Next Article in Journal
Functional Microbial Communities in Hybrid Linear Flow Channel Reactors for Desulfurization of Tannery Effluent
Next Article in Special Issue
Effects of Hops Treatment on Nitrogen Retention, Volatile Fatty Acid Accumulations, and Select Microbial Populations of Composting Poultry Litter Intended for Use as a Ruminant Feedstuff
Previous Article in Journal
Developing New Tools to Fight Human Pathogens: A Journey through the Advances in RNA Technologies
Previous Article in Special Issue
Land-Use Change and Management Intensification Is Associated with Shifts in Composition of Soil Microbial Communities and Their Functional Diversity in Coffee Agroecosystems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Consumption of N2O by Flavobacterium azooxidireducens sp. nov. Isolated from Decomposing Leaf Litter of Phragmites australis (Cav.)

1
Leibniz Centre for Agricultural Landscape Research (ZALF), Eberswalder Str. 84, D-15374 Müncheberg, Germany
2
Institute of Microbiology, Leibniz University Hannover, Herrenhäuser Str. 2, D-30419 Hannover, Germany
3
Institute for Ecoscience, Aarhus University, C.F. Møllersvej, Bygning 1331, 8000 Aarhus, Denmark
4
Leibniz-Institute of Freshwater Ecology and Inland Fisheries Berlin, Müggelseedamm 301, D-12587 Berlin, Germany
*
Authors to whom correspondence should be addressed.
Microorganisms 2022, 10(11), 2304; https://doi.org/10.3390/microorganisms10112304
Submission received: 11 October 2022 / Revised: 8 November 2022 / Accepted: 10 November 2022 / Published: 21 November 2022
(This article belongs to the Special Issue Bacterial Functions in Carbon, Nitrogen, and Sulfur Cycles)

Abstract

:
Microorganisms acting as sinks for the greenhouse gas nitrous oxide (N2O) are gaining increasing attention in the development of strategies to control N2O emissions. Non-denitrifying N2O reducers are of particular interest because they can provide a real sink without contributing to N2O release. The bacterial strain under investigation (IGB 4-14T), isolated in a mesocosm experiment to study the litter decomposition of Phragmites australis (Cav.), is such an organism. It carries only a nos gene cluster with the sec-dependent Clade II nosZ and is able to consume significant amounts of N2O under anoxic conditions. However, consumption activity is considerably affected by the O2 level. The reduction of N2O was not associated with cell growth, suggesting that no energy is conserved by anaerobic respiration. Therefore, the N2O consumption of strain IGB 4-14T rather serves as an electron sink for metabolism to sustain viability during transient anoxia and/or to detoxify high N2O concentrations. Phylogenetic analysis of 16S rRNA gene similarity revealed that the strain belongs to the genus Flavobacterium. It shares a high similarity in the nos gene cluster composition and the amino acid similarity of the nosZ gene with various type strains of the genus. However, phylogenomic analysis and comparison of overall genome relatedness indices clearly demonstrated a novel species status of strain IGB 4-14T, with Flavobacterium lacus being the most closely related species. Various phenotypic differences supported a demarcation from this species. Based on these results, we proposed a novel species Flavobacterium azooxidireducens sp. nov. (type strain IGB 4-14T = LMG 29709T = DSM 103580T).

1. Introduction

Mitigating global climate change requires a comprehensive understanding of the mechanisms that produce and consume greenhouse gases. Various biotic and abiotic processes are involved in the release of nitrous oxide (N2O), the potent greenhouse gas and the dominant source of stratospheric ozone depletion [1,2]. As part of efforts to develop strategies to control N2O emissions, increasing attention is being paid to the reduction of N2O by microorganisms. The most important biologically catalysed N2O reduction is conducted by the copper cluster-containing enzyme N2O reductase NosZ [3]. In the past, N2O consumption has been attributed primarily to canonical denitrifying bacteria. However, genomic-based studies identified “atypical” NosZ coding genes (Clade II nosZ) [4] that are often more abundant in many biomes than the well-studied “typical” nosZ genes (Clade I nosZ) [5], which denitrifying bacteria typical possess. Furthermore, diversity is generally higher for Clade II nosZ than for Clade I, which is reflected by a larger taxonomic range of Clade II nosZ among sequenced genomes [2].
Organisms containing Clade II nosZ genes possess divergent nos gene clusters with genes that are evolutionarily distinct from the typical nos genes of denitrifiers [4,6,7]. The nos gene clusters have a gene upstream nosZ, which encodes a transmembrane protein (nosB), and are often associated with c- and b-type cytochromes, as well as iron–sulfur protein encoding genes. In contrast, the nosR and nosX genes, reported as essential to maintain the activity of Clade I NosZ, are absent in Clade II nos gene clusters. Another difference lies in the signal peptide for NosZ export across the cytoplasmic membrane [3,6]. Nearly all Clade II nosZ possess Sec-dependent (secretory pathway) signal peptides to transport the enzyme in unfolded form, whereas the Clade I NosZ signal peptide is recognised by the twin-arginine translocation system (Tat), where it is transported in folded form. Overall, the question of how far these obvious genomic differences between organisms with Clade I and II nosZ are translated into physiological differences that may influence niche differentiation within and between the individual clades cannot yet be clearly answered and requires further studies [2]. There is an open debate about whether N2O-reducing bacteria of Clade II have an inherently higher N2O affinity than their Clade I counterparts [6,8,9,10,11,12]. However, the clade type does not fully explain the different N2O affinities, indicating that the physiological properties of N2O-reducing bacteria differ at the species and strain levels. However, slightly more than half of Clade II nosZ organisms are apparently non-denitrifying N2O reducers and therefore have the potential to be a sink without contributing to N2O release [2,13].
Members of the genus Flavobacterium are widespread in the environment and occur in very diverse habitats—from marine and fresh waters, sediments and soils, various plant compartments, food, and food processing plants to clinical environments [14,15,16]. Although the reduction of nitrate (NO3) to nitrite (NO2) was known to occur in many species of the genus, denitrification was first described for the species Flavobacterium denitrificans in 2005 [17,18]. However, microbial communities in N2O biofiltration systems included Flavobacterium spp. as one of the most abundant Clade II nosZ-carrying organisms [19,20]. Analysis of enrichment cultures of activated sludge samples in a continuous N2/N2O stream revealed also a dominance of Flavobacterium spp. among both nosZ gene and transcript pools [21]. These results demonstrate a high potential of flavobacteria to serve as an N2O sink in natural or man-made environments. Screening available genomes of type strains for genes encoding enzymes for the reduction of different nitrogen species [22], we found a large diversity of strains harboring such genes, suggesting that these traits are widespread in Flavobacterium species. In addition, a high degree of modularity was found in the gene inventory, which also indicates functional differences.
Accompanying a mesocosm experiment to study the litter decomposition of Phragmites australis (Cav.) under different redox conditions, the NO3/NO2 reducing bacterial community was investigated in relation to N2O and ammonia flux measurements. In this context, a group of isolates was obtained that were assigned to the genus Flavobacterium. This group was interesting both from a taxonomic point of view and with regard to functional traits related to N-transformation processes. A primarily analysis of 16S rRNA phylogeny revealed a novel species status and a screening for reduction of different nitrogen species demonstrated the ability to consume N2O. Therefore, in this study the exact taxonomic position of this group was investigated by a polyphasic approach. Furthermore, physiological characteristics and the gene inventory were studied to determine the potential to participate in N-transformation processes.

