Skip to main content

Efficient management of the nitritation-anammox microbiome through intermittent aeration: absence of the NOB guild and expansion and diversity of the NOx reducing guild suggests a highly reticulated nitrogen cycle

Abstract

Obtaining efficient autotrophic ammonia removal (aka partial nitritation-anammox, or PNA) requires a balanced microbiome with abundant aerobic and anaerobic ammonia oxidizing bacteria and scarce nitrite oxidizing bacteria. Here, we analyzed the microbiome of an efficient PNA process that was obtained by sequential feeding and periodic aeration. The genomes of the dominant community members were inferred from metagenomes obtained over a 6 month period. Three Brocadia spp. genomes and three Nitrosomonas spp. genomes dominated the autotrophic community; no NOB genomes were retrieved. Two of the Brocadia spp. genomes lacked the genomic potential for nitrite reduction. A diverse set of heterotrophic genomes was retrieved, each with genomic potential for only a fraction of the denitrification pathway. A mutual dependency in amino acid and vitamin synthesis was noted between autotrophic and heterotrophic community members. Our analysis suggests a highly-reticulated nitrogen cycle in the examined PNA microbiome with nitric oxide exchange between the heterotrophs and the anammox guild.

Introduction

Predicting and managing the composition and function of microbial communities is heralded as the holy grail of microbial ecology [1]. Composition can be imposed when working under aseptic conditions, but is challenging when communities are open, the typical scenario for most ecology-relevant and technology-interesting microbial communities. While rational engineering of microbiomes is an active field of research, it is recognized that we still lack the required mechanistic understandings to design microbiomes based on a priori knowledge [2].

On the other hand, heuristic approaches based on manipulation of environmental conditions, have a long history of success in managing microbial communities towards specific functions in food, agriculture, and environmental applications. While those management practices do not rely on first principle understanding of microbial community assembly, unravelling those communities, and seeking links with management constraints might yield insights on which to build new testable theories [2].

The common, and intuitively simple, conditions that one can impose on a microbial community are the provision/fluxes of specific electron donors versus electron acceptors and of macro-or micronutrients; provision or limitation of these will impose selective pressures and enrich communities with the desired phenotypes. In addition, community control may be further facilitated by spatio-temporal gradients that naturally occur or are artificially imposed. Gradients are especially valuable when they permit the establishment of multiple redox conditions over short spatial or temporal scales that allow the co-occurrence of functional groups that require mutually exclusive environmental conditions; the characteristic of microbial aggregates and biofilms [3, 4].

Here, we examine the management of a microbial community with the functional property of complete autotrophic conversion of ammonia (NH3) to dinitrogen gas (N2), also known as the partial nitritation/anammox process (PNA) [5]. The simplest functional PNA community would consist of only two functional groups: one that performs aerobic NH3 oxidation (using O2 as terminal electron acceptor (TEA) and forming NO2, aerobic ammonia oxidizing prokaryotes or AOBs) and one that performs anoxic NH3 oxidation (using NO2 as TEA and forming N2, anaerobic ammonia oxidizing prokaryotes or AnAOBs). The PNA process relies on the provision of controlled (limited) supply of oxygen and the spatial or temporal variation in redox conditions [6, 7]. Yet, the PNA process remains tested by microbial community management, especially the suppression of the aerobic NO2 oxidation guild (aerobic nitrite oxidizing prokaryotes or NOBs) [7, 8]. Indeed, all metagenetic and metagenomic analyses of PNA communities to date indicate the persistent presence of NOBs; and their excessive presence would deteriorate or collapse the PNA process [8, 9]. Selection for AOB against NOBs has variously relied on growth inhibition by free ammonia or nitrous acid [10] or oxygen limitation driven by the low O2 affinity of the NOB [9, 11,12,13]. In addition, while autotrophic NH3 and NO2 oxidizers are the only microbes that can grow on the influent devoid of organic carbon and are essential for a functional PNA process, their metabolism and decay will result in release of organic byproducts. As a result, heterotrophic microbes are inevitable and their abundance and potential symbiosis with autotrophs has previously been found [14, 15], yet their functional contribution is ill-documented.

We document here that periodic and limiting provision of oxygen (via intermittent aeration) to a granule-based reactor can result in complete elimination of NOBs from the community performing the PNA process. We examine the consequences of this operation on the community structure, with specific attention on alternate pathways for NOx (nitrogen oxides) metabolism. We identified a representative set of high-quality MAGs representing most of the metagenome, a few highly abundant AOB and AnAOB, the absence of NOB, and a diversity in NOx respiratory abilities across the heterotrophic MAGs. Genome analysis suggests a highly reticulated network with possibility for NO exchange between autotrophs and heterotrophs and strong evidence of auxotrophies distributed across the community members.

Results

Process performance

Prior to this experimental phase, the reactor had been operating for nearly 19 months, and the last five months using a consistent regime that included three aerated and non-aerated intervals (30% and 70% of the react phase respectively): total nitrogen (TN) removal efficiency was around 84.5% [16]. During the reported phase, the frequency of redox switching increased on a monthly basis from 3 to 4, 6, 8, 10, 16 and 25 per cycle. Throughout these six months of operation, process performance deteriorated slightly with removal rates varying from 86 ± 4% TN in early phase and reaching 75 ± 4% at highest switching frequency (increasing oxygen loading recovered 85% efficiency). No residual nitrite was ever detected in the effluent and the ratio of nitrate produced per unit of ammonium removed (RNatTot [17]) remained below 0.13, pointing at AnAOB as the only generators of NO3 [18]. Further details are reported elsewhere [16].

