the Genome Taxonomy Database. Bioinformatics.
2019;36:192527.
88. Seemann T Prokka: Rapid prokaryotic genome annotation. Bioinformatics.
2014;30:206869.
ACKNOWLEDGEMENTS
We thank Caitlin Singleton for providing early access to high-quality genomes from a
full-scale WWTP to compare our MAGs against. Metagenomic and metatranscrip-
tomic sequencing was provided through a Joint Genome Institute Community
Science Proposal (Proposal ID 873). This work was supported by funding from the
National Science Foundation (MCB-1518130) to K.D.M and D.R.N. Funding was
provided to E.A.M. by a fellowship through the Department of Bacteriology at the
University of Wisconsin Madison. Funding for B.O.O was in part provided by the
Technology Foundation of the Dutch National Science Foundation (NWO-TTW). This
research was performed in part using the Wisconsin Energy Institute computing
cluster, which is supported by the Great Lakes Bioenergy Research Center as a part of
the U.S. Department of Energy Ofce of Science (DE-SC0018409).
AUTHOR CONTRIBUTIONS
EAM and JJMVS contributed equally to this work. EAM performed metagenomic
assembly and genome curation, metatranscriptomic and functional analysis, and
software testing. JJMVS and BOO designed the TbasCO analysis software and JJMVS
performed benchmarking testing. DRN, KDM, JMR, and MHM provided critical
feedback on the manuscript. EAM, JJMVS, and BOO performed analyses, interpreta-
tion, and wrote the manuscript with input from all coauthors.
COMPETING INTERESTS
The authors declare no competing interests.
E.A. McDaniel et al.
19
ISME Communications
ADDITIONAL INFORMATION
Supplementary information The online version contains supplementary material
available at https://doi.org/10.1038/s43705-022-00189-2.
Correspondence and requests for materials should be addressed to E. A. McDaniel,
J. J. M. van Steenbrugge or B. O. Oyserman.
Reprints and permission information is available at http://www.nature.com/
reprints
Publishers note Springer Nature remains neutral with regard to jurisdictional
claims in published maps and institutional afliations.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative
Commons license, and indicate if changes were made. The images or other third party
material in this article are included in the articles Creative Commons license, unless
indicated otherwise in a credit line to the material. If material is not included in the
articles Creative Commons license and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly
from the copyright holder. To view a copy of this license, visit http://
creativecommons.org/licenses/by/4.0/.
© The Author(s) 2022
E.A. McDaniel et al.
20
ISME Communications
Article
Full-text available
Recovering nitrogen (N) from municipal wastewater is a promising approach to prevent nutrient pollution, reduce energy use, and transition toward a circular N bioeconomy, but remains a technologically challenging endeavor. Existing N recovery techniques are optimized for high-strength, low-volume wastewater. Therefore, developing methods to concentrate dilute N from mainstream wastewater will bridge the gap between existing technologies and practical implementation. The N-rich biopolymer cyanophycin is a promising candidate for N bioconcentration due to its pH-tunable solubility characteristics and potential for high levels of accumulation. However, the cyanophycin synthesis pathway is poorly explored in engineered microbiomes. In this study, we analyzed over 3,700 publicly available metagenome assembled genomes (MAGs) and found that the cyanophycin synthesis gene cphA was ubiquitous across common activated sludge bacteria. We found that cphA was present in common phosphorus accumulating organisms (PAO) Ca. ‘Accumulibacter’ and Tetrasphaera, suggesting potential for simultaneous N and P bioconcentration in the same organisms. Using metatranscriptomic data, we confirmed the expression of cphA in lab-scale bioreactors enriched with PAO. Our findings suggest that cyanophycin synthesis is a ubiquitous metabolic activity in activated sludge microbiomes. The possibility of combined N and P bioconcentration could lower barriers to entry for N recovery, since P concentration by PAO is already a widespread biotechnology in municipal wastewater treatment. We anticipate this work to be a starting point for future evaluations of combined N and P bioaccumulation, with the ultimate goal of advancing widespread adoption of N recovery from municipal wastewater.