2. Materials and Methods

2.1. Isolation and Primariy Classification

The isolates studied originate from a mesocosm experiment designed to investigate the litter decomposition of Phragmites australis (Cav.) under different redox conditions. Leaves were collected from Phragmites growing in a rewetted fen (Germany, 53°52′33.4′′ N, 12°53′20.7′′ E) placed in stainless steel litterbags (1 mm mesh size) and processed as described by Reuter, et al. [23]. The litterbags were transferred into 5 L Duran wide neck bottles with four port caps and incubated in artificial fen water medium (4.5 L). Therefore, organic-free, deionized water was supplemented with NaHCO3 and fresh detritus mud from the plant-sampling site as an inoculum. Furthermore, 200 mgL−1 NO3 or SO42− was added as an oxidant in comparison to a control without external oxidants. The mesocosms were incubated under dark conditions in a custom-built flow-through steady-state system which is described in detail by Rillig, et al. [24]. A temperature of 18 °C was maintained in the incubation vessels with the aid of a climate chamber. The headspace of the mesocosms (0.5 L) was continuously flushed with nitrogen (N2, 6 Lh−1) using a nitrogen membrane-generator NGM-11-LC/MS (CMC Instruments GmbH, Eschborn, Germany). Gas flux measurements were performed continuously during the whole incubation period according to Rillig, et al. [24].
To investigate the microbial community, leaves of a litter bag (10 g) were immersed in 90 mL of 0.2 M Sörensen sodium phosphate buffer (pH 7) and treated for 2 min at high speed in the Stomacher 400 Circulator (Seward Ltd., Thetford, UK). Serially diluted samples were inoculated onto selective growth medium G3M12, which was optimized for the isolation of NO3 reducing bacteria [25]. The medium consisted of a standard mineral base according to Stanier, et al. [26] supplemented with Hutner’s vitamin-free mineral base [27] and a thiamine and complex vitamin solution [25,28]. The C/N ratio was adjusted to 2.5, with ethanol added as a carbon source and KNO3 as a nitrogen source. The pH of the medium was corrected to 7.5 after autoclaving and addition of sterile filtered supplements. The agar plates were incubated for 2 weeks at 21 °C in an anoxic chamber [gas composition: 10% carbon dioxide (CO2), 5% hydrogen (H2), and 85% N2]. Forty isolates were randomly picked and purified on ½ strength nutrient agar II containing 15 gL−1 agar (½NAII; composition according to Sifin, Berlin, Germany) to classify the isolates according their taxonomic affiliation and functional traits in the nitrogen cycle. Classification of isolates according to their taxonomic background was performed by matrix-assisted laser desorption/ionization time of flight mass spectrometry (MALDI-TOF MS) as described by Ulrich, et al. [29].

2.2. Phylogenetic Analysis

The taxonomic classification was confirmed by sequencing the almost complete 16S rRNA gene [30]. The 16S rRNA gene was amplified using primers 8f and 1525r [31] according a protocol described by [32] and sequenced with the internal primers 1492r [31] and 782r [33]. Sequence similarity comparisons were carried out using the EzBiocloud database [34]. For the phylogenetic analysis, an alignment (1488 nt) of the 16S rRNA genes from closely related species was generated using the ClustalW algorithm with MEGA X [35]. A phylogenetic tree was constructed using the maximum-likelihood method based on evolutionary distances of the General Time Reversible model (+G+I) and the neighbour-joining algorithm applying the Kimura 2-parameter model (+G). To investigate a possible clonal origin Ribotyping was performed with restriction endonucleases EcoRI, PvuII as described by Behrendt, et al. [36]. Furthermore, ribosomal intergenic spacer regions of isolates were amplified with the primers 1492f and 115r and subsequently sequenced as described by Tokajian, et al. [37].

2.3. Genome Sequencing and Bioinformatic Analyses

Strain IGB 4-14T was cultured in nutrient broth II (NBII; Sifin, Germany) for 2 days at 25 °C. The whole genomic DNA was extracted as described by Ulrich et al. [38]. DNA was sequenced using the Pacific Biosciences (PacBio) RS II sequencing platform at Eurofins Genomics (Konstanz, Germany). Sequence reads were de novo assembled using the PacBio hierarchical genome assembly process (HGAP4). The assembly resulted in one contig with an average genome coverage of 209x. The genome sequence was circularized with Circlator ver. 1.5.5 and deposited in the GenBank database under accession no. CP096205.
For the phylogenomic analysis, a tree based on core genome phylogeny [39] was constructed as described by Behrendt et al. [40]. The analysis of 120 bacterial core marker genes resulted in a concatenated amino acid sequence alignment, which was used to calculate a maximum-likelihood tree (LG substitution model with F+G+I) with MEGA X [35]. Digital DNA-DNA hybridization (dDDH) and genomic G+C content were determined on the Type Strain Genome Server (TYGS; [41]). Average nucleotide index (ANI) values were calculated by the OrthoANIu procedure [42]. To support the taxonomic classification, average amino acid identity (AAI) as a further genome relatedness index was calculated using the EzAAI tool [43].
The genome sequences of IGB 4-14T and related species were annotated with Rapid Annotation using Subsystem Technology (RAST) version 2.0 [44,45]. Additional functional and pathway analyses were performed using the BlastKOALA web tool of the KEGG database [46] and the NCBI Prokaryotic Genome Annotation Pipeline [22]. To analyse the nosZ gene phylogeny amino acid sequences of type strains representing the genus Flavobacterium and reference strains for Clade II nosZ genes according to Sanford, et al. [4] were selected. A maximum-likelihood tree (LG substitution model with G+I) was calculated with MEGA X [35].

2.4. Phenotypic Characterisation

Phenotypic analyses were performed for strain IGB 4-14T and the reference strain Flavobacterium lacus NBRC 109715T. Standard methods for morphological and physiological characterization were conducted as described by Behrendt, et al. [47]. Unless otherwise stated, strains were cultivated on ½NAII or in the respective broth (½NBII) at 22 °C. Presence of oxidase was tested on Cytochrome Oxidase Test Strips (Merck). Production of flexirubin-type pigments and Congo red absorption were assessed following the methods of Bernardet, et al. [48]. Oxidation of carbon compounds and resistance to inhibitory chemicals were determined using GEN III MicroPlates (Biolog) according to the manufacturer’s instructions. Results were scored visually after 24 and 48 h. Additional physiological and enzymatic characteristics were determined using the API 20E and API 20NE test strip (bioMérieux) after 48 h.
Analysis of cellular fatty acids was performed at DSMZ GmbH (Braunschweig, Germany) after cultivation on tryptic soy agar (TSA) for 2 days at 20 °C. Cellular fatty acids were converted into fatty acid methyl esters (FAMEs) using minor modifications of the method of Miller [49] and Kuykendall, et al. [50]. The FAMEs were separated by gas chromatography and detected by a flame ionisation detector using Sherlock Microbial Identification System (MIDI, Microbial ID, Newark, DE, USA). Peaks were automatically integrated, and fatty acid identifications and percentages were calculated by standard software (Microbial ID, Library TSBA40, 4.10).

2.5. N2O Consumption

Strain IGB 4-14T was cultured under oxic conditions in ½NBII for 48 h at 20 °C. 100 µL of this culture was inoculated into sterile 12 mL flat bottom vials (Labco, Lampeter, UK) containing 5 mL ½NBII broth. To create anoxic or micro-oxic conditions, the headspace was replaced three times with helium. For micro-oxic conditions, sterile air was added to achieve 4% (v/v headspace) oxygen (O2). Furthermore, N2O and nitric oxide (NO) were added into vials of all experiments (oxic, micro-oxic and anoxic) at a final concentration of 0.8% (v/v headspace) and 6 ppm (v/v headspace), respectively. NO was included as an inducer molecule for transcription of the nos genes [51]. To avoid diffusion with the ambient air, an overpressure (0.05 MPa) was created with helium in the anoxic and micro-oxic experiment and with sterile air in the oxic experiment. All tests were carried out in six replicates. For each experiment, a control was performed in triplicate without inoculation. The inoculated and control vials were incubated at 20 °C and sampled for gas analysis according to the time sequence shown in Figure 4. N2O, O2 and CO2 were measured via gas chromatography (7890B GC Agilent Technologies Inc., Santa Clara, CA, USA) equipped with an autosampler (Combi Pal-xt system, CTC analytics, Zwingen, Switzerland). A thermal conductivity detector (TCD, G3440B, Agilent Technologies Inc., Santa Clara, CA, USA) was used for measurement of O2. A pulsed discharge helium ionization detector (PDHID, V1D-3-I-HP220, Valco Instruments Company Inc. VICI AG International, Houston, TX, USA) detected CO2 and N2O. Details of the method are described in Zaman, et al. [52].
Increase in optical density (OD600) was used for estimating cell growth. It was measured at the start and the end of the experiment in a microplate with a Tecan plate reader (Infinite M Plex, Tecan Trading AG, Männedorf, Switzerland) and path length correction.
Loss of N2O, O2 and CO2 due to gas sampling was accounted for in the subsequent calculations. Equilibrated headspace concentrations of gases were used to allow comparative assessments of reduction or production of gas. Statistical significance was analyzed with a Kruskal-Wallis-Test and Dunn’s post hoc test of multiple comparisons using rank sums adjusted with the Bonferroni method.