Community analysis via nxr and 16S rRNA gene targeted qPCR [16] and 16S rRNA gene amplicon sequencing (Additional file 1: Fig. S1) indicated extremely low abundances of taxa containing typical nitrite oxidizing bacteria (predominantly in smallest aggregates (< 90 µm)) Nitrobacter spp. and Nitrospira spp., but high abundance of taxa comprising AnAOB (Brocadia spp.) and AOB (Nitrosomonas spp.). AOB fractions increased with redox switching frequency, but only in the largest aggregates (> 600 µm) [16].

Quality of MAGs

Whole-community DNA sequencing from seven samples, taken at monthly intervals, from the PNA reactor, generated an average of 2.7 ± 0.5 Gbp high-quality, paired-end sequence data per sample. A total of 55 metagenome-assembled genomes (MAGs) (average completeness and contamination of 81.5% and 1.7%, respectively (Additional file 2: Table S1)) were retrieved from the co-assembled contigs. On average 79% (± 1) of the community metagenome could be assigned to the recovered MAGs indicating a solid coverage of the community by the retrieved MAGs. Temporal dynamics of the community, as inferred from relative MAG abundance, was limited, indicating resilience to the changing aeration conditions (Fig. 1).

Fig. 1
figure 1

Overview of the recovered MAGs, their relative abundance, dynamics and phylogenetic placement

Overall community composition

MAGs were classified as autotrophs if they contained the key genes of any of the previously described carbon fixation pathways (the Calvin-Benson-Bassham (CBB) cycle, the 3-hydroxypropionate cycle, the 3-hydroxypropionate-4-hydroxybutyrate cycle, the reductive tricarboxylic acid (rTCA) cycle or the Wood-Ljungdahl pathway). The community was dominated by non-autotrophic (46 MAGs, ca. 57% abundance) over autotrophic MAG types (9 MAGs, ca. 43% abundance) even though no organic carbon was fed to the system (Additional file 3:Table S2; Fig. 1). The autotrophic MAGs comprised three MAGs each in the AOB (reductive pentose phosphate (CBB)) and AnAOB (Wood-Ljungdahl (WL)) guilds, respectively, and three additional MAGs PRO3, PRO5 and CFX7 carry the CBB and WL pathways, respectively.

The ammonia oxidizing guilds

Of the core functional guilds, the AOB guild comprises three different MAGs; while AOB2 had an AAI of 97% with Nitrosomonas europaea, AOB1 and AOB3 are sufficiently divergent from known genomes to be separate species with Nitrosomonas eutropha and Nitrosomonas europaea as closest relatives (Additional file 1: Fig. S2). Although differential binning was unable to assign the amo and hao operons to the appropriate MAGs (see explanation in Additional file 1: Fig. S3), all AOB MAGs have the genes for the ammonia monooxygenase complex (amoCABDE) and hydroxylamine oxidase (hao), and typical Copper resistance/homeostasis (copCD) genes. Genes for NOx reduction were also present: The AOB MAGs contain a Copper-containing nitrite reductase (nirK) and cytochrome bc-type complex cNOR (respiratory nitric oxide reductase, norBC) (Additional file 3: Table S2).

The AnAOB comprises three different MAGs: AMX1 had an AAI of 99% with Candidatus Brocadia fulgida; AMX3 had an AAI of 94% with Candidatus Brocadia sp. UTAMX1; AMX2 constitutes a new species, with Candidatus Brocadia sp. UTAMX2 as closest relative (Additional file 1: Fig. S2). UTAMX1 and UTAMX 2 were reported as dominant AnAOB in a similar study [19].

All AnAOB MAGs harbor the hydrazine dehydrogenase (hdh), hydrazine synthase (hzsABC) and nitrate oxidoreductase (nxrAB) genes (Additional file 3: Table S2). A gene encoding nitrite reductase (nirK), the typical enzyme converting NO2 to NO was present in AMX1, but absent from both AMX2 and AMX3 (Additional file 3: Table S2). Besides these genes, multiple copies of hao-like genes were present in all AMX MAGs: ten in AMX1 and AMX2 and six in AMX3 (Additional file 1: Fig. S4). Phylogenetic analysis indicates at least eight hao clusters congruent with published hao. Both AMX1 and AMX2 harbor three homologous copies of a hao-like gene previously associated with hydroxylamine oxidation to nitric oxide [19], while AMX3 contains one copy. On the other hand, each of the AMX MAGs have three homologous copies of a hao-like gene hypothesized to be involved in nitrite reduction to either nitric oxide or hydroxylamine [19] (Additional file 1: Fig. S4). No amoA-like sequences were found that could be assigned to non-autotrophic MAGs; however some haoA-like sequences were assigned to non-AOB MAGs (PRO3, PRO5, PRO11); PRO3 has two haoA gene paralogues – and both have as closest relative a gene found in Lautropia SCN 69-89, previously identified as abundant in nitration/anammox communities[20] and suggested to be a nitrifier-denitrifier.