Preprint
Full-text available
Recovering nitrogen (N) from municipal wastewater is a promising approach to prevent nutrient pollution, reduce energy use, and transition towards a circular N bioeconomy, but remains a technologically challenging endeavor. Existing N recovery techniques are optimized for high-strength, low-volume wastewater. Therefore, developing methods to concentrate dilute N from mainstream wastewater will bridge the gap between existing technologies and practical implementation. The N-rich biopolymer cyanophycin is a promising candidate for N bioconcentration and recovery due to its solubility characteristics and potential for high levels of accumulation in a limited number of bacterial isolates. However, the cyanophycin synthesis pathway is poorly explored in natural and engineered microbiomes. In this study, we analyzed over 3700 publicly available metagenome assembled genomes (MAGs) and found that the cyanophycin synthesis gene cphA was ubiquitous across common activated sludge bacteria. Surprisingly, we found that cphA was present in all analyzed genomes of the common phosphorus accumulating organisms (PAO) Ca. 'Accumulibacter' and Tetrasphaera, suggesting potential for simultaneous N and P bioconcentration in the same organisms. Using metatranscriptomic data, we also confirmed the expression of cphA in lab-scale bioreactors enriched with PAO. Our findings suggest that cyanophycin synthesis is a ubiquitous metabolic pathway in activated sludge microbiomes and therefore may have potential for integration in existing biological nutrient removal and recovery processes. We anticipate this work to be a starting point for future evaluations of combined N and P bioaccumulation, with the ultimate goal of advancing widespread adoption of nutrient recovery from municipal wastewater.
Article
Full-text available
Common culturing techniques and priorities bias our discovery towards specific traits that may not be representative of microbial diversity in nature. So far, these biases have not been systematically examined. To address this gap, here we use 116,884 publicly available metagenome-assembled genomes (MAGs, completeness ≥80%) from 203 surveys worldwide as a culture-independent sample of bacterial and archaeal diversity, and compare these MAGs to the popular RefSeq genome database, which heavily relies on cultures. We compare the distribution of 12,454 KEGG gene orthologs (used as trait proxies) in the MAGs and RefSeq genomes, while controlling for environment type (ocean, soil, lake, bioreactor, human, and other animals). Using statistical modeling, we then determine the conditional probabilities that a species is represented in RefSeq depending on its genetic repertoire. We find that the majority of examined genes are significantly biased for or against in RefSeq. Our systematic estimates of gene prevalences across bacteria and archaea in nature and gene-specific biases in reference genomes constitutes a resource for addressing these issues in the future.
Article
Full-text available
“ Candidatus Accumulibacter” is the most studied PAO, with a primary role in biological nutrient removal. However, the species-level taxonomy of this lineage is convoluted due to the use of different phylogenetic markers or genome sequencing approaches. Here, we redefined the phylogeny of these organisms, proposing a comprehensive approach which could be used to address the classification of other diverse and uncultivated lineages.
Article
Full-text available
Phosphorus (P) is present in activated sludge from wastewater treatment plants in the form of metal salt precipitates, extracellular polymeric substances, or bound into the biomass, for example, as intracellular polyphosphate (poly-P). Several methods for a reliable quantification of the different P-fractions have recently been developed, and this study combines them to obtain a comprehensive P mass-balance of activated sludge from four enhanced biological phosphate removal (EBPR) plants. Chemical characterization by ICP-OES and sequential P fractionation showed that chemically bound P constituted 38-69% of total P, most likely in the form of Fe, Mg, or Al minerals. Raman microspectroscopy, solution state 31P NMR, and 31P MAS NMR spectroscopy applied before and after anaerobic P-release experiments, were used to quantify poly-P, which constituted 22-54% of total P and was found in approximately 25% of all bacterial cells. Raman microspectroscopy in combination with fluorescence in situ hybridization was used to quantify poly-P in known polyphosphate-accumulating organisms (PAO) (Tetrasphaera, Candidatus Accumulibacter, and Dechloromonas) and other microorganisms known to possess high level of poly-P, such as the filamentous Ca. Microthrix. Interestingly, only 1-13% of total P was stored by unidentified PAO, highlighting that most PAOs in the full-scale EBPR plants investigated are known.