3. Results and Discussion

Based on a similarity analysis of the spectral data obtained by MALDI TOF-MS, the isolated strains were classified into taxonomic units at the species level. A group of isolates that could not be assigned to any taxa using the MALDI biotyper reference database originated from the 200 mgL−1 NO3-supplemented mesocosm. The abundance of this group was 2 × 105 colony forming units per gram of fresh leaf litter. By 16S rRNA gene sequence comparison, the group could be assigned to the genus Flavobacterium of the phylum Bacteroidota, but a clear species affiliation was still not possible. Analysis of ribosomal intergenic spacer regions of five randomly selected isolates revealed identical sequences. Furthermore, the riboprint patterns showed no differences at the strain level. These results clearly indicated a clonal origin of the isolates. Therefore, only one isolate, strain IGB 4-14T, was selected for detailed taxonomic and physiological studies.

3.1. Phylogenetic Analysis

Comparative 16S rRNA gene analysis of strain IGB 4-14T revealed the highest sequence similarity of 97.3% to the species F. lacus (Table 1). Considering the species boundary of 98.2–99.0% for the 16S rRNA gene sequence similarity proposed by Meier-Kolthoff, et al. [53], a separate species position of the isolated strain is indicated. Furthermore, the high abundance of the isolates indicated an active growth under the tested redox conditions. The phylogenetic tree demonstrated that the isolated strain clearly clusters with F. lacus NP180T in both treeing methods applied, which was additionally supported by a high bootstrap value of 82% (Figure 1). Their position in relation to a cluster formed by further species of the genus Flavobacterium could also be found by both treeing methods. However, only the position related to the species Flavobacterium piscinae was confirmed by a supporting bootstrap value of 70%.
It is known that phylogenetic analyses based on the 16S rRNA gene do not always provide the desired taxonomic resolution depending on the genus studied [54]. Nevertheless, it gives a complete picture regarding the relationships within the genus because genomes are not available for all species. However, the study of evolutionary relationships using orthologous genes of the core genome can be used for a more solid and alternative way for phylogenetic analysis [39]. Therefore, the genome of the strain was sequenced. The assembled genome comprises a single circular contig with a total length of 3,823,204 bp exhibiting 100% completeness. No plasmid is present. The genome encoded 3263 protein-coding sequences and the DNA G+C content calculated from the genome sequence is 33.8%, which is consistent with the description of the genus Flavobacterium [18].
The phylogenomic analysis revealed similar results like the 16S rRNA gene analysis. Strain IGB 4-14T formed a branch with F. lacus within a joint cluster with F. piscinae, Flavobacterium orientale and Flavobacterium filum (Figure 2). Overall genome relatedness indices substantiated the novel species position of strain IGB 4-14T (Table 1). The dDDH of closely related species resulted in values far below the species boundary of 70% [53]. Accordingly, the ANI and AAI values were below the recommended species cut-off level of 95% [55,56].
In the summary of the results, it becomes clear that strain IGB 4-14T deserves its own species position from a phylogenetic point of view.

3.2. Phenotypic Analysis

Strain IGB-14T was comprehensively characterized by a morphological, physiological and chemotaxonomic analysis. The features characterizing the strain and thus the proposed novel species in detail are listed in the protologue (see Section 5). This section focuses on characteristics that allow the strain to be distinguished from the type strain F. lacus NBRC 109715T, which possess the highest phylogenetic relationship.
The fatty acid profile of strain IGB 4-14T is similar to the type strain of the related species F. lacus (Table 2). Both strains display fatty acids iso-C15:0 and iso-C15:1 G as the major components. Minor amounts were detected for iso-C15:0 3-OH, iso-C17:0 3-OH and iso-C17:1 w9c, with the proportions differing significantly. Qualitative differences were only found for fatty acids, which occurred in traces, so that a distinction between both strains was mainly based on quantitative differences. For type strains of species clustering in the phylogenomic analysis with the investigated strains, similar fatty acid compositions were described, especially with regard to the dominant acids [57,58]. The profiles match the general description of the genus Flavobacterium [18]. Furthermore, the strain IGB 4-14T can be clearly distinguished from the related species F. lacus due to several differentiating physiological characteristics (Table 3). These results provide phenotypic support for the separate species status already demonstrated in the phylogenetic analysis. Therefore, the assignment of the investigated isolate to a novel species, Flavobacterium azooxidireducens sp. nov., is proposed.

3.3. Genome Inventory with Regard to the N-Oxide Reduction

The genome of strain IGB 4-14T does not contain genes coding for NO3, NO2 and NO reductases. Only the gene of the sec-dependent Clade II NosZ (locus_tag M0M57_RS11105), followed by genes of various accessory proteins of the NosZ, could be detected. Thus, this bacterium possesses only genes for N2O reduction, and therefore has the potential to be an important sink for N2O without the capabilities to act as an N2O source.
The nosZ phylogeny followed the phylogenetic relationship (Figure 3). Strains of the genus Flavobacterium constitute a monophyletic branch separate from the reference strains for Clade II NosZ included in the analysis according to Sanford, et al. [4]. Type strains of the species F. orientale, F. filum and F. piscinae formed a cluster with the strain IGB 4-14T similar to the phylogenomic analysis (Figure 2). However, the type strain of F. lacus, the species with the closest relationship, does not possess any gene associated with N-transformation processes. These results are consistent with the general view that diversification of nosZ occurred predominantly through vertical inheritance, and on a finer scale the distribution of nosZ can be uneven, as nosZ may be present or absent in the genomes of closely related organisms [2].
The nosZ gene of strain IGB 4-14T is followed by genes encoding NosZ asseccory proteins annotated as nosL (M0M57_RS11110), nosD (M0M57_RS11115), nosF (M0M57_RS11120) and nosY (M0M57_RS11125). An identical nos gene cluster was found for the closely related species, F. denitrificans and F. crocinum (Figure 3). In contrast, the other flavobacteria involved in the study exhibit two genes at this position between nosZ and nosD. The first gene is annotated as a hypothetical protein, whereas the second gene is clearly annotated as nosL. However, a nosB gene typical for a Clade II nos gene cluster was not identified. The nosB gene encoding a yet uncharacterized polytopic membrane protein [59] is often surrounded by nosZ and nosD [4,7]. In Wollinella succinogenes, Hein, et al. [60] demonstrated that NosB is essential for N2O respiration by characterising a non-polar nosB deletion mutant. It is assumed that the protein is involved in electron transport to NosZ, but possibly also in copper management [59]. The latter is derived from the fact that in certain bacteria like Dyadobacter fermentans NosB is fused to the copper binding protein NosL [59,60].
It is likely that a fusion protein is also produced by strain IGB 4-14T and flavobacteria, which have an identical nos gene cluster (Figure 3). Sequence comparisons of the nosL gene of these bacteria with genes encoding the hypothetical protein and NosL of the other flavobacterial group revealed high similarity to the first gene segment and the following segment, respectively. A nosB amino-acid sequence comparison of Gemmatimonas aurantiaca T-27T with the flavobacterial sequences of the hypothetical protein or the corresponding stretch of the fusion protein revealed identity values between 40 and 46%. Considering conservative substitutions in addition to identical amino-acids, similarity values of around 63% were achieved. In addition, two highly conserved amino-acid stretches were found in the investigated flavobacteria, which were described on the basis of an alignment of NosB homologues by Spanning [59]. In strain IGB 4-14T, they were located at positon 76 (LNHYIGM) and 144 (YGHN) of the fusion protein (M0M57_RS11110). These results imply that the hypothetical protein found in some strains is a NosB and the multidomain fusion protein of the other flavobacteria including IGB 4-14T represents a NosB-L that combines the functions of NosB and NosL.