About half of the recovered MAGs (28/55) harbor genes homologous to nxr/nar (Additional file 3: Table S2) comprising both the cytoplasmic and periplasmic NxrA/NarG encoding operons (Additional file 1: Fig S5). All of those genes were phylogenetically distinct from those in previously characterized nitrite oxidizing bacteria, however, seven MAGs (PRO3, PRO6, PRO11, PRO12, CFX 1, CFX9, ARM1) encode a periplasmic Nxr that belong to the NOB and AnAOB phyletic group (Additional file 1: Fig. S5). Whether these nxr-like genes encode for proteins involved in nitrite oxidation or nitrate reduction remains unknown. Except for PRO3, no MAG contains both a nxr operon and a carbon fixation pathway, suggesting the absence of canonical NOBs, but possibility for non-autotrophic nitrite oxidation as noted for an ARM genome retrieved from a anammox microbiome [21].

Even though the recovered MAGs represented approximately 80% of the whole metagenome, we also screened the metagenomic reads to detect possible canonical nxr missed during the assembly or binning process. Some reads mapped against canonical nxr, especially from Nitrobacter spp., although the number was much lower compared to the reads mapping to AnAOB nxr (3.0 ± 0.8 reads per million (RPM) vs. 100.2 ± 24.6 RPM). An even lower number of reads mapped to nxr from other known NOB (Nitrolancea spp.: 0.7 ± 0.2 RPM; Nitrospira spp.: 0.3 ± 0.2 RPM; and 0 RPM to Candidatus Nitrotoga spp. and Nitrospinae spp.) (Additional file 1: Table S3).

The heterotrophic guilds–NOx respiration

As more than half of the metagenome was heterotrophic (Fig. 1, Additional file 3: Table S2) it was examined in further detail, especially for its respiratory abilities towards nitrogen oxides (Fig. 2).

Fig. 2
figure 2

Distribution of nitrogen cycling genes across the MAGs

Almost half of the MAGs (23/55) encode genes for respiratory nitrate reductase (narGHIJ, 28/55), one encodes the periplasmic nitrate reductase gene (napAB), some carry the genes for dissimilatory nitrite reduction to ammonia (DNRA) via the periplasmic nitrite reductase (nrfHA; 12/55) or the cytoplasmic nitrite reductase (nirBD; 2/55) (Additional file 3: Table S2), although complete DNRA (carriage of nitrate reductases gene in addition to nrfHA or nirBD)) was found in just 5 MAGs (ACD2, CFX8, CFX11, CLB2 and CLB3) (Additional file 3: Table S2). If the above identified nxr-like genes (in ARM1 and CFX1) encode a nitrate-reductase, the number of MAGs that can reduce nitrate increases to 25/55.

Of the MAGs with capacity for nitrate reduction, there is only one MAG that encodes the complete set of genes for respiration of all reduced nitrogen oxides (the proteobacterial MAG PRO3, which carries, in addition to narGHI, also nirS, norBC, and nosZ of the clade I type). Two of the narGH encoding MAGs (ACT2 at 0.23% and CFX15 at 0.34% relative metagenome abundance) have no additional genes related to NOx respiration.

The two most abundant heterotrophic MAG, CFX2 (at 8.75% relative abundance) and CLB1 (at 5.20% relative abundance), have, in addition, to a narGH operon, solely a nirK and a nosZ (class II) gene, respectively, clearly indicating incomplete denitrification pathways.

While 23 MAGs encoded genes for respiratory nitrate reduction, 27 MAGs encoded genes for respiratory nitrite reduction (8 nirS, 19 nirK). Only 7 of these MAGs carried both nir and nar genes. 8 MAGs encoded genes for nitric oxide reduction (3 with norZ, 5 via norBC). Most of these also carried nir genes (7/8), but only half (4/8) carried both nir and nar genes. Only one MAG carried a class I nosZ gene (PRO 3). As stated above, PRO3 was the only MAG with all genes for a complete denitrification pathway. On the other hand, many MAGs encoded a class II nosZ gene (20/55), revealing a very high genomic potential for high-affinity N2O reduction. None of these MAGs carried genes for NO reduction (i.e. the norBC or norZ genes), while 7 of these MAGs carried genes for nitrite reduction (nirS or nirK, CFX6, PRO1, BCT2, PRO5, CFX9, BCT5, BCT6) or for nitrate reduction (narGHI, IGN1, IGN2, PRO5, BCG1, CFX8, BCT11, BCD2), respectively.

While MAGs with the capability of NO3 to NO2 reduction and N2O to N2 were abundant, MAGs with incomplete denitrification pathways prevailed (Figs. 2 and 3). MAGs that were characterized as NO2 to N2O reducers consisted of AOB, consistent with process observations [22]. In addition, a large number of MAGs with potential to exclusively reduce NO2 to NO or with potential for NO2 to NO, and N2O to N2 reduction were recovered. Several MAGs with potential to consume NO were retrieved, but especially the AnAOB MAGs AMX2 and AMX3 stand out as they did not encode the expected nitrite reductase (nirK). The abundance of MAGs with genetic potential for NO production (and not consumption) suggest that NO is exchanged in the community; AMX2 and AMX3 are the obvious NO consumers, suggesting growth of AnAOB on NO (not NO2) as electron acceptor as recently documented [23]. While there are MAGs with the exclusive ability to reduce N2O, most of them also encode genes (nar or nrf) allowing for NO3 to NO2 or NO2 to NH4+ reduction (Additional file 3: Table S2).