3.4. Consumption of N2O

Strain IGB 4-14T consumed significant amounts of N2O under anoxic conditions (Figure 4a). Following a high consumption rate within the first 14 days, N2O consumption plateaued out until almost all N2O has been consumed by the end of the experiment. The N2O consumption activities were significantly affected by O2 levels (Figure 4c). Under microoxic conditions, N2O was consumed within the first 6 days. At the same rate, the O2 concentration in the test tubes decreased and the CO2 production (Figure 4b) increased. With this, the maximum of N2O consumption was reached and in the following course of the experiment there were no further changes in the investigated parameters. The average N2O consumption rate up to day 6 was 0.090 ± 0.066 µmol d−1 and was marginal higher than under anoxic conditions with 0.062 ± 0.022 µmol d−1. In contrast, consumption of N2O under oxic conditions started around day 6, after most of the O2 amount had been consumed. Until then, only low consumption rates of 0.006 ± 0.016 µmol d−1 on average were achieved, which increased to 0.054 ± 0.019 µmol d−1 in the further course of the experiment. The increase in CO2 production correlated negatively with the O2 amount in the test tube. When CO2 reached its maximum at day 20, N2O consumption ceased.
N2O reduction by NosZ usually requires anoxic conditions, as NosZ copper centers are susceptible to O2 [6]. Nevertheless, N2O reduction under oxic and microoxic conditions has been described, for example, for strains of the species Pseudomonas stutzeri (Clade I nosZ) and Azospira sp. (Clade II nosZ) [12,61,62,63]. N2O reducing bacteria are often facultatively aerobic bacteria isolated from anoxic to microoxic habitats that possess a respiratory chain involved in micro-aerobic respiration. It is assumed that this property gives the cells a certain aerotolerance and protects the N2O reductase from inactivation by O2 [6]. This could be an explanation for the moderate consumption of N2O under microoxic conditions with decreasing O2 content in this experiment. Nevertheless, the question arises why N2O consumption does not increase after the O2 content has been completely reduced. The cessation of CO2 formation could indicate exhausted growth. On the other hand, Nos recovery after exposure to O2 might be restricted. The latter is considered an important parameter for the selection of strains for N2O mitigation in engineered systems [6,12].
Another important result is the lack of CO2 formation during anaerobic N2O consumption. This indicates that the cells do not grow, which is also reflected in the optical density. During the incubation period, no significant increase in OD was recorded (data not shown). This result implies that N2O consumption by strain IGB 4-14T is not coupled to energy conservation for growth, which is in contrast to the common assumption that the reduction of N2O is associated with energy conservation but in line with recent studies demonstrating that this is not the case for all microorganisms [10]. G. aurantiaca T-27T is one of these organisms that has been extensively studied in terms of physiological properties [64,65]. Although it has been characterized as an obligate aerobic bacterium [66], it is able to reduce N2O not only under oxic, but also under microoxic and anoxic conditions when O2 was regularly refed [64,65]. The reduction of N2O under anoxic conditions apparently acts as an electron sink for metabolism to sustain viability during transient anoxia and is not associated with growth. However, it is not known how widespread this NosZ function is, as it has not yet been tested on a broader diversity of isolates [10]. A further alternative NosZ function is the detoxification of high N2O concentrations [2,10,67]. Sullivan, et al. [68] demonstrated the cytotoxic effect of elevated N2O concentrations by inactivation of vitamin B12, thereby preventing enzyme reactions essential for microbial metabolisms. Thus, in Paracoccus denitrificans the B12-requiring methionine synthase MetH was inhibited, and the bacterium switched to the vitamin B12-independent methionine synthase MetE pathway, through the transcriptional modulation of genes controlled by vitamin B12 riboswitches. The two synthases occur together in a large number of bacteria, but in some species only one is present but not the other. In the Bacteroidota phylum in particular, MetH is the predominant single enzyme [69]. Correspondingly, the strain IGB 4-14T also possesses only the gene encoding the B12-depent synthase MetH (locus tag M0M57_09420), similar to G. aurantiaca T-27T, which belong to the same clade. Therefore, NosZ can play an important role in the detoxification of high N2O concentrations in these organisms. However, the almost complete consumption of N2O indicates that strain IGB 4-14T is capable of consuming N2O even in the range of natural concentrations in the environment, and along with the demonstrated absence of other denitrification associated enzymes (see above), could be a sink without contributing to N2O formation.
Figure 4. (a) Cumulative net N2O reduction; (b) cumulative net CO2 production; and (c) O2 depletion by strain IGB 4-14T under anoxic, microoxic (4% O2) and oxic (20% O2) conditions. The plotted values are averages of six repetitions, corrected by the average of three controls. Error bars represent the standard deviations. Different letters indicate significant differences (p < 0.05).
Figure 4. (a) Cumulative net N2O reduction; (b) cumulative net CO2 production; and (c) O2 depletion by strain IGB 4-14T under anoxic, microoxic (4% O2) and oxic (20% O2) conditions. The plotted values are averages of six repetitions, corrected by the average of three controls. Error bars represent the standard deviations. Different letters indicate significant differences (p < 0.05).
Microorganisms 10 02304 g004

4. Protologue—Description of Flavobacterium azooxidireducens sp. nov.

Flavobacterium azooxidireducens (a.zo.o.xi.di.re.du’cens. N.L. neut. n. azooxidum, dinitrogen oxide; L. pres. part. reducens, reducing from L. v. reduco, reduce, bring back to a condition; N.L. part. adj. azooxidireducens, reducing nitrous oxide).
Cells are Gram-stain-negative, nonmotile rods, which are approximately 0.3–0.6 µm in diameter and 1.0–1.9 µm long. A capsule is present. Colonies are deep yellow, convex, circular, 1–4 mm in diameter with entire margins. Growth occurs at 4 to 30 °C (optimum 20–25 °C). No growth is observed at 37 °C. It is able to grow on TSA, Reasoner’s 2A agar and nutrient agar. Flexirubin-type pigments are produced, and Congo red is absorbed by colonies. Catalase and oxidase activities are present. Hydrolyses gelatin but not tyrosine, casein, aesculin, starch and DNA. Nitrate is not reduced to nitrite. Arginine dihydrolase, lysine and ornithin decarboxylase, tryptophan deaminase and β-galactosidase is not present. Indole and acetoin production is negative.
In the Biolog GEN III Microplate assay following substrates are oxidized: l-arginine, l-aspartic acid, acetoacetic acid, acetic acid, citric acid, dextrin, d-fructose-6-PO4, α-d-glucose, d-galactose, 3-methyl glucose, l-glutamic acid, d-galacturonic acid, l-galactonic acid lactone, d-glucuronic acid, glucuronamide, l-histidine, d-lactic acid methyl ester, l-lactic acid, d-maltose, n-acetyl-β-d-mannosamine, glycyl-l-proline, p-hydroxy-phenylacetic acid, propionic acid, and l-serine. Weak reactions are observed for d-arabitol and d-mannose. No reactions are obtained for d-aspartic acid, l-alanine, γ-amino-butryric acid, α-hydroxy butyric acid, β-hydroxy-d,l butyric acid, α-keto-butyric acid, d-cellobiose, d-fructose, d- and l-fucose, formic acid, gentiobiose, β-methyl-d-glucoside, n-acetyl-d-glucosamine, n-acetyl-d-galactosamine, glycerol, d-glucose-6-PO4, d-gluconic acid, α-keto-glutaric acid, inosine, myo-inositol, α-d-lactose, d-melibiose, d-mannitol, mucic acid, d-malic acid, l-malic acid, n-acetyl neuraminic acid, l-pyroglutamic acid, pectin, methyl pyruvate, quinic acid, d-raffinose, l-rhamnose, sucrose, stachyose, d-salicin, d-sorbitol, d-serine, d-saccharic acid, bromo-succinic acid, d-trehalose, d-turanose, and tween 40. The strain is able to grow in the presence of 1% sodium lactate, d-serine, troleandomycin, rifamycin SV, minocycline, lincomycin, vancomycin, nalidixic acid, lithium chloride, potassium tellurite, aztreonam, sodium butyrate and tetrazolium violet. It did not tolerate fusidic acid, guanidine HCl, niaproof 4, tetrazolium blue, and sodium bromate. The strain tolerates pH 6 and a NaCl concentration of 1%, but not pH 5 and 4% NaCl. In the API 20NE test system, d-glucose, d-maltose, potassium gluconate are assimilated, but not l-arabinose, d-mannose, n-acetyl-glucosamine, potassium gluconate, capric acid, adipic acid, malic acid, trisodium citrate, and phenylacetic acid.
The predominant fatty acids are iso-C15:0 and iso-C15:1 G. Minor amounts were detected for iso-C15:0 3-OH, iso-C17:1 w9c, iso-C17:0 3-OH, and anteiso-C15:0.
The type strain IGB 4-14T (LMG 29709T = DSM 103580T) was isolated from a mesocosm experiment with decomposing leaf litter of Phragmites australis (Cav.) in artificial fen water enriched with fresh detritus originating from a rewetted fen (Germany, 53°52′33.4′′ N, 12°53′20.7′′ E). The DNA G+C content of the type strain is 33.8%. The genome sequence accession number is CP096205 and the 16S rRNA gene accession number is LT598610.

5. Conclusions

In this study, a bacterium associated with decomposing leaf litter of Phragmites australis (Cav.) was clearly differentiated from related Flavobacterium species in a polyphasic approach applying phylogenomics and phenotypic analysis. Based on these data, a novel species Flavobacterium azooxidireducens sp. nov. with the type strain IGB 4-14T was proposed. The type strain is a non-denitrifying bacterium that was able to consume significant amounts of N2O under anoxic conditions by a N2O reductase of the Clade II. Due to the absence of other enzymes associated with denitrification, it has the potential to be a sink without contributing to N2O formation. The N2O consumption served as an electron sink for metabolism and/or to detoxify high N2O concentrations. It shows a high similarity in nos gene cluster composition and amino acid similarity of the nosZ gene with different type strains of the genus, so it can be assumed that the N2O reductase in these strains might also underlie this function and is typical for the genus Flavobacterium. However, in contrast to the NosZ function for energy conservation through anaerobic respiration, the importance and distribution of these alternative NosZ functions in the environment is not known and needs further investigation.

Author Contributions

Conceptualization, U.B., S.K., J.A. and A.U.; methodology, U.B., T.S., J.A., D.H.Z., M.A.H. and A.U.; investigation, U.B., T.S., and A.U.; writing—original draft preparation, U.B.; writing—review and editing, A.U., T.S., J.A., D.H.Z., M.A.H., S.K.; All authors have read and agreed to the published version of the manuscript.

Funding

The research was funded by the DFG, grant no. ZA 742/2-1, and the DFG Priority Program 1374 “Infrastructure-Biodiversity-Exploratories”, project no. 324639010. The publication of this article was supported by funds from the Leibniz Association’s Open Access Fund.

Data Availability Statement

The GenBank/EMBL/DDBJ accession numbers for the 16S rRNA gene sequences and the IGS spacer region of type-strain Flavobacterium azooxidireducens IGB 4-14T are LT598610 and LT616948, respectively. The genome sequence data for the type-strain are available under the GenBank/EMBL/DDBJ accession number CP096205. It has been deposited in the German Collection of Microorganisms and Cell Cultures GmbH (=DSM 103580T) and in BCCM/LMG Bacteria Collection in Belgian (=LMG 29709T).