Fig. 3
figure 3

Overview of the genomic potential for production and consumption of NOx intermediates as distributed across the MAGs

Auxotrophy across MAGs

The heterotrophs in anammox or nitritation-anammox systems can either provide growth factors to [14] or depend on growth factors from [15] the autotrophic community members. Hence, we examined the genomic potential for synthesis of AA and B-vitamins across the MAGs. Of all the recovered MAGs, only one (PRO4) was fully prototrophic for its AA; all other MAGs were at least auxotrophic in one and up to all (PAT1) amino acids (Fig. 4). The dominant autotrophs (AOB) were nearly prototrophic, all three missing the cysteine, plus the alanine (AOB2) and leucine (AOB1) biosynthesis pathway, respectively. Similarly, among the AnAOB, AMX1 and AMX2 only missed the methionine biosynthetic pathway, while AMX3, in addition, lacked complete histidine and proline biosynthetic pathways. Similarly, none of the MAGs was totally prototrophic for its B-vitamin synthesis, with some MAGs (PAT1, PAT3, PRO10, PLA1, VER1, VER2, VER3 GMM2) completely devoid of this genetic potential (Fig. 4). The potential for cobalamin biosynthesis (Vit B12) was only retrieved in the AnAOB MAGs AMX1 and AMX2. Correlations between the degree of prototrophy and MAG completeness or MAG abundance were low: both nearly complete MAGs (> 95%) and very rare MAGs (< 5%) ranged in AA prototrophy from 25 to 100%; yet the most abundant MAG (AMX2) was the most prototrophic (Additional file 1: Fig. S6).

Fig. 4
figure 4

Presence/absence of biosynthetic pathways for amino acid and B-vitamins across MAGs

Discussion

Performance and NOB suppression

PNA has been successfully attained using various reactor configurations/operations and resulting biomass morphologies [5]. In general, attached growth (biofilms, aggregates) seems necessary to obtain sufficient AnAOB in the system, yet the same systems are challenged by the retention of unwanted NOB. NOB control is easier to obtain in suspended growth (or hybrid suspended/attached growth) systems [11]. Here we document that granular systems—obtained by sequential feeding and subjected to intermittent aeration—permit both retention of AOB and AnAOB (regulated by diffusional transport of O2, NH4+ and NO2[24])) but also control NOB density. While causes for repression of NOB by aeration switching have been proposed [12, 25,26,27], direct proof has remained elusive and this operational control towards microbiome engineering remains heuristic.

Presence and diversity of AOB and AnAOB

In the resulting PNA community, operated on a synthetic feed with NH4+ as the sole energy source, AOB and AnAOB amounted to 13% (stdev: 5%) and 22% (stdev: 6%) of the community (as fractions of the metagenome). This is similar to the fractions observed in other PNA communities by Speth et al. (AOB max 4% AnAOB max 20%) and Wang et al. (AOB ca. 25% AnAOB ca. 40%) [28]. While Speth et al. [29] and Wang et al. [28] identified only one MAG as AOB and AnAOB, we identified 3 MAGs each that could be classified as AOB and AnAOB. This may, in part, be due to the higher fraction of the MG that could be assigned to the different MAGs (79% here vs only 59% in [29]); even though the distribution across MAGs was more equitable in our study compared to Wang et al. [28]. The dominant AnAOB and AOB in the current system were closely related to Candidatus Brocadia sp., and Nitrosomonas europaea and N. eutropha as typically found in these highly loaded synthetic PNA or (for AnAOB) anammox communities [14, 19, 28, 29].

Absence of NOB

Although 16S rRNA gene targeted qPCR ([16, 30]) and 16S rRNA gene amplicon targeted community analysis (Additional file 1: Fig. S1) indicated a small Nitrospira presence (< 0.5%), no MAGs encoding autotrophic nitrite oxidation were recovered, and the presence in the whole metagenome was also minimal (ca. 4 RPM mapped to canonical NOB nxr). Speth et al. [29], on the other hand, detected a Nitrospira MAG (at 1.6 to 2.8%); yet also Wang et al. did not detect NOB MAGs in their PNA community MG [28]. As both our study and [28] applied sequential (instead of continuous low-rate) aeration to support the PNA community, this may be an effective strategy for NOB counter selection. We also note that, in fact, Nitrospira detection based on 16S rRNA gene or on nxrA, as done here, is not necessarily indicative of the presence of strict NOB since comammox Nitrospira have also been recovered from PNA systems [31]. Therefore the low fraction of Nitrospira we detect might be an overestimate of NOB abundance.

The types of HB and the role of auxotrophy vs. prototrophy of HB

Heterotrophic bacteria were abundant in this study (57% of MG vs autotrophs 43%); consistent with other studies [14, 15, 29]. They were distributed across a diverse set of phylotypes but with notable abundance in Chloroflexi (MAG CFX 1-15 at 23%), Ignavibacteriales/Chlorobi (MAG CLB 1-4 at 5%), Armatimonadates (MAG ARM1 at 4%), Bacteroidetes/Flavobacteria (MAG BCD 1,2 at 4%), and a few Proteobacteria (MAG PRO 2 and 6 at 3 and 2%). These taxa are all typically found in PNA and/or anammox communities. Heterotrophs are assumed to be supported by soluble microbial products actively or passively released by autotrophic PNA members [32]. In addition, others have suggested heterotrophs as essential in providing growth factors to autotrophs [14, 33], a claim not consistent with our findings. Auxotrophy for AA and vitamin biosynthesis were present in both autotrophs as heterotrophs; the most auxotrophic MAGs were heterotrophs, and AnAOB MAGs were the only MAGs encoding potential for cobalamin biosynthesis (Fig. 4). Clearly mutual dependencies beyond exchange of N species drive the composition of the PNA microbiome [15, 34].

Denitrification pathways

The current metagenome analysis indicates that heterotrophic MAGs have varying abilities for NOx respiration, with only one MAG encoding a complete denitrification pathway. Similar observations were made before: rare MAGs encoding full denitrification, but wide potential (and expression) for NO3 to NO2 respiration across multiple MAGs [15, 19, 29]. These observations have supported the notion that heterotrophs in PNA systems support a nitrite loop [35]. Our analysis results suggest, in addition, an abundance of MAGs with NO as the predicted end product of NOx respiration. In combination with the fact that the two AnAOB MAGs AMX2 and AMX3 (Additional file 3: Table S2) lack NIR encoding genes, this raises the possibility for NO cycling between autotrophic and heterotrophic MAGs. While the ability to support anaerobic ammonium oxidation supported by NO reduction (instead of NO2) has been shown in pure culture [23], direct proof in a PNA microbiome awaits confirmation.

In conclusion, our metagenomics analysis indicates that intermittent aeration is a highly-efficient control strategy to suppress NOB presence in a PNA process. The resulting microbiome presents mutual dependencies between the AOB and AnAOB autotrophs and heterotrophs, and a N cycle network that involves NO exchange.

Materials and methods

Sample collection and extraction of DNA

Mixed biomass was collected as 2 mL grab samples at the end of each set frequency period (monthly) and at the end of a react phase. The samples were centrifuged at 10,000xg, supernatant removed and stored as pellets at − 20 °C until DNA extraction. DNA was extracted using the MP FastDNA Spin Kit (MP Biomedicals LLC, Solon, USA) following manufacturer instructions. DNA concentration and quality were measured by NanoDrop (NanoDrop Technologies, Wilmington, USA).

Library preparation, sequencing and de novo assembly

DNA-shearing and library preparation were conducted based on the NEXTflex Rapid DNA-Seq Kit, V13.08 (Bioo Scientific, Austin, TX, USA) as described [36]. Sequencing was performed as a 100-bp pair-end run on HiSeq 2000 (Illumina Int., San Diego, CA, USA) at BGI (Copenhagen, Denmark). FastQC [37] was used for quality control, while Trimmomatic v0.22 [38] was run to remove adapters and trim the reads (threshold quality = 30; minimum length = 45). Assembly of high-quality reads from each sample into contigs was performed using IDBA-UD [39] with default parameters.

Metagenomic assembled genomes recovery and annotation

Genome binning was conducted based on pentanucleotide signatures using VizBin [40] and differential coverage using mmgenome [41]. Resulting metagenomic assembled genomes (MAGs) were manually evaluated through contig depth and GC content. Completeness and potential contamination of each MAG was evaluated using CheckM [42]. When the same draft MAG was obtained from several samples, comparison between them was implemented, retaining the one best assembled, most complete, and with the lowest contamination. MAGs with at least 70% of completeness, or an inferior number without contamination were further analyzed. The relative abundance of the MAGs was calculated with CoverM v0.6.1 (https://github.com/wwood/CoverM).

Taxonomic and functional annotation

MAGs were classified using the classify workflow of the GTDB-Tk v.0.1.3 tool [50]. Predicted coding sequences retrieved using Prodigal 2.50 [43] were annotated using USEARCH [44] -ublast against the manually created databases of reference proteins encoded by genes of interest (best hit with E < 1e−5, Bitscore > 60 and sequence similarity > 30%). Furthermore, to confirm protein functional assignment, Kyoto encyclopedia of genes and genomes (KEGG) annotations of the predicted proteins in each CG were obtained using the WebMGA server [45]. Presence of the complete operon was evaluated in MG-RAST [46]. The presence of carbon fixation pathways in each MAG was evaluated using METABOLIC [47]. In addition, GhostKOALA was used to assign KEGG orthology (KO) to each predicted ORFs of retrieved MAGs [48]. KEGG mapper was then used to process KO annotation results and reconstruct metabolic pathways of the retrieved MAGs. The presence/absence of an amino acid or vitamin biosynthesis pathway was estimated based on the encoded genes for the steps within a given pathway. If a reaction step could be catalyzed by more than one enzyme, the presence of a gene encoding one of the enzymes was regarded as the presence of this reaction step. The presence of a metabolic pathway in a given MAG was estimated as follows: If a pathway includes less than two steps, all steps are required to be encoded. If a pathway included more than two steps, only one missing step was allowed.

To detect canonical nxr in the whole metagenome, VSEARCH (–usearch_global, –id 0.85, –query_cov 0.9) [49] was used to map filtered metagenomic reads against a custom database containing the nxrA gene of previously described canonical NOB (Candidatus Nitrospira defluvii, Nitrospira lenta, Nitrobacter hamburgensis, Nitrobacter winogradskyi, Nitrolancea hollandica, Candidatus Nitrotoga sp. KNB and Candidatus Nitronauta litoralis). Same analysis was conducted to detect nxr from AnAOB (using the nxrA from Candidatus Brocadia pituitae and Candidatus Kuenenia stuttgartiensis).