Acknowledgments

We greatly acknowledge Peter Schumann for analysis of whole cell fatty acids and riboprint patterns (DSMZ-Braunschweig). We would also like to thank Aharon Oren and Bernhard Schink for reviewing the Latin name of the novel species.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ravinshakara, A.; Daniel, J.; Portmann, R. Nitrous oxide (N2O): The dominant ozone-depleting substance emitted in the 21st century. Science 2009, 326, 123–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hallin, S.; Philippot, L.; Löffler, F.E.; Sanford, R.A.; Jones, C.M. Genomics and ecology of novel N2O-reducing microorganisms. Trends Microbiol. 2018, 26, 43–55. [Google Scholar] [CrossRef] [PubMed]
  3. Pauleta, S.R.; Carepo, M.S.P.; Moura, I. Source and reduction of nitrous oxide. Coord. Chem. Rev. 2019, 387, 436–449. [Google Scholar] [CrossRef]
  4. Sanford, R.A.; Wagner, D.D.; Wu, Q.; Chee-Sanford, J.C.; Thomas, S.H.; Cruz-Garcia, C.; Rodriguez, G.; Massol-Deya, A.; Krishnani, K.K.; Ritalahti, K.M.; et al. Unexpected nondenitrifier nitrous oxide reductase gene diversity and abundance in soils. Proc. Natl. Acad. Sci. USA 2012, 109, 19709–19714. [Google Scholar] [CrossRef] [Green Version]
  5. Orellana, L.H.; Rodriguez-R, L.M.; Higgins, S.; Chee-Sanford, J.C.; Sanford, R.A.; Ritalahti, K.M.; Löffler, F.E.; Konstantinidis, K.T. Detecting nitrous oxide reductase nosZ genes in soil metagenomes: Method development and implications for the nitrogen cycle. Mbio 2014, 5, e01193-14. [Google Scholar] [CrossRef] [Green Version]
  6. Simon, J. Mitigation of laughing gas emissions by nitrous oxide respiring microorganisms. In Enzymes for Solving Humankind’s Problems: Natural and Artificial Systems in Health, Agriculture, Environment and Energy; Moura, J.J.G., Moura, I., Maia, L.B., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 185–211. [Google Scholar] [CrossRef]
  7. Torres, M.J.; Simon, J.; Rowley, G.; Bedmar, E.J.; Richardson, D.J.; Gates, A.J.; Delgado, M.J. Nitrous oxide metabolism in nitrate-reducing bacteria: Physiology and regulatory mechanisms. In Advances in Bacterial Electron Transport Systems and Their Regulation; Poole, R.K., Ed.; Advances in Microbial Physiology; Academic Press Ltd-Elsevier Science Ltd.: London, UK, 2016; Volume 68, pp. 353–432. [Google Scholar] [CrossRef] [Green Version]
  8. Yoon, S.; Nissen, S.; Park, D.; Sanford, R.A.; Löffler, F.E. Nitrous oxide reduction kinetics distinguish bacteria harboring clade I NosZ from those harboring clade II NosZ. Appl. Environ. Microbiol. 2016, 82, 3793–3800. [Google Scholar] [CrossRef] [Green Version]
  9. Hein, S.; Simon, J. Chapter Four—Bacterial nitrous oxide respiration: Electron transport chains and copper transfer reactions. Adv. Microb. Physiol. 2019, 75, 137–175. [Google Scholar] [CrossRef]
  10. Shan, J.; Sanford, R.A.; Chee-Sanford, J.; Ooi, S.K.; Löffler, F.E.; Konstantinidis, K.T.; Yang, W.H. Beyond denitrification: The role of microbial diversity in controlling nitrous oxide reduction and soil nitrous oxide emissions. Glob. Change Biol. 2021, 27, 2669–2683. [Google Scholar] [CrossRef]
  11. Suenaga, T.; Hori, T.; Riya, S.; Hosomi, M.; Smets, B.F.; Terada, A. Enrichment, isolation, and characterization of high-affinity N2O-reducing bacteria in a gas-permeable membrane reactor. Environ. Sci. Technol. 2019, 53, 12101–12112. [Google Scholar] [CrossRef]
  12. Suenaga, T.; Riya, S.; Hosomi, M.; Terada, A. Biokinetic Characterization and activities of N2O-reducing bacteria in response to various oxygen levels. Front. Microbiol. 2018, 9, 697. [Google Scholar] [CrossRef]
  13. Graf, D.R.H.; Jones, C.M.; Hallin, S. Intergenomic comparisons highlight modularity of the denitrification pathway and underpin the importance of community structure for N2O emissions. PLoS ONE 2014, 9, e114118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Bernardet, J.-F. Flavobacteriaceae. In Bergey’s Manual of Systematics of Archaea and Bacteria; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2015. [Google Scholar] [CrossRef]
  15. Kolton, M.; Erlacher, A.; Berg, G.; Cytryn, E. The Flavobacterium genus in the plant holobiont: Ecological, physiological, and applicative insights. In Microbial Models: From Environmental to Industrial Sustainability; Castro-Sowinski, S., Ed.; Springer: Singapore, 2016; pp. 189–207. [Google Scholar] [CrossRef]
  16. Bernardet, J.-F.; Bowman, J.P. The genus Flavobacterium. In The Prokaryotes: Volume 7: Proteobacteria: Delta, Epsilon Subclass; Dworkin, M., Falkow, S., Rosenberg, E., Schleifer, K.-H., Stackebrandt, E., Eds.; Springer: New York, NY, USA, 2006; pp. 481–531. [Google Scholar] [CrossRef]
  17. Horn, M.A.; Ihssen, J.; Matthies, C.; Schramm, A.; Acker, G.; Drake, H.L. Dechloromonas denitrificans sp. nov., Flavobacterium denitrificans sp. nov., Paenibacillus anaericanus sp. nov. and Paenibacillus terrae strain MH72, N2O-producing bacteria isolated from the gut of the earthworm Aporrectodea caliginosa. Int. J. Syst. Evol. Microbiol. 2005, 55, 1255–1265. [Google Scholar] [CrossRef] [PubMed]
  18. Bernardet, J.-F.; Bowman, J.P. Flavobacterium. In Bergey’s Manual of Systematics of Archaea and Bacteria; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2015. [Google Scholar] [CrossRef]
  19. Yoon, H.; Song, M.J.; Kim, D.D.; Sabba, F.; Yoon, S. A serial biofiltration system for effective removal of low-concentration nitrous oxide in oxic gas streams: Mathematical modeling of reactor performance and experimental validation. Environ. Sci. Technol. 2019, 53, 2063–2074. [Google Scholar] [CrossRef] [PubMed]
  20. Yoon, H.; Song, M.J.; Yoon, S. Design and feasibility analysis of a self-sustaining biofiltration system for removal of low concentration N2O emitted from wastewater treatment plants. Environ. Sci. Technol. 2017, 51, 10736–10745. [Google Scholar] [CrossRef]
  21. Kim, D.D.; Park, D.; Yoon, H.; Yun, T.; Song, M.J.; Yoon, S. Quantification of nosZ genes and transcripts in activated sludge microbiomes with novel group-specific qPCR methods validated with metagenomic analyses. Water Res. 2020, 185, 116261. [Google Scholar] [CrossRef]
  22. Li, W.; O’Neill, K.R.; Haft, D.H.; DiCuccio, M.; Chetvernin, V.; Badretdin, A.; Coulouris, G.; Chitsaz, F.; Derbyshire, M.K.; Durkin, A.S.; et al. RefSeq: Expanding the Prokaryotic Genome Annotation Pipeline reach with protein family model curation. Nucleic Acids Res. 2021, 49, D1020–D1028. [Google Scholar] [CrossRef]
  23. Reuter, H.; Gensel, J.; Elvert, M.; Zak, D. Evidence for preferential protein depolymerization in wetland soils in response to external nitrogen availability provided by a novel FTIR routine. Biogeosciences 2020, 17, 499–514. [Google Scholar] [CrossRef] [Green Version]
  24. Rillig, M.C.; Hoffmann, M.; Lehmann, A.; Liang, Y.; Lück, M.; Augustin, J. Microplastic fibers affect dynamics and intensity of CO2 and N2O fluxes from soil differently. Microplastics Nanoplastics 2021, 1, 3. [Google Scholar] [CrossRef]
  25. Heylen, K.; Vanparys, B.; Wittebolle, L.; Verstraete, W.; Boon, N.; De Vos, P. Cultivation of denitrifying bacteria: Optimization of isolation conditions and diversity study. Appl. Environ. Microbiol. 2006, 72, 2637–2643. [Google Scholar] [CrossRef] [Green Version]
  26. Stanier, R.Y.; Palleroni, N.J.; Doudoroff, M. The aerobic pseudomonads: A taxonomic study. J. Gen. Microbiol. 1966, 43, 159–271. [Google Scholar] [CrossRef] [Green Version]
  27. Cohen-Bazire, G.; Sistrom, W.R.; Stanier, R.Y. Kinetic studies of pigment synthesis by non-sulphur purple bacteria. J. Cell. Comp. Physiol. 1957, 49, 25–68. [Google Scholar] [CrossRef] [PubMed]
  28. Kniemeyer, O.; Probian, C.; Rossello-Mora, R.; Harder, J. Anaerobic Mineralization of Quaternary Carbon Atoms: Isolation of Denitrifying Bacteria on Dimethylmalonate. Appl. Environ. Microbiol. 1999, 65, 3319–3324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Ulrich, K.; Becker, R.; Behrendt, U.; Kube, M.; Ulrich, A. A comparative analysis of ash leaf-colonizing bacterial communities identifies putative antagonists of Hymenoscyphus fraxineus. Front. Microbiol. 2020, 11, 966. [Google Scholar] [CrossRef] [PubMed]