Phylogenetic analysis

Phylogenetic analyses of the recovered MAGs were conducted with the GTDB-Tk v.0.1.3 tool [50] using the de novo workflow with a set of 120 single copy marker proteins and the genome taxonomy database (GTDB) [51]. Predicted NxrA/NarG and HaoA/HzoA amino-acid sequences were independently aligned with reference sequences using MUSCLE [52]. These alignments were used to construct maximum likelihood trees using RAxML v. 8.2.11 (the number of bootstraps was determined using the autoMRE option) [53]. For NxrA/NarG, the tree was built using the PROTGAMMAILGF model of sequence evolution, while for HaoA/HzoA the tree was constructed using the PROTGAMMAIWAG model. In both cases the best model was determined using ProtTest v. 3.4.2 [54]. All trees were visualized using the online web tool from the Interactive Tree of Life (iTol) [55].

Data availability

16S rRNA gene sequences, shotgun metagenomic sequences and metagenome-assembled genomes retrieved from the bioreactor have been deposited at NCBI under the project PRJNA791618.

References

  1. Widder S, Allen RJ, Pfeiffer T, Curtis TP, Wiuf C, Sloan WT, et al. Challenges in microbial ecology: building predictive understanding of community function and dynamics. ISME J. 2016;10:2557–68.

    Article  PubMed  PubMed Central  Google Scholar 

  2. Lawson CE, Harcombe WR, Hatzenpichler R, Lindemann SR, Löffler FE, O’Malley MA, et al. Common principles and best practices for engineering microbiomes. Nat Rev Microbiol. 2019;17:725–41.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Terada A, Lackner S, Tsuneda S, Smets BF. Redox-stratification controlled biofilm (ReSCoBi) for completely autotrophic nitrogen removal: The effect of co- versus counter-diffusion on reactor performance. Biotechnol Bioeng. 2007;97:40–51.

    Article  PubMed  CAS  Google Scholar 

  4. Vlaeminck E, Terada A, Smets BF, De Clippeleir H, Schaubroeck T, Bolca S, et al. Aggregate size and architecture determine biomass activity for one-stage partial nitritation and anammox. Appl Environ Microbiol. 2010;76:900–9.

    Article  PubMed  CAS  Google Scholar 

  5. Vlaeminck SE, De Clippeleir H, Verstraete W. Microbial resource management of one-stage partial nitritation/anammox. Microb Biotechnol. 2012;5:433–48.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  6. Lackner S, Gilbert EM, Vlaeminck SE, Joss A, Horn H, van Loosdrecht MCM. Full-scale partial nitritation/anammox experiences–an application survey. Water Res. 2014;55:292–303.

    Article  PubMed  CAS  Google Scholar 

  7. Joss A, Derlon N, Cyprien C, Burger S, Szivak I, Traber J, et al. Combined nitritation–anammox: advances in understanding process stability. Environ Sci Technol. 2011;45:9735–42.

    Article  PubMed  CAS  Google Scholar 

  8. Weissbrodt DG, Wells GF, Laureni M, Agrawal shelesh, Goel R, Russo G, et al. Systems microbiology and engineering of aerobic-anaerobic ammonium oxidation. ChemRxiv. Cambridge: Cambridge Open Engage; 2020.

  9. Pérez J, Lotti T, Kleerebezem R, Picioreanu C, van Loosdrecht MCM. Outcompeting nitrite-oxidizing bacteria in single-stage nitrogen removal in sewage treatment plants: a model-based study. Water Res. 2014;66:208–18.

    Article  PubMed  CAS  Google Scholar 

  10. Vadivelu VM, Keller J, Yuan Z. Effect of free ammonia on the respiration and growth processes of an enriched Nitrobacter culture. Water Res. 2007;41:826–34.

    Article  PubMed  CAS  Google Scholar 

  11. Agrawal S, Seuntjens D, De CP, Lackner S, Vlaeminck SE. Success of mainstream partial nitritation/anammox demands integration of engineering, microbiome and modeling insights. Curr Opin Biotechnol. 2018;50:214–21.

    Article  PubMed  CAS  Google Scholar 

  12. Gilbert EM, Agrawal S, Brunner F, Schwartz T, Horn H, Lackner S. Response of different nitrospira species to anoxic periods depends on operational DO. Environ Sci Technol. 2014;48:2934–41.

    Article  PubMed  CAS  Google Scholar 

  13. Ma Y, Peng Y, Wang S, Yuan Z, Wang X. Achieving nitrogen removal via nitrite in a pilot-scale continuous pre-denitrification plant. Water Res. 2009;43:563–72.

    Article  PubMed  CAS  Google Scholar 

  14. Zhao Y, Liu SS, Jiang B, Feng Y, Zhu T, Tao H, et al. Genome-centered metagenomics analysis reveals the symbiotic organisms possessing ability to cross-feed with anammox bacteria in anammox consortia. Environ Sci Technol. 2018;52:11285–96.

    Article  PubMed  CAS  Google Scholar 

  15. Wang YY, Niu Q, Zhang X, Liu L, Wang YY, Chen Y, et al. Exploring the effects of operational mode and microbial interactions on bacterial community assembly in a one-stage partial-nitritation anammox reactor using integrated multi-omics. Microbiome. 2019;7:1–15.

    Article  CAS  Google Scholar 

  16. Mutlu AG. Management of microbial community composition, architecture and performance in autotrophic nitrogen removing bioreactors through aeration regimes Denmark. PhD Thesis, Department of Environmental Engineering, Technical University of Denmark. 2015.

  17. Mutlu AG, Vangsgaard AK, Sin G, Smets BF. An operational protocol for facilitating start-up of single-stage autotrophic nitrogen removing reactors based on process stoichiometry. Water Sci Technol. 2012. https://doi.org/10.2166/wst.2013.157.

    Article  Google Scholar 

  18. Lotti T, Kleerebezem R, Lubello C, van Loosdrecht MCM. Physiological and kinetic characterization of a suspended cell anammox culture. Water Res. 2014;60:1–14.

    Article  PubMed  CAS  Google Scholar 

  19. Lawson CE, Wu S, Bhattacharjee AS, Hamilton JJ, McMahon KD, Goel R, et al. Metabolic network analysis reveals microbial community interactions in anammox granules. Nat Commun. 2017;8:15416.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Sun Y, Guan Y, Zeng D, He K, Wu G. Metagenomics-based interpretation of AHLs-mediated quorum sensing in Anammox biofilm reactors for low-strength wastewater treatment. Chem Eng J. 2018;344:42–52.

    Article  CAS  Google Scholar 

  21. Okubo T, Toyoda A, Fukuhara K, Uchiyama I, Harigaya Y, Kuroiwa M, et al. The physiological potential of anammox bacteria as revealed by their core genome structure. DNA Res. 2021;28:1–12.

    Article  CAS  Google Scholar 

  22. Domingo-Félez C, Mutlu AG, Jensen MM, Smets BF. Aeration strategies to mitigate nitrous oxide emissions from single-stage nitritation/anammox reactors. Environ Sci Technol. 2014;48:8679–87.

    Article  PubMed  CAS  Google Scholar 

  23. Hu Z, Wessels HJCTCT, van Alen T, Jetten MSMM, Kartal B. Nitric oxide-dependent anaerobic ammonium oxidation. Nat Commun. 2019;10:1244.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Vlaeminck SE, Terada A, Smets BF, De Clippeleir H, Schaubroeck T, Bolea S, et al. Aggregate size and architecture determine microbial activity balance for one-stage partial nitritation and anammox. Appl Environ Microbiol. 2010;76:900–9.

    Article  PubMed  CAS  Google Scholar 

  25. Ma Y, Felez CD, Plósz BG, Smets BF. Suppression of nitrite-oxidizing bacteria in intermittently aerated biofilms: a model-based explanation. Poster session presented at MEWE and biofilms IWA specialist conference, Copenhagen, Denmark. 2016.

  26. Ma B, Bao P, Wei Y, Zhu G, Yuan Z, Peng Y. Suppressing nitrite-oxidizing bacteria growth to achieve nitrogen removal from domestic wastewater via anammox using intermittent aeration with low dissolved oxygen. Sci Rep. 2015;5:13048.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Ma Y, Piscedda A, Veras ADLC, Domingo-Félez C, Smets BF. Intermittent aeration to regulate microbial activities in membrane-aerated biofilm reactors: Energy-efficient nitrogen removal and low nitrous oxide emission. Chem Eng J. 2021;433:133630.

    Article  CAS  Google Scholar 

  28. Wang Y, Niu Q, Zhang X, Liu L, Wang Y, Chen Y, et al. Exploring the effects of operational mode and microbial interactions on bacterial community assembly in a one-stage partial-nitritation anammox reactor using integrated multi-omics. Microbiome. 2019;7:122.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  29. Speth DR, In’t Zandt MH, Guerrero-Cruz S, Dutilh BE, Jetten MSM. Genome-based microbial ecology of anammox granules in a full-scale wastewater treatment system. Nat Commun. 2016;7:11172.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Graham DW, Knapp CW, Van Vleck ES, Bloor K, Lane TB, Graham CE. Experimental demonstration of chaotic instability in biological nitrification. ISME J. 2007;1:385–93.

    Article  PubMed  CAS  Google Scholar 

  31. van Kessel MAHJ, Speth DR, Albertsen M, Nielsen PH, den Camp HJMO, Kartal B, et al. Complete nitrification by a single microorganism. Nature. 2015;528:555–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Ni BJ, Ruscalleda M, Smets BF. Evaluation on the microbial interactions of anaerobic ammonium oxidizers and heterotrophs in Anammox biofilm. Water Res. 2012;46:4645–52.

    Article  PubMed  CAS  Google Scholar 

  33. Ji XM, Zheng C, Wang YL, Jin RC. Decoding the interspecies interaction in anammox process with inorganic feeding through metagenomic and metatranscriptomic analysis. J Clean Prod. 2021;288:125691.

    Article  CAS  Google Scholar 

  34. Morris JJ, Lenski RE, Zinser ER. The black queen hypothesis: evolution of dependencies through adaptative gene loss. MBio. 2012;3:1–7.

    Article  Google Scholar 

  35. Lawson CE, Wu S, Bhattacharjee AS, Hamilton JJ, McMahon KD, Goel R, et al. Metabolic network analysis reveals microbial community interactions in anammox granules. Nat Commun. 2017. https://doi.org/10.1038/ncomms15416.