  30. Behrendt, U.; Ulrich, A.; Schumann, P. Fluorescent pseudomonads associated with the phyllosphere of grasses; Pseudomonas trivialis sp. nov., Pseudomonas poae sp. nov. and Pseudomonas congelans sp. nov. Int. J. Syst. Evol. Microbiol. 2003, 53, 1461–1469. [Google Scholar] [CrossRef] [PubMed]
  31. Lane, D.J. 16S/23S rRNA sequencing. In Nucleic Acid Techniques in Bacterial Systematics; Stackebrandt, E., Goodfellow, M., Eds.; John Wiley and Sons: Chichester, UK, 1991; pp. 115–175. [Google Scholar]
  32. Ulrich, A.; Wirth, S. Phylogenetic diversity and population densities of culturable cellulolytic soil bacteria across an agricultural encatchment. Microb. Ecol. 1999, 37, 238–247. [Google Scholar] [CrossRef]
  33. Chun, J.; Goodfellow, M. A phylogenetic analysis of the genus Nocardia with 16S rRNA gene sequences. Int. J. Syst. Bacteriol. 1995, 45, 240–245. [Google Scholar] [CrossRef] [Green Version]
  34. Yoon, S.H.; Ha, S.M.; Kwon, S.; Lim, J.; Kim, Y.; Seo, H.; Chun, J. Introducing EzBioCloud: A taxonomically united database of 16S rRNA gene sequences and whole-genome assemblies. Int. J. Syst. Evol. Microbiol. 2017, 67, 1613–1617. [Google Scholar] [CrossRef]
  35. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  36. Behrendt, U.; Ulrich, A.; Schumann, P. Chryseobacterium gregarium sp. nov., isolated from decaying plant material. Int. J. Syst. Evol. Microbiol. 2008, 58, 1069–1074. [Google Scholar] [CrossRef] [Green Version]
  37. Tokajian, S.; Issa, N.; Salloum, T.; Ibrahim, J.; Farah, M. 16S-23S rRNA gene intergenic spacer region variability helps resolve closely related sphingomonads. Front. Microbiol. 2016, 7, 149. [Google Scholar] [CrossRef] [Green Version]
  38. Ulrich, K.; Kube, M.; Becker, R.; Schneck, V.; Ulrich, A. Genomic analysis of the endophytic Stenotrophomonas strain 169 reveals features related to plant-growth promotion and stress tolerance. Front. Microbiol. 2021, 12, 1542. [Google Scholar] [CrossRef] [PubMed]
  39. Parks, D.H.; Chuvochina, M.; Waite, D.W.; Rinke, C.; Skarshewski, A.; Chaumeil, P.-A.; Hugenholtz, P. A standardized bacterial taxonomy based on genome phylogeny substantially revises the tree of life. Nat. Biotechnol. 2018, 36, 996–1004. [Google Scholar] [CrossRef] [PubMed]
  40. Behrendt, U.; Wende, S.; Kolb, S.; Ulrich, A. Genome-based phylogeny of the genera Proteus and Cosenzaea and description of Proteus terrae subsp. terrae subsp. nov. and Proteus terrae subsp. cibarius subsp. nov. Int. J. Syst. Evol. Microbiol. 2019, 71, 004651. [Google Scholar] [CrossRef]
  41. Meier-Kolthoff, J.P.; Göker, M. TYGS is an automated high-throughput platform for state-of-the-art genome-based taxonomy. Nat. Commun. 2019, 10, 2182. [Google Scholar] [CrossRef] [PubMed]
  42. Yoon, S.H.; Ha, S.M.; Lim, J.; Kwon, S.; Chun, J. A large-scale evaluation of algorithms to calculate average nucleotide identity. Antonie Van Leeuwenhoek 2017, 110, 1281–1286. [Google Scholar] [CrossRef] [PubMed]
  43. Kim, D.; Park, S.; Chun, J. Introducing EzAAI: A pipeline for high throughput calculations of prokaryotic average amino acid identity. J. Microbiol. 2021, 59, 476–480. [Google Scholar] [CrossRef]
  44. Overbeek, R.; Olson, R.; Pusch, G.D.; Olsen, G.J.; Davis, J.J.; Disz, T.; Edwards, R.A.; Gerdes, S.; Parrello, B.; Shukla, M.; et al. The SEED and the rapid annotation of microbial genomes using subsystems technology (RAST). Nucleic Acids Res. 2013, 42, D206–D214. [Google Scholar] [CrossRef]
  45. Aziz, R.K.; Bartels, D.; Best, A.A.; DeJongh, M.; Disz, T.; Edwards, R.A.; Formsma, K.; Gerdes, S.; Glass, E.M.; Kubal, M.; et al. The RAST Server: Rapid annotations using subsystems technology. BMC Genom. 2008, 9, 75. [Google Scholar] [CrossRef] [Green Version]
  46. Kanehisa, M.; Sato, Y.; Kawashima, M.; Furumichi, M.; Tanabe, M. KEGG as a reference resource for gene and protein annotation. Nucleic Acids Res. 2015, 44, D457–D462. [Google Scholar] [CrossRef] [Green Version]
  47. Behrendt, U.; Ulrich, A.; Schumann, P.; Erler, W.; Burghardt, J.; Seyfarth, W. A taxonomic study of bacteria isolated from grasses: A proposed new species Pseudomonas graminis sp. nov. Int. J. Syst. Bacteriol. 1999, 49, 297–308. [Google Scholar] [CrossRef] [Green Version]
  48. Bernardet, J.F.; Nakagawa, Y.; Holmes, B. Proposed minimal standards for describing new taxa of the family Flavobacteriaceae and emended description of the family. Int. J. Syst. Evol. Microbiol. 2002, 52, 1049–1070. [Google Scholar] [CrossRef] [PubMed]
  49. Miller, L.T. Single derivatization method for routine analysis of bacterial whole-cell fatty acid methyl esters, including hydroxy acids. J. Clin. Microbiol. 1982, 16, 584–586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Kuykendall, L.D.; Roy, M.A.; O’Neill, J.J.; Devine, T.E. Fatty acids, antibiotic resistance, and deoxyribonucleic acid homology groups of Bradyrhizobium japonicum. Int. J. Syst. Evol. Microbiol. 1988, 38, 358–361. [Google Scholar] [CrossRef] [Green Version]
  51. Arai, H.; Mizutani, M.; Igarashi, Y. Transcriptional regulation of the nos genes for nitrous oxide reductase in Pseudomonas aeruginosa. Microbiology 2003, 149, 29–36. [Google Scholar] [CrossRef]
  52. Zaman, M.; Kleineidam, K.; Bakken, L.; Berendt, J.; Bracken, C.; Butterbach-Bahl, K.; Cai, Z.; Chang, S.X.; Clough, T.; Dawar, K.; et al. Automated Laboratory and Field Techniques to Determine Greenhouse Gas Emissions. In Measuring Emission of Agricultural Greenhouse Gases and Developing Mitigation Options Using Nuclear and Related Techniques: Applications of Nuclear Techniques for GHGs; Zaman, M., Heng, L., Müller, C., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 109–139. [Google Scholar] [CrossRef]
  53. Meier-Kolthoff, J.P.; Auch, A.F.; Klenk, H.-P.; Göker, M. Genome sequence-based species delimitation with confidence intervals and improved distance functions. BMC Bioinform. 2013, 14, 60. [Google Scholar] [CrossRef] [Green Version]
  54. Hördt, A.; López, M.G.; Meier-Kolthoff, J.P.; Schleuning, M.; Weinhold, L.-M.; Tindall, B.J.; Gronow, S.; Kyrpides, N.C.; Woyke, T.; Göker, M. Analysis of 1,000+ type-strain genomes substantially improves taxonomic classification of Alphaproteobacteria. Front. Microbiol. 2020, 11, 468. [Google Scholar] [CrossRef]
  55. Konstantinidis, K.T.; Tiedje, J.M. Towards a genome-based taxonomy for prokaryotes. J. Bacteriol. 2005, 187, 6258–6264. [Google Scholar] [CrossRef] [Green Version]
  56. Palmer, M.; Steenkamp, E.T.; Blom, J.; Hedlund, B.P.; Venter, S.N. All ANIs are not created equal: Implications for prokaryotic species boundaries and integration of ANIs into polyphasic taxonomy. Int. J. Syst. Evol. Microbiol. 2020, 70, 2937–2948. [Google Scholar] [CrossRef]
  57. Chen, W.-M.; Guo, Y.-P.; Kwon, S.-W.; Sheu, S.-Y. Flavobacterium piscinae sp. nov., isolated from a fish pond. Int. J. Syst. Evol. Microbiol. 2019, 69, 1775–1782. [Google Scholar] [CrossRef]
  58. Li, A.; Liu, H.; Sun, B.; Zhou, Y.; Xin, Y. Flavobacterium lacus sp. nov., isolated from a high-altitude lake, and emended description of Flavobacterium filum. Int. J. Syst. Evol. Microbiol. 2014, 64, 933–939. [Google Scholar] [CrossRef] [Green Version]
  59. van Spanning, R.J.M. Structure, function, regulation and evolution of the nitrite and nitrous oxide reductases: Denitrification enzymes with a Beta-propeller fold. In Nitrogen Cycling in Bacteria Molecular Analysis; Moir, J.W.B., Ed.; Caister Academic Press: Norfolk, UK, 2011; pp. 135–161. [Google Scholar]
  60. Hein, S.; Witt, S.; Simon, J. Clade II nitrous oxide respiration of Wolinella succinogenes depends on the NosG,-C1,-C2, -H electron transport module, NosB and a Rieske/cytochrome bc complex. Environ. Microbiol. 2017, 19, 4913–4925. [Google Scholar] [CrossRef] [PubMed]
  61. Desloover, J.; Roobroeck, D.; Heylen, K.; Puig, S.; Boeckx, P.; Verstraete, W.; Boon, N. Pathway of nitrous oxide consumption in isolated Pseudomonas stutzeri strains under anoxic and oxic conditions. Environ. Microbiol. 2014, 16, 3143–3152. [Google Scholar] [CrossRef] [PubMed]