    Article  PubMed  PubMed Central  Google Scholar 

  36. Palomo A, Fowler SJ, Gülay A, Rasmussen S, Sicheritz-Ponten T, Smets BF. Metagenomic analysis of rapid gravity sand filter microbial communities suggests novel physiology of Nitrospira spp. ISME J. 2016;10:2569–81.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Andrews S. FastQC a quality control tool for high throughput sequence data. http://www.bioinformatics.babraham.ac.uk/projects/fastqc/. 2010.

  38. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30:2114–20.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Peng Y, Leung HCM, Yiu SM, Chin FYL. IDBA-UD: a de novo assembler for single-cell and metagenomic sequencing data with highly uneven depth. Bioinformatics. 2012;28:1420–8.

    Article  PubMed  CAS  Google Scholar 

  40. Laczny CC, Sternal T, Plugaru V, Gawron P, Atashpendar A, Margossian H, et al. VizBin - an application for reference-independent visualization and human-augmented binning of metagenomic data. Microbiome. 2015;3:1.

    Article  PubMed  PubMed Central  Google Scholar 

  41. 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;31:533–8.

    Article  PubMed  CAS  Google Scholar 

  42. 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;25:1043–55.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  43. Hyatt D, Chen G-L, Locascio PF, Land ML, Larimer FW, Hauser LJ. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinform. 2010;11:119.

    Article  CAS  Google Scholar 

  44. Edgar RC. Supplementary material - search and clustering orders of magnitude faster than BLAST. Bioinformatics. 2010;26:2460–1.

    Article  PubMed  CAS  Google Scholar 

  45. Wu S, Zhu Z, Fu L, Niu B, Li W. WebMGA: a customizable web server for fast metagenomic sequence analysis. BMC Genomics. 2011;12:444.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Keegan KP, Glass EM, Meyer F. MG-RAST, a metagenomics service for analysis of microbial community structure and function. In: Methods in molecular biology, vol. 1399. New York, NY: Humana Press; 2016.

    Google Scholar 

  47. Zhou Z, Tran PQ, Breister AM, Liu Y, Kieft K, Cowley ES, et al. METABOLIC: high-throughput profiling of microbial genomes for functional traits, metabolism, biogeochemistry, and community-scale functional networks. Microbiome. 2020;10:33.

    Article  CAS  Google Scholar 

  48. Kanehisa M, Sato Y, Morishima K. BlastKOALA and GhostKOALA: KEGG tools for functional characterization of genome and metagenome sequences. J Mol Biol. 2016;428:726–31.

    Article  PubMed  CAS  Google Scholar 

  49. Rognes T, Flouri T, Nichols B, Quince C, Mahé F. VSEARCH: a versatile open source tool for metagenomics. PeerJ. 2016;4:e2584.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Chaumeil PA, Mussig AJ, Hugenholtz P, Parks DH. GTDB-Tk: a toolkit to classify genomes with the Genome Taxonomy Database. Bioinformatics. 2020;36:1925–7.

    CAS  Google Scholar 

  51. Parks DH, Chuvochina M, Waite DW, Rinke C, Skarshewski A, Chaumeil P-A, et al. A standardized bacterial taxonomy based on genome phylogeny substantially revises the tree of life. Nat Biotechnol. 2018;36:996–1004.

    Article  PubMed  CAS  Google Scholar 

  52. Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004;32:1792–7.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30:1312–3.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  54. Darriba D, Taboada GL, Doallo R, Posada D. ProtTest 3: fast selection of best-fit models of protein evolution. Bioinformatics. 2011;27:1164–5.

    Article  PubMed  CAS  Google Scholar 

  55. Letunic I, Bork P. Interactive tree of life (iTOL) v3: an online tool for the display and annotation of phylogenetic and other trees. Nucleic Acids Res. 2016;44:W242–5.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

None

Funding

Danish Agency for Science Technology and Innovation; Research Project LaGas (12-132633) and Research Centre EcoDesign (09-067230), EU-FP7 People: Marie-Curie Actions, REA 607492, MERMAID ITN, Villlum Fonden, Expa-N 13391.

Author information

Authors and Affiliations

Authors

Contributions

AP, BDAS, MTS, AD, performed bioinformatic analysis, created figures; CDF, AGM operated and analyzed the reactor performance; YW and TZ performed additional bioinformatic analysis, BFS and AP coordinated the overall data interpretation, BFS, supported by AP and AD wrote the manuscript, all authors agreed on the final content of the manuscript.

Corresponding author

Correspondence to Barth F. Smets.

Ethics declarations

Ethics approval and consent to participate

Not applicable

Consent for publication

All authors have consented to publication

Competing interests

Authors declare no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1:

 Supplemental Information, Supplementary Figures and Table S3.

Additional file 2: Table S1:

 Quality,  abundance and taxonomy of the recovered metagenome-assembled genomes.

Additional file 3: Table S2:

Overview of the genetic content of the recovered metagenome-assembled genomes.

Rights and permissions

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 licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence 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 licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Palomo, A., Azevedo, D., Touceda-Suárez, M. et al. Efficient management of the nitritation-anammox microbiome through intermittent aeration: absence of the NOB guild and expansion and diversity of the NOx reducing guild suggests a highly reticulated nitrogen cycle. Environmental Microbiome 17, 39 (2022). https://doi.org/10.1186/s40793-022-00432-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40793-022-00432-2

Keywords