  62. Zhao, B.; Cheng, D.Y.; Tan, P.; An, Q.; Guo, J.S. Characterization of an aerobic denitrifier Pseudomonas stutzeri strain XL-2 to achieve efficient nitrate removal. Bioresour. Technol. 2018, 250, 564–573. [Google Scholar] [CrossRef] [PubMed]
  63. Zhou, Y.; Suenaga, T.; Qi, C.; Riya, S.; Hosomi, M.; Terada, A. Temperature and oxygen level determine N2O respiration activities of heterotrophic N2O-reducing bacteria: Biokinetic study. Biotechnol. Bioeng. 2021, 118, 1330–1341. [Google Scholar] [CrossRef] [PubMed]
  64. Park, D.; Kim, H.; Yoon, S.; Kivisaar, M. Nitrous oxide reduction by an obligate aerobic bacterium, Gemmatimonas aurantiaca strain T-27. Appl. Environ. Microbiol. 2017, 83, e00502–e00517. [Google Scholar] [CrossRef] [Green Version]
  65. Chee-Sanford, J.; Tian, D.; Sanford, R. Consumption of N2O and other N-cycle intermediates by Gemmatimonas aurantiaca strain T-27. Microbiology 2019, 165, 1345–1354. [Google Scholar] [CrossRef]
  66. Zhang, H.; Sekiguchi, Y.; Hanada, S.; Hugenholtz, P.; Kim, H.; Kamagata, Y.; Nakamura, K. Gemmatimonas aurantiaca gen. nov., sp. nov., a Gram-negative, aerobic, polyphosphate-accumulating micro-organism, the first cultured representative of the new bacterial phylum Gemmatimonadetes phyl. nov. Int. J. Syst. Evol. Microbiol. 2003, 53, 1155–1163. [Google Scholar] [CrossRef]
  67. Conthe, M.; Wittorf, L.; Kuenen, J.G.; Kleerebezem, R.; van Loosdrecht, M.C.M.; Hallin, S. Life on N2O: Deciphering the ecophysiology of N2O respiring bacterial communities in a continuous culture. ISME J. 2018, 12, 1142–1153. [Google Scholar] [CrossRef] [Green Version]
  68. Sullivan, M.J.; Gates, A.J.; Appia-Ayme, C.; Rowley, G.; Richardson, D.J. Copper control of bacterial nitrous oxide emission and its impact on vitamin B12-dependent metabolism. Proc. Natl. Acad. Sci. USA 2013, 110, 19926–19931. [Google Scholar] [CrossRef] [Green Version]
  69. Ferla, M.P.; Patrick, W.M. Bacterial methionine biosynthesis. Microbiology 2014, 160, 1571–1584. [Google Scholar] [CrossRef]
Figure 1. Maximum-likelihood tree of the 16S rRNA gene sequences showing the position of strain IGB 4-14T among phylogenetically related species of genus Flavobacterium. Filled circles indicate branches of the tree that were also obtained using neighbour-joining algorithm. The sequence GU166749 of Lutibacter flavus IMCC1507T was used as an outgroup. Numbers at branch nodes refer to bootstrap values ≥70%. Bar: substitutions per nucleotide site. Accession numbers (NCBI or IMG database) are indicated in brackets.
Figure 1. Maximum-likelihood tree of the 16S rRNA gene sequences showing the position of strain IGB 4-14T among phylogenetically related species of genus Flavobacterium. Filled circles indicate branches of the tree that were also obtained using neighbour-joining algorithm. The sequence GU166749 of Lutibacter flavus IMCC1507T was used as an outgroup. Numbers at branch nodes refer to bootstrap values ≥70%. Bar: substitutions per nucleotide site. Accession numbers (NCBI or IMG database) are indicated in brackets.
Microorganisms 10 02304 g001
Figure 2. Phylogenomic tree showing the position of strain IGB 4-14T within the genus Flavobacterium. The strain Lutibacter flavus IMCC1507T was used as an outgroup. The maximum-likelihood tree is based on concatenated 120 core marker proteins. Numbers at branch nodes refer to bootstrap values >70%. Bar: amino acid substitutions per position. Assembly accession numbers are indicated in brackets.
Figure 2. Phylogenomic tree showing the position of strain IGB 4-14T within the genus Flavobacterium. The strain Lutibacter flavus IMCC1507T was used as an outgroup. The maximum-likelihood tree is based on concatenated 120 core marker proteins. Numbers at branch nodes refer to bootstrap values >70%. Bar: amino acid substitutions per position. Assembly accession numbers are indicated in brackets.
Microorganisms 10 02304 g002
Figure 3. Maximum-Likelihood based phylogeny of NosZ amino-acid sequences showing the position of strain IGB 4-14T among a selection of type strains of Flavobacterium species and reference strains according to Sanford, et al. [4]. The scale bar indicates the number of amino acid substitutions per site.
Figure 3. Maximum-Likelihood based phylogeny of NosZ amino-acid sequences showing the position of strain IGB 4-14T among a selection of type strains of Flavobacterium species and reference strains according to Sanford, et al. [4]. The scale bar indicates the number of amino acid substitutions per site.
Microorganisms 10 02304 g003
Table 1. Overall genome relatedness indices and 16S rRNA gene similarities calculated for the strain F. azooxidireducens IGB 4-14T and closely related Flavobacterium species.
Table 1. Overall genome relatedness indices and 16S rRNA gene similarities calculated for the strain F. azooxidireducens IGB 4-14T and closely related Flavobacterium species.
Strain IGB 4-14T vs. Species16S rRNA Gene SimilaritydDDH
Values (%)
ANI (%) 2AAI (%)
F. lacus CGMCC 1.12504T (NP180T) 197.2932.286.6889.78
F. piscinae ICH-30T96.9227.884.0187.92
F. orientale CGMCC 1.12506T (SP3T) 196.8024.980.3181.02
F. filum DSM 17961T (EMB34T) 194.5724.180.5283.20
1 Strain number for analysis of 16S rRNA gene similarity if different from the genome analysis. 2 Coverage 37–48%.
Table 2. Whole cell fatty acid compositions of strain F. azooxidireducens IGB 4-14T and the type strain F. lacus NBRC 109715T.
Table 2. Whole cell fatty acid compositions of strain F. azooxidireducens IGB 4-14T and the type strain F. lacus NBRC 109715T.
Fatty Acid 1/StrainIGB 4-14TNBRC 109715T
Saturated
iso-C13:01.32Tr
anteiso-C13:0TrND
C15:0Tr2.4
iso-C15:044.1836.11
anteiso-C15:02.361.1
iso-C15:0 3-OH8.576.2
iso-C17:0 3-OH5.308.49
Unsaturated
iso-C15:1 G17.6521.58
C15:1 w6cTr2.61
iso-C17:1 w9c5.639.88
anteiso-C17:1 w9cTrND
C17:1 w6cNDTr
C17:1 w8cNDTr
Summed features 2
Feature 11.52ND
Feature 31.921.71
Feature 4Tr1.68
TR, trace (<1.0%). ND, not detected. 1 Fatty acids amounting to less than 1.0% of the total fatty acids in both strains are not mentioned in the table: C14:0, iso-C14:0, iso-C14:0 E, C15:0 2-OH, C16:0, iso-C16:0, iso-C16:0 3-OH, iso-C17:0, C17:0 2-OH, anteiso-C15:1 A, iso-C16:1 H, and C18:1 w5c. 2 Summed feature 1: iso-C15:0 H/C13:0 3-OH; summed feature 3: C16:1 ω7c/iso-C15:1 2-OH; summed feature 4: iso-C17:1 I/anteiso-C17:1 B.
Table 3. Physiological characteristics distinguishing strain F. azooxidireducens IGB 4-14T and the type strain F. lacus NBRC 109715T.
Table 3. Physiological characteristics distinguishing strain F. azooxidireducens IGB 4-14T and the type strain F. lacus NBRC 109715T.
CharacteristicIGB 4-14TNBRC 109715T
Flexirubin-type pigments+-
absorption of Congo red+-
Oxidation (GEN III) of:
N-Acetyl-β-D-Mannosamine +-
D-Galactose+-
3-Methyl Glucose+-
D-Fructose-6-PO4+-
L-Arginine +-
L-Histidine+-
Pectin -+
D-Galacturonic Acid +-
L-Galactonic Acid Lactone +-
D-Glucuronic Acid+-
Glucuronamide+-
Mucic Acid-+
p-Hydroxy-Phenylacetic Acid+-
L-Lactic Acid +-
D-Malic Acid-+
L-Malic Acid-+
Propionic Acid+-
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Behrendt, U.; Spanner, T.; Augustin, J.; Zak, D.H.; Horn, M.A.; Kolb, S.; Ulrich, A. Consumption of N2O by Flavobacterium azooxidireducens sp. nov. Isolated from Decomposing Leaf Litter of Phragmites australis (Cav.). Microorganisms 2022, 10, 2304. https://doi.org/10.3390/microorganisms10112304

AMA Style

Behrendt U, Spanner T, Augustin J, Zak DH, Horn MA, Kolb S, Ulrich A. Consumption of N2O by Flavobacterium azooxidireducens sp. nov. Isolated from Decomposing Leaf Litter of Phragmites australis (Cav.). Microorganisms. 2022; 10(11):2304. https://doi.org/10.3390/microorganisms10112304

Chicago/Turabian Style

Behrendt, Undine, Tobias Spanner, Jürgen Augustin, Dominik H. Zak, Marcus A. Horn, Steffen Kolb, and Andreas Ulrich. 2022. "Consumption of N2O by Flavobacterium azooxidireducens sp. nov. Isolated from Decomposing Leaf Litter of Phragmites australis (Cav.)" Microorganisms 10, no. 11: 2304. https://doi.org/10.3390/microorganisms10112304

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop