Skip to main content
mBio logoLink to mBio
. 2024 Oct 30;15(12):e02540-24. doi: 10.1128/mbio.02540-24

Unprecedented N2O production by nitrate-ammonifying Geobacteraceae with distinctive N2O isotopocule signatures

Zhenxing Xu 1,2,, Shohei Hattori 3,4, Yoko Masuda 2,5,, Sakae Toyoda 6, Keisuke Koba 7, Pei Yu 8, Naohiro Yoshida 9,10, Zong-Jun Du 1, Keishi Senoo 2,5
Editors: Derek R Lovley11, Daniel R Bond12
PMCID: PMC11633192  PMID: 39475233

ABSTRACT

Dissimilatory nitrate reduction to ammonium (DNRA), driven by nitrate-ammonifying bacteria, is an increasingly appreciated nitrogen-cycling pathway in terrestrial ecosystems. This process reportedly generates nitrous oxide (N2O), a strong greenhouse gas with ozone-depleting effects. However, it remains poorly understood how N2O is produced by environmental nitrate-ammonifiers and how to identify DNRA-derived N2O. In this study, we characterize two novel enzymatic pathways responsible for N2O production in Geobacteraceae strains, which are predominant nitrate-ammonifying bacteria in paddy soils. The first pathway involves a membrane-bound nitrate reductase (Nar) and a hybrid cluster protein complex (Hcp–Hcr) that catalyzes the conversion of NO2 to NO and subsequently to N2O. The second pathway is observed in Nar-deficient bacteria, where the nitrite reductase (NrfA) generates NO, which is then reduced to N2O by Hcp–Hcr. These enzyme combinations are prevalent across the domain Bacteria. Moreover, we observe distinctive isotopocule signatures of DNRA-derived N2O from other established N2O production pathways, especially through the highest 15N-site preference (SP) values (43.0‰–49.9‰) reported so far, indicating a robust means for N2O source partitioning. Our findings demonstrate two novel N2O production pathways in DNRA that can be isotopically distinguished from other pathways.

IMPORTANCE

Stimulation of DNRA is a promising strategy to improve fertilizer efficiency and reduce N2O emission in agriculture soils. This process converts water-leachable NO3 and NO2 into soil-adsorbable NH4+, thereby alleviating nitrogen loss and N2O emission resulting from denitrification. However, several studies have noted that DNRA can also be a source of N2O, contributing to global warming. This contribution is often masked by other N2O generation processes, leading to a limited understanding of DNRA as an N2O source. Our study reveals two widespread yet overlooked N2O production pathways in Geobacteraceae, the predominant DNRA bacteria in paddy soils, along with their distinctive isotopocule signatures. These findings offer novel insights into the role of the DNRA bacteria in N2O production and underscore the significance of N2O isotopocule signatures in microbial N2O source tracking.

KEYWORDS: nitrate-ammonifying bacteria, DNRA, Geobacteraceae, N2O production, N2O isotopocule signatures, paddy soils

INTRODUCTION

Nitrous oxide (N2O) is a powerful greenhouse gas and the dominant ozone-depleting substance throughout the 21st century (1, 2). The largest current emissions of N2O into the atmosphere are driven by microbial activities in agricultural soils, exacerbated by the application of nitrogen fertilizers (3). Denitrification and nitrification are well-established as the dominant processes, accounting for approximately two-thirds of all soil-derived N2O emissions (4). However, another microbial process, dissimilatory nitrate reduction to ammonium (DNRA), has also been implicated in contributing to global N2O emission (5, 6). DNRA, also termed nitrate (NO3) ammonification, is a nitrogen (N)-retaining process in which water-leachable NO3- is anaerobically reduced to soil-adsorbable NH4+ via the intermediate nitrite (NO2). This reduction is catalyzed by cytoplasmic and/or periplasmic NO3 reductases (Nar and/or Nap) and cytochrome c552 and/or NADH-dependent NO2 reductases (NrfA and/or NirB) in a two-step process (Fig. 1)(7, 8).

Fig 1.

Nitrogen transformation processes in aerobic and anaerobic conditions include pathways for nitrification, denitrification, nitrogen fixation, and DNRA. Key compounds like NH3, NO3−, NO2−, NO, N2O, and N2 are represented with arrows.

Simplified diagram showing N2O production pathways in the nitrogen cycle under both aerobic and anaerobic conditions. SPbD, SPfD, SPNI, and SPDN: 15N-site preference of N2O derived from the bacterial denitrification, fungal denitrification, nitrification, and DNRA processes, respectively. The value ranges of SPbD, SPfD, and SPNI are retrieved from Ref. (9). Key enzymes of the nitrogen cycle are indicated as nitrate reductase (Nar/Nap), nitrite reductase (Nrf/NirB/NirS/NirK), NO dioxygenase (Hmp/Fdp), nitric oxide reductase (Nor), nitrous oxide reductase (Nos), nitrogenase (Nif), ammonia monooxygenase (AMO), hydroxylamine oxidoreductase (HAO), and nitrate oxidoreductase (Nxr). Dotted lines represent the pathways with predicted enzymes and intermediates.

The phenotype of N2O production in DNRA was first documented in 1981, where 163 out of 168 studied nitrate-ammonifying soil bacteria were found to produce N2O (10). Subsequent research has identified many nitrate-ammonifiers, primarily affiliated with Proteobacteria and the genus Bacillus, as N2O producers (1113). The widely accepted pathway for N2O production in these bacteria involves the generation of N₂O from NO₂ via a two-step process, with NO as an intermediate. This process is catalyzed by Nar and NO detoxification-related enzymes, including flavodiiron proteins (Fdp), flavorubredoxin (NorVW), and flavohemoglobin (Hmp), as demonstrated in Nar-containing bacteria Escherichia coli and Salmonella typhimurium (Fig. 1) (1418). However, several studies have reported contradictory findings. For instance, the Hmp and NorV mutations in E. coli did not impair NO reduction (19). Additionally, a recent study involving soil DNRA bacteria suggested that N2O production from Nap and Nar was unlikely, with the NO2-to-NH4+ reaction being a more plausible N2O source (20). This raises questions about the mechanism of N2O generation in Nar-containing ammonifiers. To date, there has been no attempt to explore the N2O generation mechanism in Nap-driven nitrate-ammonifiers (Nar-absent), although their phenotype of N2O production from NO3 has been observed in the DNRA model organism Wolinella succinogenes (21). Therefore, a comprehensive understanding of DNRA-derived N2O production requires further investigation of DNRA bacteria that use Nar versus Nap enzymes for NO3 reduction.

The analysis of N2O isotopocule signatures (δ15Nbulk, δ18O, and 15N-site preference [SP]) is a promising tool for tracing the N2O origins and quantifying the contributions of distinct production processes, as they rely on the natural abundance of N and O isotopes without perturbing in situ conditions (22). Particularly, SP refers to the difference in the 15N isotopic composition of αN compared with the βN in the linear N2O molecule (βN-αN-O)(22). Unlike conventional isotopic modes like δ15Nbulk and δ18O, SP is independent of the concentrations and isotope ratios of substrates, making it a unique indicator for distinguishing N2O production pathways (9). Although the N₂O isotopocule signatures of most known N₂O generation pathways have been documented (Fig. 1) (9, 23), those associated with DNRA remain unexplored. N2O production via DNRA and other pathways, especially denitrification, often occurs simultaneously under anaerobic conditions (6, 24), but few studies have evaluated the contribution of DNRA to N2O emissions due to the lack of a means for distinguishing DNRA-derived N2O from bulk N2O emissions. Characterizing the N2O isotopocule signatures associated with DNRA could enable the partitioning of bulk N₂O emissions into contributions from DNRA and other pathways.

Paddy soils are the largest anthropogenic wetlands on Earth and contribute greatly to global N2O emissions (25, 26). These soils are rich in DNRA drivers due to the reducing conditions caused by waterlogging during rice cultivation (27). Geobacteraceae is a strictly anaerobic and ubiquitous bacterial group in terrestrial ecosystems (28) and one of the dominant nitrate-ammonifying groups in paddy soils (27, 29). Recently, we isolated dozens of Geomonas and Oryzomonas strains (both belonging to the family Geobacteraceae) from paddy soils and nearby sediments (3034). They were genome-annotated as different types of nitrate-ammonifiers, as Geomonas strains mainly contain only Nar or both Nar and Nap for NO3 reduction, while Oryzomonas strains only possess Nap (Fig. 2). The co-presence of Nar- and Nap- driven strains in these closely related genera provides an ideal model for investigating the diverse features of DNRA, including the N2O generation mechanisms.

Fig 2.

Phylogenetic tree with bacterial strains, isolation sources, and nitrogen-related gene functions. Heatmap depicts gene abundance for NO3− to NO2− reduction, NH4+ production, and nitrogen fixation with color intensity representing number of genes.

Phylogenomics and nitrogen metabolism-related gene inventory of Geobacteraceae species. The tree is a maximum-likelihood (ML) phylogeny based on 92 concatenated core proteins in the UBCG database (35). Bootstrap values shown in circle shapes at branching nodes were calculated using 100 replicates. The scale bar represents 0.1 substitutions per amino acid position. The studied genera Geomonas and Oryzomonas were labelled by gray color. The nitrogen metabolism-related gene inventory is shown in the heatmap and the copy numbers of every functional gene are denoted by the blue bar. The accession numbers of genome sequences used in tree construction and detailed information of the strain habitat are listed in Table S4.

In this study, we investigated the N2O production pathways and the key enzymes involved in strictly anaerobic nitrate-ammonifiers Geobacteraceae and synchronously measured the isotopocule signatures of DNRA-derived N2O. We hypothesized that these nitrate-ammonifiers would exhibit novel N2O production pathways or N2O-generating enzymes, given the absence of fdp, norVW, and hmp genes related to NO detoxification reported in enteric facultative anaerobes (1618), and thus possibly create distinctive N2O isotopocule signatures. This study provides a comprehensive description of DNRA-derived N2O production in environmental microorganisms, enhancing our understanding of N2O emission pathways and the role of DNRA microorganisms in greenhouse gas production.

RESULTS AND DISCUSSION

DNRA activity and N2O production in representative Geomonas and Oryzomonas strains

Geomonas and Oryzomonas species were genome-annotated as nitrate-ammonifiers, as they contain the key marker genes nar/nap and nrfA of the DNRA, except for two strains, Geomonas bremensis R1 and Geomonas bemidjiensis DSM 16622T, which lack nar and nap genes and are solely nitrite-ammonifiers (Fig. 2). These strains exhibit various combinations of functional genes related to N metabolism, but these combinations are conserved within phylogenetic clusters at the species level (Fig. 2). We selected three representative species, including two Nar-containing bacteria Geomonas sp. Red32 and Geomonas terrae Red111 and one Nar-absent but Nap-containing bacterium Oryzomonas rubra Red88 to assay DNRA activities using either NO3 (8 mM) or NO2 (2 mM) as substrates.

G. terrae Red111 and O. rubra Red88 were observed to utilize NO3 for growth, producing NH4+, NO2, and N2O as end products, thereby demonstrating DNRA activity (Fig. 3a through c). In contrast, Geomonas sp. Red32 produced only NO2 and N2O without NH4+ production or significant biomass increase (Fig. 3a and d), likely due to its weak NO2-reducing activity, which may be attributed to the single copy of the nrfA gene (Fig. 2). When NO2 was the substrate, G. terrae Red111 and O. rubra Red88 completely consumed NO2, producing NH4+ and N2O along with bacterial growth (Fig. 3e through g). In contrast, Geomonas sp. Red32 only partially consumed NO2, producing N2O but no NH4+ (Fig. 3e and h). None of the studied strains produced NO3 from the added NO2 (Fig. 3f through h), indicating that NO2 oxidation did not occur in any of the cultures. Notably, the N2O produced by these nitrate-ammonifiers accounted for only 1% or less of consumed NO3 and 1%–3% of consumed NO2, far less than the main product NH4+ (>60% of substrates for G. terrae Red111 and O. rubra Red88) and the accumulated NO2 during NO3 reduction (3%–20%) (Fig. 3i). This suggests that N2O production during DNRA is a byproduct rather than a key intermediate or main product.

Fig 3.

Graphs depict biomass growth and nitrogen compound levels over time for different bacterial strains under varying NaNO3 and NaNO2 conditions. Bar chart at the bottom compares nitrogen composition across strains for NO3− and NO2− reduction.

Dissimilatory NO3 and NO2 reduction by three Geobacteraceae strains. (a–h) Dynamics of growth conditions and concentrations of NO3, NO2, NH4+, and N2O in the cultures after NO3 (8 mM, left panel) or NO2 (2 mM, right panel) additions for Geomonas terrae Red111, Oryzomonas rubra Red88, and Geomonas sp. Red32. The start time was from the bacterial inoculation for the NO3 reduction process and from the NO2 addition for the NO2 reduction process. The values are shown as mean ± standard deviation (SD). (i) The relative abundance of N compositions in the cultures at the final reaction stage during NO3 and NO2 reduction by the three strains.

The facultative anaerobic nitrate-ammonifiers have been reported to produce N2O at levels accounting for 3%–36% of consumed NO3 (10, 36, 37), indicating that these Geobacteraceae strains have a relatively low N2O production ratio from NO3 transformation. However, such ratios are comparable to those observed in the strict anaerobe Wolinella succinogenes, which converted approximately 0.15% of NO3 to N2O via DNRA (21). This finding indicates a difference in N2O production capacities between strictly and facultative anaerobic nitrate-ammonifiers, possibly resulting from the difference in the pathways or enzymes involved in N2O production. Moreover, among the biological processes of N2O production, denitrification exhibits the highest N2O production ratio from NO3, as N2O is a key intermediate and can be the sole product, accounting for all consumed NO3, depending on culture conditions and denitrifier types (38). Previous studies have reported that ammonia-oxidizing bacteria (AOB) converted 0.3%–10% of substrate NH3 to N2O and its precursor NO during both NH3 oxidation to NO2 and NO2 reduction to N2O (39). In contrast, ammonia-oxidizing archaea (AOA) and comammox bacteria yield approximately 0.08% N2O from substrate NH3 (40, 41), and anammox bacteria converted 0.1% of substrates NO2 and NH4+ to N2O (42). These findings suggest that nitrate-ammonifiers produce N2O at levels comparable to AOB, and much higher than AOA and comammox and anammox bacteria. Given the wide distribution and diverse species of nitrate-ammonifying bacteria (6), DNRA has great potentials as a contributor to greenhouse gas emission and should not be overlooked.

N2O production coupled to NO2 and NO accumulation during DNRA

To determine the direct substrate contributing to N2O during DNRA, we conducted a comparative analysis of N2O production across different strains and substrates based on prior culture experiments. First, N2O production from NO3 in G. terrae Red111 and O. rubra Red88 suggested five potential sources for N2O: NO3, NO2, NH4+, and the intermediate reactions between NO3 and NO2, and between NO2 and NH4+ (Fig. 3b and c; Path_1 in Fig. 4a). Given that N2O production also occurred from NO2 in G. terrae Red111 and O. rubra Red88 (Fig. 3f and g), we ruled out NO3 and the intermediate processes between NO3 and NO2 as N2O sources (Path_2 in Fig. 4a). Then, Geomonas sp. Red32 partially consumed NO3 and NO2 with concurrent N2O generation, but no NH4+ was detected in the medium (Fig. 3d and h), eliminating NH4+ and the intermediate process between NO2 and NH4+ as N2O sources (Path_3 in Fig. 4a). Collectively, NO2 remains the most likely substrate for N2O (or its precursor) production during DNRA (Fig. 4a).

Fig 4.

Nitrogen pathways lead to N2O production with graphs comparing NO2− and N2O levels over time between control and acetate-treated cultures. Violin plots display distribution of nitrogen isotope ratios for NO2−, NO3−.

Comparison of N2O production and SP values showing the unique N2O production from NO2 via the intermediate NO. (a) A comparison diagram showing the possible N2O production source on DNRA. Three pathways, labeled by Path_1, Path_2, and Path_3, represent the experimental results from the NO3 reduction by all studied Geomonas strains except for Red32, NO2 reduction by all studied Geomonas strains except for Red32, and NO3 reduction by strain Red32, respectively. The red dots on the arrows indicate the possible source of N2O production. (b and c) Dynamics of NO2 and N2O concentrations during the NO3 reduction process driving by Geomonas terrae Red111 with (gray squares) and without (black circles) acetate (5 mM) addition. Arrows indicate the time when the extra acetate was added to the medium. (d and e) N2O production from the NO3 reduction process driven by Geomonas strain Red111 and Oryzomonas strain Red88 across time with (gray squares) and without (black circles) c-PTIO additions. c-PTIO was added to the medium at the same time to bacteria inoculation. CK indicates control experiments, asterisks indicate significant difference, *** P < 0.001, *P < 0.05, ns indicates insignificant difference, P > 0.05. The values are shown as mean ± standard deviation (SD). (f and g) The comparison of SP values between different genera (Geomonas and Oryzomonas) and substrates (NO3 and NO2). Med and n indicate the median and number of every data set, respectively.

To verify the relationship between NO2 and N2O in DNRA, we compared the production and consumption rates of N2O and other N compounds in cultures. N2O production was slower than NO3 consumption and NH4+ production (Fig. 3b and c) but comparable to NO2 accumulation with significantly positive linear relations (R2>0.7, P < 0.0001) (Fig. S1). Additionally, more carbon sources (high C:N ratios) were found to reduce the NO2 accumulation during NO3 reduction in G. terrae Red111 (Fig. S2). In a further study, the absence of NO2 accumulation halted N2O production (Fig. 4b and c). These results indicate that NO2 is the direct substrate contributing to N2O production (or its precursor) during DNRA in Geobacteraceae strains, consistent with previous findings that N2O is generated from NO2 in nitrate-ammonifying Enterobacteriaceae (15, 16, 36).

Previous studies have reported that NO is produced before N2O in nitrate-ammonifying Enterobacteriaceae (1418), raising the question of whether intermediates, such as NO and hydroxylamine, exist between NO2 and N2O in Geobacteraceae strains. We measured hydroxylamine concentrations in the cultures and used NO inhibitors (c-PTIO and L-NMMA) and donors (SNP) for treatments. c-PTIO, a NO scavenger, removes NO from the medium, while L-NMMA, a NO synthase inhibitor, prevents endogenous NO generation from amino acids (43, 44). Hydroxylamine was undetected during bacterial growth, but c-PTIO (100 µM) significantly inhibited N2O production (P < 0.05; Fig. 4d and e). Notably, c-PTIO had no impact on bacterial growth or the DNRA rate, even at high concentrations of 300 µM (P > 0.05, Fig. S3), and L-NMMA (100 µM) did not affect N2O production (P > 0.05, Fig. S4). These results indicate that c-PTIO inhibits N2O production by reducing exogenous NO (derived from supplied NO3) rather than by affecting bacterial activity. Furthermore, G. terrae Red111 and O. rubra Red88 converted approximately half of the supplied extra NO (80 µM SNP solution) to N2O (Fig. S5), demonstrating their ability to reduce NO to N2O. Altogether, we believe that NO is an intermediate between NO2 and N2O, contributing partially or entirely to N2O production during DNRA. Notably, c-PTIO reduced N2O emission by 39.9% in Oryzomonas strains, but only by 19.8% in Geomonas strains (Fig. 4b and c). This difference suggests that Oryzomonas and Geomonas strains probably contain different pathways or enzymes for converting NO3 to N2O.

The process proposed for N2O production from NO2 via NO during DNRA is plausible for Geobacteraceae strains. However, three questions remain unanswered: 1) Is NO2 the sole substrate contributing to N2O in all studied strains? 2) Are there other N2O production pathways besides the NO-mediated one? 3) Does N2O reduction occur during N2O generation in the studied strains? To address these questions, we measured and analyzed SP values, which depend solely on the structures of enzymatic precursors right before N2O production (22). The SP values of N2O produced from NO3 ranged from 44.8‰ to 47.7‰ in 3-day cultures and from 40.1‰ to 46.6‰ in 7-day cultures, while those from NO2 ranged from 43.4‰ to 49.9‰ in 3-day cultures and from 43.0‰ to 47.3‰ in 7-day cultures (Table 1; Table S1). Comparing different substrates, the median SP values for NO2- and NO3- derived N2O were 46.2‰ and 46.4‰, respectively, with no significant difference (P > 0.05) (Fig. 4f). Similarly, the SP values were insignificant (P > 0.05) among strains in the two genera Geomonas and Oryzomonas, with close medians of 46.4‰ and 46.2‰, respectively (Fig. 4g). These similar SP values indicate a shared N2O production pathway in these strains, regardless of whether NO2 or NO3 is used as the substrate. Additionally, to prove the uniqueness of the N2O production pathway, we compared the SP values of N2O collected at different reaction times, 3 and 7 days. Geomonas sp. Red32 and O. rubra Red88 showed constant SP values for both NO3 and NO2 reduction processes, whereas G. terrae Red111 only showed consistent SP values for NO2 reduction (Table 1). It is known that mixed N2O production/reduction pathways can alter their relative contribution ratios over time, thereby changing SP values. However, the consistent SP values observed in this study refute this proposal and demonstrate a single, consistent pathway for N2O generation in the studied strains, except for NO3 reduction driven by G. terrae Red111, where mixed processes for N2O production may occur (see the last section for explanation). These findings prove a unique N2O production pathway (NO2-NO-N2O) without N2O reduction in the studied nitrate-ammonifiers, addressing the questions above.

TABLE 1.

Isotopocule signatures of N2O produced from NO3 and NO2 reduction processes by Geobacteraceae strains and abiotic reactionsa

Substrate Strainb Incubation time (d)c δ18O (‰) δ15Nbulk (‰) SP (‰)
NO3- Geomonas sp. Red32 3 16.2 (0.1) −39.8 (0) 46.5 (0.1)
(8 mM) 7 14.6 (0.4) −27.5 (1.2) 45.4 (1.4)
G. diazotrophica Red69 3 26.4 (4.1) −29.6 (4.5) 45.5 (0.7)
7 23.6 (3.9) −28.2 (4.9) 41.9 (0.9)
G. terrae Red111 3 15.2 (1.2) −19.7 (2.1) 47.0 (0.3)
7 9.6 (0.7) −23.0 (3.4) 41.1 (1.4)
G. silvestris Red330 3 18.6 (4.1) −21.1 (7.0) 46.8 (1.3)
7 15.5 (1.5) −18.0 (7.2) 46.5 (0.1)
O. rubra Red88 3 23.4 (0.8) −31.4 (3.7) 46.3 (0.6)
7 19.2 (1.5) −22.4 (5.4) 46.4 (0.3)
NO2- Geomonas sp. Red32 3 27.8 (0.2) −23.7 (0) 46.8 (0.2)
(2 mM) 7 27.0 (1.0) −21.3 (0.1) 46.1 (1.0)
G. diazotrophica Red69 3 28.8 (0.1) −23.7 (0) 45.8 (0.7)
7 28.6 (0) −23.5 (0.3) 46.0 (0.1)
G. terrae Red111 3 27.5 (0.3) −24.3 (0.1) 47.2 (0.7)
7 29.5 (0) −23.9 (0.1) 46.3 (0.2)
G. silvestris Red330 3 28.5 (0.4) −22.5 (0.5) 46.5 (0.2)
7 28.6 (0.2) −21.6 (0.5) 46.4 (0.2)
O. rubra Red88 3 28.8 (0.4) −21.6 (0.7) 44.8 (2.0)
7 28.8 (0.3) −21.0 (0.1) 44.4 (2.0)
G. bemidjiensis DSM 16622 3 28.6 (0.6) −22.9 (0.2) 47.4 (2.3)
Geob. anodireducens JCM 30203 3 28.7 (0.7) −23.0 (0) 46.7 (0.7)
Abiotic 1 20.1 (0.2) −28.8 (0.4) 20.0 (1.2)
a

All N2O isotope values are the isotopic composition of the N2O produced, the δ15N and δ18O values are the difference from those of the substrate NO315NNO3- = 2.4 ± 0.1‰, δ18ONO3- = 17.9 ± 0.3‰) or NO215NNO2- = −1.8 ± 0.3‰, δ18ONO2- = 3.7 ± 0.2‰). The values in parentheses represent the standard deviation obtained from replicate experiments (refer to the raw data in Table S1).

b

Abiotic represents the N2O production from a chemical reaction without bacteria inoculation under low pH conditions (pH < 2).

c

Sample collection at different incubation times, the start time is from that the substrate NO3 or NO2 was injected into the cultures.

Enzymatic pathways of N2O production during DNRA

To elucidate the enzymatic pathways of nitrate-ammonifying N2O production, we cultured G. terrae Red111 (containing both Nar and Nap) and O. rubra Red88 (containing only Nap). Initially, abiotic sources of N2O were excluded, as little or no N2O production was observed in bacteria-free and autoclaved cultures (Fig. 5a and b), consistent with the differing SP values in nitrate-ammonifiers from the abiotic reaction (Table 1). In contrast, N2O production was observed in filtered cultures, where live bacteria had been removed, but free and partial periplasmic enzymes were present, suggesting the enzymatic reactions responsible for N2O production during DNRA (Fig. 5a and b). Growth curves and fluorescence microscopy revealed that kanamycin (100ug/mL) reduced biomass and increased cell mortality of G. terrae Red111 and O. rubra Red88 in non-concentrated cultures (Fig. S6). Consequently, kanamycin was used to prepare cell lysates. Kanamycin addition significantly increased N2O production in G. terrae Red111 (P < 0.01) but decreased it in O. rubra Red88 (P < 0.01) (Fig. 5a and b). This increase in G. terrae Red111 possibly resulted from the release of more intracellular enzymes, whereas the decrease in O. rubra Red88 may be attributed to the cessation of extracellular/periplasmic enzyme synthesis, suggesting different enzyme profiles between these strains, consistent with their varying responses to c-PTIO (Fig. 4b and c). Moreover, kanamycin slightly increased N2O production with SNP addition in both strains (P > 0.05) (Fig. S5), suggesting the intracellular enzymes in both strains facilitating the reduction of NO to N2O. The addition of oxygen significantly decreased N2O production in both strains with NO2 or SNP (P < 0.001, Fig. 5a and b; Fig. S5), implying that an oxygen-sensitive enzyme is responsible for N2O production, aligning with the anaerobic metabolism of DNRA.

Fig 5.

Bar charts compare N2O production under various treatments for G. terrae and O. rubra. Heatmap displays gene expression changes for NO3− vs CK and NO2− vs CK. Line graphs plot N2O production over time. Final line graph compares two strains for N2O levels.

Comparisons of N2O productions, gene expressions, and gene structures of Geobacteraceae strains. (a and b) N2O production from the NO2 reduction process driven by strains Red111 and Red88 under different treatments: Abiotic, bacteria-free medium; CK, bacteria-containing medium; Filter, filtered bacteria-containing medium; Autoclaved, autoclaved bacteria-containing medium; Oxygen, bacteria-containing medium gassed air; Kan+, bacteria-containing medium supplemented with 100 µg/mL kanamycin; Kan + Oxy, bacteria-containing medium gassed air and supplemented with 100 µg/mL kanamycin. The default medium is NNFM_N. (c) The expression levels of nitrogen metabolism related genes from transcriptome analysis in Geomonas sp. Red32 with NO3 (8 mM, left panel) and NO2 (2 mM, right panel) inductions. Red colors represent the upregulated genes, while blue colors represent the downregulated genes. Asterisk indicates significant difference, adjusted P < 0.05. (d and e) The expression levels of representative genes related to DNRA and nitrosative stress processes from RT-qPCR in strains Red111 and Red88 with NO3 (8 mM, right panel) and NO2 (2 mM, left panel) inductions for 4 and 24 h. (f and g) N2O production from the NO2 reduction process driven by strains Red111 and Red88 over time with (grey squares) and without (black circles) chlorate additions. (h) structure comparison of DNRA-related genes (nap and nrf clusters) in three Oryzomonas strains, Red88, Red96, and Red100. Arrows indicate open reading frames and their orientations in the genomes. (i) N2O production from the NO2 reduction process driven by nar/nap-absent strains Geomonas bemidjiensis DSM 16622 and Geobacter anodireducens JCM 30203 at different culture times. Asterisks indicate significant difference, *** P < 0.001, **P < 0.01, ns indicates insignificant difference, P > 0.05.

To identify the key enzymes catalyzing N2O production, we quantified the expression levels of genes related to DNRA and nitrosative stress processes under NO3, NO2, and SNP treatments in G. terrae Red111 and O. rubra Red88. The target genes were identified based on transcriptomic results from Geomonas sp. Red32, which showed high conversion ratios from NO3 and NO2 to N2O (Fig. 3d and h) and was selected for transcriptomic study (Table S2). A total of 1,699 and 1,134 differentially expressed genes were detected in Geomonas sp. Red32 under NO3 and NO2 treatments, respectively (Fig. S7). Genes related to nitrogen fixation and ammonium assimilation were globally downregulated under both NO3 and NO2 treatments, whereas genes involved in DNRA and nitrosative stress were upregulated (Fig. 5c; Table S3). Given that Geomonas sp. Red32 produced N2O following NO3 or NO2 addition (Fig. S8), the upregulated genes in the DNRA and nitrosative stress pathways are connected to N2O production.

In G. terrae Red111 and O. rubra Red88, most genes associated with DNRA and nitrosative stress pathways were significantly upregulated, except for the genes napA, norB, and qnor in G. terrae Red111 and norB and qnor in O. rubra Red88, which exhibited low or negative expression levels (Fig. 5d and e; Fig. S9). Denitrifying N2O production driven by cNor (encoded by norBC) and qNor has been reported with SP values less than 0‰ (23, 45), significantly lower than the SP values observed for DNRA-derived N2O. These findings excluded the possibility of cNor and qNor involved in NO reduction in Geobacteraceae strains. Interestingly, the hcp-hcr gene cluster (encoding the hybrid cluster protein complex, Hcp–Hcr) in the nitrosative stress group showed high expression levels under all conditions (Fig. 5c through e; Fig. S9). Hcp–Hcr is an unusual redox endoenzyme complex with oxygen-sensitive hydroxylamine and NO reductase activity (46), and its function in NO detoxication to N2O has been confirmed by Hcp mutants in Methylobacter species and E. coli (47, 48). Moreover, hcp-hcr is the only remaining gene cluster related to NO detoxification present in all Geobacteraceae genome sequences (Fig. 2). Given the constant SP values suggesting similar enzymes catalyzing NO to N2O in different Geobacteraceae strains, we propose the Hcp–Hcr complex as the primary driver of NO reduction to N2O in these strains. This finding is distinct from the reported enteric nitrate-ammonifiers that use NO reductase (NorVW) or NO dioxygenase (Hmp or Fhp) for NO detoxication (16, 17).

For the enzyme catalyzing NO2 to NO, we first assume Nar in Geomonas strains, due to its high expression and central role in the gene co-expression network under the NO2- treatment in Geomonas sp. Red32 (Fig. 5c and d; Fig. S10). Although Nar is known for catalyzing NO3 to NO2, its role in catalyzing NO2 to NO is less established. To investigate this, we added the Nar inhibitor chlorate (ClO3) to the cultures (49). ClO3 showed little effect on bacterial growth and NO reduction to N2O but completely inhibited NO3 reduction in Geomonas diazotrophica Red69 (Nar-containing but Nap-absent) and G. terrae Red111 (Nar- and Nap-containing), with no effect on Nap-driven O. rubra Red88 (Fig. S11 and S12). This indicates the Nar-driven type (Nap deactivated) of G. terrae Red111 related to DNRA. ClO3 addition further completely inhibited N2O production from NO2 in G. terrae Red111 but had no effect on O. rubra Red88 (Fig. 5f and g), underscoring the critical role of Nar in N2O production. After numerous unsuccessful attempts to construct genetic mutants of Geobacteraceae strains, we moved to use another nitrate-ammonifying strain, E. coli MG1655, as an alternative and constructed mutant strains MG1655 ∆narG and MG1655 ∆napA. Compared with the wild type, strain MG1655 ∆narG produced much less N2O from NO2, while MG1655 ∆napA produced similar amounts (Fig. S13). Nar is an oxygen-sensitive membrane-bound enzyme in which the catalytic subunit faces the cytoplasm (50), meeting the proposed characteristics of intracellular enzymes. Thus, Nar is proposed as the enzyme responsible for NO2 to NO conversion in Nar-containing Geomonas strains.

For Nar-deficient Oryzomonas strains, Nap and NrfA are the potential candidates for catalyzing NO2 to NO, as both are periplasmic enzymes and showed high expression levels in O. rubra Red88 under both NO3 and NO2 treatments (Fig. 5e). To determine the responsible one, we cultured the other two bacteria Oryzomonas japonica Red96 and Oryzomonas sagensis Red100 and compared their NO3 and NO2 reduction products. All three strains could produce N2O from NO3 reduction, but only O. rubra Red88 produced N2O from NO2 reduction (Fig. 3c and g; Fig. S14), indicating that the target enzyme is present in all three strains but in difference from O. rubra Red88 to O. japonica Red96 and O. sagensis Red100. Gene cluster comparisons revealed that the napDAB cluster is structurally conserved and shares high sequence similarities (>95%) among the three strains, whereas the nrfAH cluster was distinctly different, as O. rubra Red88 harbors two paralogs of nrfAH (nrfAH1Red88 and nrfAH2Red88), and O. japonica Red96 and O. sagensis Red100 harbor only one (nrfAH1Red88 orthologous) (Fig. 5h; Fig. S15). Moreover, the gene nrfAH2Red88 showed higher gene expression than nrfAH1Red88 under NO3 and NO2 treatments (Fig. 5e), highlighting the difference of NrfA from O. rubra Red88 to O. japonica Red96 and O. sagensis Red100. These comparisons indicate that NrfA is more likely than Nap to be involved in N2O production. In addition, two other Geobacteraceae strains, G. bemidjiensis DSM 16622 and Geobacter anodireducens JCM 30203, which lack both nar and nap genes and are incapable of NO3 reduction, also produced N2O from NO2 reduction with comparable N2O ratios and SP values to other studied nitrate-ammonifiers (Table 1; Fig. 5i; Fig. S16). This finding excluded the necessity of Nap in N2O production. With these results and the reported crystal structure of NrfA that predicted NO occurrence during NO2 reduction to NH4+ (51), NrfA is supposed as the target enzyme catalyzing NO production in Nap-driven nitrate-ammonifiers.

Although Geobacteraceae strains contain one or more copies of the nrfA operon, their phenotypes with respect to NO2 reduction are not consistent. For example, Geomonas sp. Red32, O. japonica Red96 and O. sagensis Red100 each contain a single copy of nrfA, but only Geomonas sp. Red32 cannot produce NH4+ from NO3. This variability in NrfA functionality suggests evolutionary divergence among these strains. Phylogenetic analyses based on NrfA sequences reveal that the NrfA proteins of Geobacteraceae strains are distributed within Clade I and the outgroup (52), displaying phylogenetic differences at the single species level (Fig. S17). Notably, in Oryzomonas strains (Red96, Red100, and Red88-NrfA1), NrfA forms a robust branch within Clade I-I (bootstrap value 0.9) and shares <75% similarity at the amino acid level with other NrfA enzymes. This sequence specificity of NrfA in Clade I-I likely underlies the differential activities observed between Oryzomonas and Geomonas strains. Most Geomonas strains contain both Nar and NrfA enzymes; however, there is currently no evidence supporting the involvement of these Nrf enzymes in N2O production. Previous studies have shown that nitrate-ammonifying strains deleting nrfA (Nar-containing type) do not influence N2O production (16, 17). Given that the Nar complex contains specialized NO2/NO3 membrane transporters (NarK)(53), we assume that Nar transforms pericellular NO2 in high concentration into bacterial cells, thereby alleviating NO2 pressure on NrfA proteins and inhibiting NO release from NrfA enzymes in Nar-containing bacteria. Although the enzymatic combinations Nar-Hcp and NrfA-Hcp related to DNRA-derived N2O production were functionally identified in Geobacteraceae strains, these gene combinations narG-hcp and nrfA-hcp are widely distributed across the bacterial domain, including within the phyla Pseudomonadota, Bacteroidota, Desulfobacterota, and Bacillota (Fig. S18). Thus, this study reveals a previously unrecognized, yet potentially widespread, N₂O production pathway in various bacteria.

Distinctive isotopocule signatures of DNRA-derived N2O with the role in quantifying N2O production processes

The isotopocule signatures of DNRA-derived N2O (except abnormal values from mixed processes) range from 43.0‰ to 49.9‰ for SP, −39.9‰ to −5.8‰ for δ15Nbulk, and 14.1‰ to 30.5‰ for δ18O (Table 1; Fig. 6a through c). These values are notably distinctive compared to those from other N2O generation processes (9), particularly the SP values, which surpass those reported for any other processes (Fig. 6a and b), suggesting that SP is a robust and independent tool for identifying DNRA-derived N2O. In contrast, the δ15Nbulk and δ18O values (isotopic fractionation from substrates) of the produced N2O show greater variability (δ15Nbulk = −24.6 ± 6.0‰, δ18O = 23.6 ± 6.3‰) than the SP values and overlap with values from other known N2O production processes (Table 1; Fig. 5a through c), indicating their limited effectiveness as indicators for N2O partitioning. However, the combined δ18O/δ15Nbulk plot enables clear differentiation of DNRA-related N2O from other processes, including nitrification, nitrifying-denitrification, and fungal denitrification (Fig. 6c). This highlights the utility of dual isotope plots in distinguishing N2O sources. Previous studies have noted that the N2O reduction process mediated by N2O reductase (NosZ) can elevate SP values appreciably (54), which may obscure the identification of high SP value processes. Our results show that the slope of SP/δ15Nbulk for residual N2O caused by N2O reduction from nitrification and fungal denitrification sources overlaps with the DNRA zone (Fig. 6a), while this slope of SP/δ18O does not overlap with the DNRA zone (Fig. 6b). Thus, SP/δ18O provides a more precise means of distinguishing DNRA-derived N2O from other pathways, offering enhanced resolution compared to SP alone. Indeed, high SP values over 45‰ have been reported in a nitritation-anammox reactor (55), which could be attributed to DNRA-derived N2O as characterized in this study.

Fig 6.

Isotope plots depict relationships between δ15N and δ18O values of N2O for different strains and substrates over time. Box plot compares 15N-SP between strains at 3 and 7 days, with statistical significance noted.

Comparisons of N2O isotopocule signatures from different N2O production reactions. (a–c) SP/δ15Nbulk, SP/δ18O, and δ18O/ δ15Nbulk plots showing the isotopocule signatures of N2O from various nitrogen metabolism processes. nD, nitrifier denitrification; bD, bacterial denitrification; fD, fungal denitrification; cD, chemical denitrification; NI, nitrification; DN, DNRA. The black arrows denote the N2O reduction (N2O-R) by bacterial denitrification based on early studies [slope ε(SP) / ε15N) = 0.96, slope ε(SP) / ε18O) = 0.45, slope ε18O) / ε15N) = 2.21]. The value ranges (except for DNRA obtained in this study) are retrieved from Ref (9). (d) SP value comparisons among strains Red69/Red111 (7 days), strain Red330 (7 days), and all Geomonas strains (3 days) with the substrate NO3. The raw data are presented in Table S1.

N2O isotopocule signatures also serve as tools for identifying and quantifying mixed N2O production processes through isotope mass balance. For instance, the N2O production from NO3 reduction by G. terrae Red111 at 7-day culture suggests mixed processes, given its lower SP values (41.1‰–41.9‰) than others (Fig. 6d). To elucidate the contributing processes, we introduced another two strains, G. diazotrophica Red69 and Geomonas silvestris Red330, and measured their N2O isotopocule signatures. Both G. terrae Red111 and G. diazotrophica Red69 showed similar decreases in SP value over prolonged culture periods, whereas G. silvestris Red330 did not (Table 1), despite all strains showing similar NO3/NO2 reduction rates (Fig. S18). This discrepancy indicates the differences between G. silvestris Red330 and those two strains causing the decrease in SP values. Genetic comparison revealed that G. diazotrophica Red69 and G. terrae Red111 contain norBC clusters, while G. silvestris Red330 does not (Fig. 2). cNor has been documented to catalyze NO to N2O during denitrification, with known SP values ranging from 7.5‰ to 3.7‰ (9). Given that NO is a precursor to N2O in DNRA, the partial involvement of cNor in G. diazotrophica Red69 and G. terrae Red111 likely contributes to the observed SP value reductions over time. Quantitative analysis of mixed N2O production processes revealed that DNRA-related enzymatic reactions account for 93.9% and 94.0% of N2O production during NO3 reduction at the 7-day culture of G. diazotrophica Red69 and G. terrae Red111, respectively. This approach has also been applied to quantify four mixed N2O production/reduction processes using N2O isotope mass balance (56, 57), displaying its significant potentials for environmental research. However, estimating pathways based on only two or three N2O isotopic dimensions can lead to an underdetermined set of equations. Thus, additional measurements, such as microbial ecological data or pathway modelling approaches, are necessary for accurate determination of extra pathways.

Previous studies on pure cultures of nitrifiers and fungal denitrifiers have demonstrated that N2O formed from asymmetrical precursors (N-N-O) typically exhibits high SP values (>20‰) owing to sequential binding mechanisms, whereas low SP values around 0‰ occurring in bacterial denitrification involve dimerization of NO (23, 58). The sequential reaction of asymmetrical precursors is thus supposed to account for the synthesis of DNRA-derived N2O molecules. Although the mechanism for the high SP values in DNRA-derived N2O is not yet fully elucidated, we hypothesize that specific enzyme activities or catalytic centers govern the sequential binding of NO, resulting in distinct fractionations during the addition of the first or second NO precursor. Further studies involving enzymatic structures is needed to bridge this gap. Coincidentally, the predicted equilibrium SP value of N2O at room temperature was reported as 45‰ based on theoretical modeling (59), comparable to that of DNRA-derived N2O. We thus believe that the high SP values of DNRA-derived N2O may provide novel insight into N2O formation mechanisms in follow-up studies.

Our results elucidate the N2O production pathways (NO2-NO-N2O) in both Nar- and Nap-type DNRA bacteria within the Geobacteraceae family. We identified two novel enzymatic pathways for N2O generation in nitrate-ammonifiers: 1) Nar and the Hcp–Hcr complex catalyzed the NO2-NO-N2O process in Nar-type nitrate-ammonifiers; 2) NrfA and the Hcp–Hcr complex catalyzed the NO2-NO-N2O process in Nap-type nitrate-ammonifiers (Fig. 7). Furthermore, DNRA-derived N2O exhibits the highest SP values reported to date across various N2O production pathways in pure culture experiments. Cross plots of SP with δ15Nbulk and δ18O values clearly distinguish DNRA-derived N2O from other known production pathways. These results are crucial for precisely modeling global greenhouse emissions. Nevertheless, the limited nitrate-ammonifiers in one group of Geobacteraceae are insufficient to obtain a comprehensive understanding of DNRA-derived N2O production, as environmental nitrate-ammonifiers exhibit a range of enzyme combinations related to N metabolism. On this basis, large-scale nitrate-ammonifier screening with N2O production pathway parsing will be considered in future studies.

Fig 7.

N2O production in soil nitrate-ammonifying bacteria contrasts Nar-type and Nap-type ammonifiers. It depicts pathways of NO3−, NO2−, NH4+, and N2O within the periplasm and cytoplasm, along with enzymes like NrfA and NapAB.

Schematic description of DNRA-derived N2O production. (a) N2O production pathway in Nar-driven Geomonas strains. (b) N2O production pathway in Nap-driven Oryzomonas strains. Red arrows with relative thin lines indicate the N2O production as a byproduct from NO2 during the DNRA process. Key enzymes are indicated as nitrate/nitrite antiporter (NarK), respiratory nitrate reductase (NarGHI), periplasmic nitrate reductase (NapAB), cytochrome c nitrite reductase (NrfA), and hybrid cluster protein complex (Hcp–Hcr). The enzymes in different colors denote different enzyme types, green labels the exoenzyme (Nap) or the outer membrane protein (Nrf), brown labels the endoenzyme (NarGHI and Hcp–Hcr complex), and blue labels the membrane transport proteins (NarK and others). The shown ranges of SP values were summarized from the NO3/NO2 reduction process driven by Geomonas or Oryzomonas strains in this study.

MATERIALS AND METHODS

Bacterial strains and cultivation

Nine bacterial strains of the family Geobacteraceae and a reference strain E. coli K-12 MG1655 were cultured for this study. Among these, Geomonas sp. Red32 was recently isolated from paddy soils and has been deposited in the Japan Collection of Microorganisms (JCM 33031) and Marine Culture Collection of China (MCCC 1K03692). G. bemidjiensis DSM 16622 was obtained from the Deutsche Sammlung von Mikroorganismen und Zellkulturen (Braunschweig, Germany), while Geobacter anodireducens JCM 30203 was sourced from the Japan Collection of Microorganisms (Tsukuba, Japan). Others, including three Geomonas strains Red69, Red111, and Red330 and three Oryzomonas strains Red88, Red96, and Red100, were described in our previous works (3032, 34) and reactivated for this study. E. coli K-12 MG1655 was acquired from the National BioResource Project (NBRP)-E. coli (Mishima, Japan). Gene knockout mutants of E. coli K-12 MG1655 were constructed as described previously (60) and hereafter termed as MG1655 ∆nar and MG1655 ∆nap, representing the markerless narG and napA deletion mutants, respectively.

Unless otherwise specified, the nine Geobacteraceae strains were routinely cultured in 50 mL glass serum bottles containing 20 mL of anoxic nitrogen-free freshwater medium (NFFM, pH 6.8) supplemented with 10 mM acetate as the electron donor and carbon source and 10 mM fumarate (NFFM_F) or nitrate (NFFM_N) as the electron acceptor. Cultures were incubated at 30°C without shaking. The composition of NFFM was consistent with that described previously (34), except that NH4Cl was omitted due to the N2-fixing capability of these strains. The only non-N2-fixing strain, Geob. anodireducens JCM 30203, was cultured equally with the addition of 5 mM NH4Cl to the medium. Serum bottles containing NFFM were neck-sealed with butyl-rubber stoppers and aluminum crimps, with the gaseous phase replaced by N2/CO2 (80:20, v/v). E. coli K-12 MG1655 and its mutants were routinely cultured using aerobic LB broth (Nihon Pharmaceutical, Japan) at 37°C with shaking at 120 rpm and anaerobically cultured when monitoring NO2 reduction using LB broth supplemented with 2 mM NO2 .

Measurement of NO3 and NO2 reduction to ammonium

NO3 reduction experiments were conducted using the strains cultured in NNFM_N, while NO2 reduction experiments involved culturing the strains in NNFM_F for 2 days until the biomass reached an optical density of >0.02 (OD600) and then 2 mM NaNO2 was injected into the bottles. The sampling was started from the time of NaNO2 addition. For sample collection, 1 mL of culture was withdrawn from each bottle at designated time intervals, and the bacterial biomass was measured at OD600. After that, the samples were filtered through 0.22 µm syringe filters prior to preserved in a freezer (−20°C) for subsequent measurement. The concentrations of NO3 and NO2 were measured using an ion chromatograph (Dionex ICS-900) equipped with a Dionex IonPac AS12A analytical column (Thermo Fisher Scientific, USA). NH4+ concentrations were determined using the indophenol-blue method followed by colorimetry (61). The produced N2O in the headspace gas of the cultural bottles (1 mL) was measured using a gas chromatograph equipped with a Porapack Q column and an electron capture detector (GC-ECD, GC-2014, Shimadzu, Japan). The temperature of the injector, detector, and column were 90°C, 345°C, and 80°C, respectively, with an Ar/CH4 carrier gas flow rate of 150 mL/min. The N2O dissolved in the liquid medium was calculated using the Bunsen absorption coefficient (α = 0.544, 25°C)(62). The effect of carbon-to-nitrate (C:N) ratios on DNRA activities and bacterial growth was investigated using G. terrae Red111 in NNFM_N for 1 week. Different C:N ratios were prepared using a constant NO3 concentration of 10 mM as the nitrogen source and varying acetate concentrations (5–40 mM) as the carbon source.

Control experiments to explore N2O production pathways

To assess the contribution of NO to N2O, a NO scavenger, 2-(4-carboxyphenyl)−4,4,5,5-tetramethylimidazoline-1-oxyl-3-oxide (c-PTIO, Santa Cruz Biotechnology, USA), was applied during bacterial culture in NNFM_N. The effect of c-PTIO on bacterial growth was evaluated using G. terrae Red111 at various concentrations (0, 50, 100, and 300 µM). NO inhibition tests were performed using 100 µM c-PTIO and quantified by the comparison of N2O concentration with c-PTIO-free cultures. L-NG-monomethyl arginine acetate (L-NMMA, Dojindo, Japan), used to investigate the role of endogenous NO in N2O production, was applied at a work concentration of 100 µM. Hydroxylamine concentrations were measured as described previously (63). The effect of kanamycin (100 μg/mL) on bacterial growth was monitored using growth curves and cell viability assays, with the latter performed via fluorescence microscopy (Carl Zeiss Axio Scope A1, Germany) following bacterial staining with a live/dead cell viability assay kit (Bestbio, China). For treatments under filtered, autoclaved, oxygen-containing and kanamycin-supplemented conditions, cultures were prepared using NNFM_N inoculated with target strains, and then dispensed into 30 serum vials (5 mL culture in 20 mL vials, five replicates/treatment) of each strain once the biomass reached an OD600 >0.02. Filtered cultures were treated using 0.22 µm syringe filters, autoclaved cultures were prepared by autoclaving at 121°C for 20 min, and kanamycin-supplemented cultures were prepared by injecting kanamycin solution to a final concentration of 100 µg/mL. The gaseous phase in the vials was then replaced with N2/CO2 (80:20, v/v), except for the oxygen-containing cultures. After that, 2 mM NO2 or 80 µM sodium nitroprusside (SNP, Sigma, USA; NO donor) was injected into sub-cultures or cell-free media as required, which were then cultured at 30 °C for 3 days before N2O sampling. The effect of sodium chlorate (NaClO3; 20 mM) on Nar, Nap, and both Nar and Nap containing strains, G. diazotrophica Red69, O. rubra Red88, and G. terrae Red111, respectively, was examined using NNFM_N and NNFM_F supplemented with 2 mM NO2, the consumed and produced N compounds were measured as described above. ClO3 concentrations were determined using an ion chromatograph as described above.

Isotopic analysis of N2O, NO3, and NO2

The N2O isotopocule ratios were measured using an automated system containing an autosampler (GX-222 XL liquid handler; Gilson, USA), a customized purge and trap system, a gas chromatographs (6850; Agilent Technologies, USA) equippled with Pola PLOT column, and an isotope ratio mass spectrometer (IRMS, MAT 253, Thermo Fisher Scientific, Germany) (64). High-purity N2O (Showa Denko Co., Ltd., purity >99.999%), which was calibrated previously with respect to the international isotopic standards air-N2 for δ15N and Vienna standard mean ocean water (VSMOW) for δ18O, were diluted with high-purity N2 and measured parallelly to correct and calibrate the ratios. Values of δ15Nbulk, δ15Nα, and δ18O were calculated as described previously (65). Values of δ15Nβ and SP were obtained as follows: δ15Nβ =2δ15Nbulk - δ15Nα; SP = δ15Nα - δ15Nβ. The typical analytical precisions are 0.2‰ for δ15Nα and 0.1‰ for δ15Nbulk and δ18O. Abiotic control N2O samples were obtained from bacteria-free NNFM supplemented with 2 mM NaNO2, acidified to pH <2 using 6M HCl.

The δ15NNO3- and δ18ONO3- values of the substrate NO3 in the medium were measured by the denitrifier method, where NO3 was converted into N2O by a denitrifying bacterium Pseudomonas aureofaciens ATCC 13985. The δ15NNO2- and δ18ONO2- values of the substrate NO2 in the medium were measured by chemical conversion of NO2 into N2O using the azide method. Detailed descriptions of these approaches and the procedures used for calibration of multiple standards (NO3, NO2, and N2O) for isotopomer ratios are indicated in Ref (66). The isotopologues of the substrate NO3 were determined as δ15NNO3- of 2.4 ± 0.1‰ and δ18ONO3- of 17.9 ± 0.3‰, while those of the substrate NO2 were δ15NNO2- of −1.8 ± 0.3‰ and δ18ONO2- of 3.7 ± 0.2‰. These isotopologues of the substrates were then used for correcting the absolute δ15Nbulk and δ18O values of N2O.

To clarify the mixed N2O production processes during NO3 reduction by strains Red69 and Red111, a quantitative estimation was performed using isotope mass balance. The fractional contributions (f value) to total N2O production based on SP were expressed as:

SP7day=SP3day×fDNRA+SPcNor×(1fDNRA)

where SP7-day and SP3-day represent the measured SP values of strains Red69 and Red111 in NO3 reduction at the culture time of 7 and 3 days, respectively. SPcNor denotes the SP values of denitrifying N2O produced from cNor (−5.9 ± 2.1‰)(67), and fDNRA indicates the relative contribution of DNRA-mediated process.

Phylogenomic analysis and comparative genomics

The reference genome sequences utilized in this study were retrieved from the NCBI database with accession numbers provided in Table S4. The phylogenomic tree was constructed using the up-to-date bacterial core gene set (UBCG) pipelines equipped with RAxML tool based on the amino acid sequences of 92 concatenated core genes (35). The presence and absence of functional genes related to the N metabolism in Geobacterales strains were determined through genome annotations by BlastKOALA server of the KEGG database (68) and RAST server (ClassicRAST annotation scheme) of the SEED database (69). In cases of inconsistencies between the two annotation servers, manual examination was performed using BLASTp against the KEGG GENES and NCBI-nr databases. For NrfA-based phylogenetic tree construction, amino acid sequences were retrieved from sequenced genomes available in the NCBI database, following the guidance in Ref (52, 70). MEGA X (71) and IQ-TREE (72) were used to align the sequences and generate the trees, respectively. AnnoTree v1.2, a tool used for exploration of functional genes across microbial tree of life, was used to search narG-hcp and nrfA-hcp gene combinations across the bacterial domain (Table S5)(73). All phylogenetic trees were polished by the interactive tree of life (iTOL) v5 (74).

Transcriptome analysis and real-time quantitative PCR

Geomonas sp. Red32 was inoculated into nine serum bottles (5% inoculum size, v/v) containing 20 mL NFFM_F and cultured for 2 days at 30°C. Then, NO3 and NO2 stock solutions were individually injected to three of the serum bottles to a final concentration of 10 and 2 mM, respectively. The remaining three serum bottles supplemented with the same amount of autoclaved water were the control group. After a further 4-h incubation, biomass was harvested by centrifugation at 10,000×g for 10 min at 4°C and then frozen in liquid nitrogen prior to storage in a freezer at −80°C. Total RNA was extracted using the total RNA extraction kit (Isogen II, Nippon Gene, Japan) and subsequently sent to Novogene Co. Ltd. (Beijing, China) for transcriptome sequencing on the Illumina NovaSeq 6000 platform. Differential gene expression analysis was performed using Bowtie2 (75) and DESeq2 (version 1.24.0) (76). Gene expression was quantified using the transcripts per million (TPM) method. Differentially regulated genes were defined with the criteria of a more than twofold change in expression and a false discovery rate-adjusted P < 0.05. Co-expressional networks were conducted by Cytoscape (version 3.8.2)(77) based on the Pearson correlation coefficient (Pearson’s r) calculated using R (version 4.0.4).

cDNA was synthesized using the ReverTra AceTM qPCR RT Master Mix with gDNA Remover (Toyobo, Japan). Real-time quantitative PCR (RT-qPCR) was performed using the StepOnePlus Real-Time PCR System (Applied Biosystems, USA) using a 20 µL reaction mixture containing 10 µL of PowerUp SYBR Green PCR Master Mix (2Χ, Thermo fisher), 1 µL of the primer (10 μΜ), 1 µL of the cDNA, and 8 µL of distilled water. The reaction condition was the default program of the instrument with the Fast-cycling model. The primer pairs for RT-qPCR were listed in Table S6. Fold changes of gene expression were calculated by comparative CT method. The reference gene rpoB was used to normalize the expression levels of target genes in this study (reference gene selection see Supplementary method). The consistency assessment of transcriptional and RT-qPCR results was described in the Supplementary method.

Statistical analyses

Statistical analyses and graphical visualizations were performed using the computing environment R v4.0.4 and GraphPad Prism v8.0.2. Significant differences (P- values) in N2O/NH4+ productions, bacterial biomass, and chlorate consumptions were assessed using the Student’s t-test for comparisons between two groups, or one-way ANOVA followed by LSD significant difference test for comparisons involving more than two groups. For RT-qPCR results and N2O isotopocule signatures, statistical significance was determined using the Mann–Whitney U tests for two-group comparisons or Kruskal–Wallis H tests for comparisons involving more than two groups.

ACKNOWLEDGMENTS

We thank Hideomi Itoh of National Institute of Advanced Industrial Sciences and Technology (Hokkaido, Japan) providing Geobacteraceae strains used in this study, we thank Shigeto Otsuka of The University of Tokyo (Tokyo, Japan) and Kazuo Isobe of Peking University (Peking, China) for helpful discussion.

This study was financially supported by National Natural Science Foundation of China (92351301), Japan Society for the Promotion of Science KAKENHI (JP21F21091, JP20H05679, JP20H00409, JP20K15423, JP18K19850, JP22H00383, and JP17H06105) and the New Energy and Industrial Technology Development Organization (NEDO) (JPNP18016). This study was also supported by Center for Ecological Research, Kyoto University, a Joint Usage/Research Center. Z.X. was supported by the JSPS Postdoctoral Fellowship (21P21091). S.H. was supported by MEXT/JSPS KAKENHI (20H04305), the Fundamental Research Funds for the Central Universities (International Collaboration Program (2024300346) and Cemac “GeoX” Interdisciplinary Program (2024300245), and start-up funding from Nanjing University.

Contributor Information

Zhenxing Xu, Email: xuzx.ut@gmail.com, xuzhenxing@sdu.edu.cn.

Yoko Masuda, Email: ygigico@gmail.com, yokomasuda@g.ecc.u-tokyo.ac.jp.

Derek R. Lovley, University of Massachusetts Amherst, Amherst, Massachusetts, USA

Daniel R Bond, University of Minnesota, St. Paul, Minnesota, USA.

DATA AVAILABILITY

The transcriptomic data generated in this study are publicly available in the NCBI under the BioProject accession number PRJNA801072.

SUPPLEMENTAL MATERIAL

The following material is available online at https://doi.org/10.1128/mbio.02540-24.

Supplemental Material. mbio.02540-24-s0001.pdf.

Supplemental methods; Figures S1 to S21.

DOI: 10.1128/mbio.02540-24.SuF1
Table S1. mbio.02540-24-s0002.xlsx.

Detailed isotopocule signatures of N2O produced from NO3- and NO2- reduction processes by Geobacterales strains and abiotic reactions.

mbio.02540-24-s0002.xlsx (23.4KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF2
Table S2. mbio.02540-24-s0003.xlsx.

Summary statistics of transcriptomic data and mapping results.

mbio.02540-24-s0003.xlsx (10.3KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF3
Table S3. mbio.02540-24-s0004.xlsx.

Detailed transcriptomic data focus on the nitrogen metabolism process of strain Red32 under NO3- and NO2- treatments.

mbio.02540-24-s0004.xlsx (22.9KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF4
Table S4. mbio.02540-24-s0005.xlsx.

Detailed information of species in the family Geobacteraceae.

mbio.02540-24-s0005.xlsx (16.6KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF5
Table S5. mbio.02540-24-s0006.xlsx.

Hit numbers and hit ratios of narG-hcp and nrfA-hcp gene combinations in the Bacteria domain.

mbio.02540-24-s0006.xlsx (41.2KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF6
Table S6. mbio.02540-24-s0007.xlsx.

Primers used in this study.

mbio.02540-24-s0007.xlsx (12.5KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF7

ASM does not own the copyrights to Supplemental Material that may be linked to, or accessed through, an article. The authors have granted ASM a non-exclusive, world-wide license to publish the Supplemental Material files. Please contact the corresponding author directly for reuse.

REFERENCES

  • 1. Myhre G, Shindell D, Bréon FM, Collins W, Fuglestvedt J, Huang J, Koch D, Lamarque JF, Lee D, Mendoza B, Nakajima T, Robock A, Stephens G, Takemura T, Zhang H. 2013. Anthropogenic and natural radiative forcing. In Stocker TF, Qin D, Plattner GK, Tignor M, Allen SK, Boschung J, Nauels A, Xia Y, Bex V, Midgley PM (ed), Climate change 2013: the physical science basis. contribution of working group I to the fifth assessment report of the intergovernmental panel on climate change. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. [Google Scholar]
  • 2. Ravishankara AR, Daniel JS, Portmann RW. 2009. Nitrous oxide (N2O): the dominant ozone-depleting substance emitted in the 21st century. Science 326:123–125. doi: 10.1126/science.1176985 [DOI] [PubMed] [Google Scholar]
  • 3. Tian H, Xu R, Canadell JG, Thompson RL, Winiwarter W, Suntharalingam P, Davidson EA, Ciais P, Jackson RB, Janssens-Maenhout G, et al. 2020. A comprehensive quantification of global nitrous oxide sources and sinks. Nature 586:248–256. doi: 10.1038/s41586-020-2780-0 [DOI] [PubMed] [Google Scholar]
  • 4. Thomson AJ, Giannopoulos G, Pretty J, Baggs EM, Richardson DJ. 2012. Biological sources and sinks of nitrous oxide and strategies to mitigate emissions. Philos Trans R Soc Lond B Biol Sci 367:1157–1168. doi: 10.1098/rstb.2011.0415 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Giles M, Morley N, Baggs EM, Daniell TJ. 2012. Soil nitrate reducing processes - drivers, mechanisms for spatial variation, and significance for nitrous oxide production. Front Microbiol 3:407. doi: 10.3389/fmicb.2012.00407 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Rütting T, Boeckx P, Müller C, Klemedtsson L. 2011. Assessment of the importance of dissimilatory nitrate reduction to ammonium for the terrestrial nitrogen cycle. Biogeosciences 8:1779–1791. doi: 10.5194/bg-8-1779-2011 [DOI] [Google Scholar]
  • 7. Pandey CB, Kumar U, Kaviraj M, Minick KJ, Mishra AK, Singh JS. 2020. DNRA: a short-circuit in biological N-cycling to conserve nitrogen in terrestrial ecosystems. Sci Total Environ 738:139710. doi: 10.1016/j.scitotenv.2020.139710 [DOI] [PubMed] [Google Scholar]
  • 8. Giblin AE, Tobias CR, Song B, Weston N, Banta GT, Rivera-Monroy VH. 2013. The importance of dissimilatory nitrate reduction to ammonium (DNRA) in the nitrogen cycle of coastal ecosystems. Oceanography 26:124–131. doi: 10.5670/oceanog.2013.54 [DOI] [Google Scholar]
  • 9. Yu L, Harris E, Lewicka‐Szczebak D, Barthel M, Blomberg MRA, Harris SJ, Johnson MS, Lehmann MF, Liisberg J, Müller C, Ostrom NE, Six J, Toyoda S, Yoshida N, Mohn J. 2020. What can we learn from N2O isotope data? – Analytics, processes and modelling. Rapid Comm Mass Spectrom 34:e8858. doi: 10.1002/rcm.8858 [DOI] [PubMed] [Google Scholar]
  • 10. Smith MS, Zimmerman K. 1981. Nitrous oxide production by nondenitrifying soil nitrate reducers. Soil Science Soc of Amer J 45:865–871. doi: 10.2136/sssaj1981.03615995004500050008x [DOI] [Google Scholar]
  • 11. Bleakley BH, Tiedje JM. 1982. Nitrous oxide production by organisms other than nitrifiers or denitrifiers. Appl Environ Microbiol 44:1342–1348. doi: 10.1128/aem.44.6.1342-1348.1982 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12. Torres MJ, Simon J, Rowley G, Bedmar EJ, Richardson DJ, Gates AJ, Delgado MJ. 2016. Nitrous oxide metabolism in nitrate-reducing bacteria: physiology and regulatory mechanisms. Adv Microb Physiol 68:353–432. doi: 10.1016/bs.ampbs.2016.02.007 [DOI] [PubMed] [Google Scholar]
  • 13. Stremińska MA, Felgate H, Rowley G, Richardson DJ, Baggs EM. 2012. Nitrous oxide production in soil isolates of nitrate-ammonifying bacteria. Environ Microbiol Rep 4:66–71. doi: 10.1111/j.1758-2229.2011.00302.x [DOI] [PubMed] [Google Scholar]
  • 14. Smith MS. 1982. Dissimilatory reduction of NO2- to NH4+ and N2O by Citrobacter sp. Appl Environ Microbiol 43:854–860. doi: 10.1128/aem.43.4.854-860.1982 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Smith MS. 1983. Nitrous oxide production by Escherichia coli is correlated with nitrate reductase activity. Appl Environ Microbiol 45:1545–1547. doi: 10.1128/aem.45.5.1545-1547.1983 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Gilberthorpe NJ, Poole RK. 2008. Nitric oxide homeostasis in Salmonella typhimurium: roles of respiratory nitrate reductase and flavohemoglobin. J Biol Chem 283:11146–11154. doi: 10.1074/jbc.M708019200 [DOI] [PubMed] [Google Scholar]
  • 17. Rowley G, Hensen D, Felgate H, Arkenberg A, Appia-Ayme C, Prior K, Harrington C, Field SJ, Butt JN, Baggs E, Richardson DJ. 2012. Resolving the contributions of the membrane-bound and periplasmic nitrate reductase systems to nitric oxide and nitrous oxide production in Salmonella enterica serovar Typhimurium . Biochem J 441:755–762. doi: 10.1042/BJ20110971 [DOI] [PubMed] [Google Scholar]
  • 18. Mills PC, Rowley G, Spiro S, Hinton JCD, Richardson DJ. 2008. A combination of cytochrome c nitrite reductase (NrfA) and flavorubredoxin (NorV) protects Salmonella enterica serovar Typhimurium against killing by NO in anoxic environments. Microbiology 154:1218–1228. doi: 10.1099/mic.0.2007/014290-0 [DOI] [PubMed] [Google Scholar]
  • 19. Vine CE, Cole JA. 2011. Nitrosative stress in Escherichia coli: reduction of nitric oxide. Biochem Soc Trans 39:213–215. doi: 10.1042/BST0390213 [DOI] [PubMed] [Google Scholar]
  • 20. Heo H, Kwon M, Song B, Yoon S. 2020. Involvement of NO3- in Involvement of NO3- in ecophysiological regulation of dissimilatory nitrate/nitrite reduction to ammonium (DNRA) is implied by physiological characterization of soil DNRA bacteria isolated via a colorimetric screening method. Appl Environ Microbiol 86:e01054-20. doi: 10.1128/AEM.01054-20 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Luckmann M, Mania D, Kern M, Bakken LR, Frostegård Å, Simon J. 2014. Production and consumption of nitrous oxide in nitrate-ammonifying Wolinella succinogenes cells. Microbiology 160:1749–1759. doi: 10.1099/mic.0.079293-0 [DOI] [PubMed] [Google Scholar]
  • 22. Yoshida N, Toyoda S. 2000. Constraining the atmospheric N2O budget from intramolecular site preference in N2O isotopomers. Nature 405:330–334. doi: 10.1038/35012558 [DOI] [PubMed] [Google Scholar]
  • 23. Toyoda S, Yoshida N, Koba K. 2017. Isotopocule analysis of biologically produced nitrous oxide in various environments. Mass Spectrom Rev 36:135–160. doi: 10.1002/mas.21459 [DOI] [PubMed] [Google Scholar]
  • 24. Saghaï A, Pold G, Jones CM, Hallin S. 2023. Phyloecology of nitrate ammonifiers and their importance relative to denitrifiers in global terrestrial biomes. Nat Commun 14:8249. doi: 10.1038/s41467-023-44022-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Bahram M, Espenberg M, Pärn J, Lehtovirta-Morley L, Anslan S, Kasak K, Kõljalg U, Liira J, Maddison M, Moora M, Niinemets Ü, Öpik M, Pärtel M, Soosaar K, Zobel M, Hildebrand F, Tedersoo L, Mander Ü. 2022. Structure and function of the soil microbiome underlying N2O emissions from global wetlands. Nat Commun 13:1430. doi: 10.1038/s41467-022-29161-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Qian H, Zhu X, Huang S, Linquist B, Kuzyakov Y, Wassmann R, Minamikawa K, Martinez-Eixarch M, Yan X, Zhou F, Sander BO, Zhang W, Shang Z, Zou J, Zheng X, Li G, Liu Z, Wang S, Ding Y, van Groenigen KJ, Jiang Y. 2023. Greenhouse gas emissions and mitigation in rice agriculture. Nat Rev Earth Environ 4:716–732. doi: 10.1038/s43017-023-00482-1 [DOI] [Google Scholar]
  • 27. Nojiri Y, Kaneko Y, Azegami Y, Shiratori Y, Ohte N, Senoo K, Otsuka S, Isobe K. 2020. Dissimilatory nitrate reduction to ammonium and responsible microbes in Japanese rice paddy soil. Microbes Environ 35:1–7. doi: 10.1264/jsme2.ME20069 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Lovley DR, Ueki T, Zhang T, Malvankar NS, Shrestha PM, Flanagan KA, Aklujkar M, Butler JE, Giloteaux L, Rotaru AE, Holmes DE, Franks AE, Orellana R, Risso C, Nevin KP. 2011. Geobacter: the microbe electric’s physiology, ecology, and practical applications. Adv Microb Physiol 59:1–100. doi: 10.1016/B978-0-12-387661-4.00004-5 [DOI] [PubMed] [Google Scholar]
  • 29. Masuda Y, Itoh H, Shiratori Y, Isobe K, Otsuka S, Senoo K. 2017. Predominant but previously-overlooked prokaryotic drivers of reductive nitrogen transformation in paddy soils, revealed by metatranscriptomics. Microbes Environ 32:180–183. doi: 10.1264/jsme2.ME16179 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Xu Z, Masuda Y, Wang X, Ushijima N, Shiratori Y, Senoo K, Itoh H. 2021. Genome-based taxonomic rearrangement of the order Geobacterales including the description of Geomonas azotofigens sp. nov. and Geomonas diazotrophica sp. nov. Front Microbiol 12:737531. doi: 10.3389/fmicb.2021.737531 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Xu Z, Masuda Y, Hayakawa C, Ushijima N, Kawano K, Shiratori Y, Senoo K, Itoh H. 2020. Description of three novel members in the family Geobacteraceae, Oryzomonas japonicum gen. nov., sp. nov., Oryzomonas sagensis sp. nov., and Oryzomonas ruber sp. nov. Microorganisms 8:634. doi: 10.3390/microorganisms8050634 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Itoh H, Xu Z, Masuda Y, Ushijima N, Hayakawa C, Shiratori Y, Senoo K. 2021. Geomonas silvestris sp. nov., Geomonas paludis sp. nov. and Geomonas limicola sp. nov., isolated from terrestrial environments, and emended description of the genus Geomonas. Int J Syst Evol Microbiol 71. doi: 10.1099/ijsem.0.004607 [DOI] [PubMed] [Google Scholar]
  • 33. Zhang Z, Xu Z, Masuda Y, Wang X, Ushijima N, Shiratori Y, Senoo K, Itoh H. 2021. Geomesophilobacter sediminis gen. nov., sp. nov., Geomonas propionica sp. nov. and Geomonas anaerohicana sp. nov., three novel members in the family Geobacterecace isolated from river sediment and paddy soil. Syst Appl Microbiol 44:126233. doi: 10.1016/j.syapm.2021.126233 [DOI] [PubMed] [Google Scholar]
  • 34. Xu Z, Masuda Y, Itoh H, Ushijima N, Shiratori Y, Senoo K. 2019. Geomonas oryzae gen. nov., sp. nov., Geomonas edaphica sp. nov., Geomonas ferrireducens sp. nov., Geomonas terrae sp. nov., four ferric-reducing bacteria isolated from paddy soil , and reclassification of three species of the genus Geobacter as members of the genus Geomonas gen. nov. Front Microbiol 10:1–17. doi: 10.3389/fmicb.2019.02201 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35. Na SI, Kim YO, Yoon SH, Ha SM, Baek I, Chun J. 2018. UBCG: up-to-date bacterial core gene set and pipeline for phylogenomic tree reconstruction. J Microbiol 56:280–285. doi: 10.1007/s12275-018-8014-6 [DOI] [PubMed] [Google Scholar]
  • 36. Satoh T, Hom SSM, Shanmugam KT. 1983. Production of nitrous oxide from nitrite in Klebsiella pneumoniae: mutants altered in nitrogen metabolism. J Bacteriol 155:454–458. doi: 10.1128/jb.155.2.454-458.1983 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Cruz-García C, Murray AE, Klappenbach JA, Stewart V, Tiedje JM. 2007. Respiratory nitrate ammonification by Shewanella oneidensis MR-1. J Bacteriol 189:656–662. doi: 10.1128/JB.01194-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Hu HW, Chen D, He JZ. 2015. Microbial regulation of terrestrial nitrous oxide formation: understanding the biological pathways for prediction of emission rates. FEMS Microbiol Rev 39:729–749. doi: 10.1093/femsre/fuv021 [DOI] [PubMed] [Google Scholar]
  • 39. Lipschultz F, Zafiriou OC, Wofsy SC, McElroy MB, Valois FW, Watson SW. 1981. Production of NO and N2O by soil nitrifying bacteria. Nature 294:641–643. doi: 10.1038/294641a0 [DOI] [Google Scholar]
  • 40. Stieglmeier M, Mooshammer M, Kitzler B, Wanek W, Zechmeister-Boltenstern S, Richter A, Schleper C. 2014. Aerobic nitrous oxide production through N-nitrosating hybrid formation in ammonia-oxidizing archaea. ISME J 8:1135–1146. doi: 10.1038/ismej.2013.220 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Kits KD, Jung M-Y, Vierheilig J, Pjevac P, Sedlacek CJ, Liu S, Herbold C, Stein LY, Richter A, Wissel H, Brüggemann N, Wagner M, Daims H. 2019. Low yield and abiotic origin of N2O formed by the complete nitrifier Nitrospira inopinata. Nat Commun 10:1836. doi: 10.1038/s41467-019-09790-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Okabe S, Oshiki M, Takahashi Y, Satoh H. 2011. N2O emission from a partial nitrification–anammox process and identification of a key biological process of N2O emission from anammox granules. Water Res 45:6461–6470. doi: 10.1016/j.watres.2011.09.040 [DOI] [PubMed] [Google Scholar]
  • 43. Planchet E, Sonoda M, Zeier J, Kaiser WM. 2006. Nitric oxide (NO) as an intermediate in the cryptogein-induced hypersensitive response--a critical re-evaluation. Plant Cell Environ 29:59–69. doi: 10.1111/j.1365-3040.2005.01400.x [DOI] [PubMed] [Google Scholar]
  • 44. Jung MY, Well R, Min D, Giesemann A, Park SJ, Kim JG, Kim SJ, Rhee SK. 2014. Isotopic signatures of N2O produced by ammonia-oxidizing archaea from soils. ISME J 8:1115–1125. doi: 10.1038/ismej.2013.205 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Well R, Braker G, Giesemann A, Flessa H. 2010. Are isotopologue signatures of N2O from bacterial denitrifiers indicative of NOR type? AGU Fall Meet Abstr B51B-0347. [Google Scholar]
  • 46. Hagen WR. 2022. Structure and function of the hybrid cluster protein. Coord Chem Rev 457:214405. doi: 10.1016/j.ccr.2021.214405 [DOI] [Google Scholar]
  • 47. Yu Z, Pesesky M, Zhang L, Huang J, Winkler M, Chistoserdova L. 2020. A complex interplay between nitric oxide, quorum sensing, and the unique secondary metabolite tundrenone constitutes the hypoxia response in Methylobacter. mSystems 5:e00770-19. doi: 10.1128/mSystems.00770-19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Wang J, Vine CE, Balasiny BK, Rizk J, Bradley CL, Tinajero-Trejo M, Poole RK, Bergaust LL, Bakken LR, Cole JA. 2016. The roles of the hybrid cluster protein, Hcp and its reductase, Hcr, in high affinity nitric oxide reduction that protects anaerobic cultures of Escherichia coli against nitrosative stress. Mol Microbiol 100:877–892. doi: 10.1111/mmi.13356 [DOI] [PubMed] [Google Scholar]
  • 49. Rusmana I, Nedwell DB. 2004. Use of chlorate as a selective inhibitor to distinguish membrane-bound nitrate reductase (Nar) and periplasmic nitrate reductase (Nap) of dissimilative nitrate reducing bacteria in sediment. FEMS Microbiol Ecol 48:379–386. doi: 10.1016/j.femsec.2004.02.010 [DOI] [PubMed] [Google Scholar]
  • 50. Bertero MG, Rothery RA, Palak M, Hou C, Lim D, Blasco F, Weiner JH, Strynadka NCJ. 2003. Insights into the respiratory electron transfer pathway from the structure of nitrate reductase A. Nat Struct Mol Biol 10:681–687. doi: 10.1038/nsb969 [DOI] [PubMed] [Google Scholar]
  • 51. Einsle O, Messerschmidt A, Stach P, Bourenkov GP, Bartunik HD, Huber R, Kroneck PMH. 1999. Structure of cytochrome c nitrite reductase. Nature 400:476–480. doi: 10.1038/22802 [DOI] [PubMed] [Google Scholar]
  • 52. Welsh A, Chee-Sanford JC, Connor LM, Löffler FE, Sanford RA. 2014. Refined NrfA phylogeny improves PCR-based nrfA gene detection. Appl Environ Microbiol 80:2110–2119. doi: 10.1128/AEM.03443-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53. Clegg S, Yu F, Griffiths L, Cole JA. 2002. The roles of the polytopic membrane proteins NarK, NarU and NirC in Escherichia coli K-12: two nitrate and three nitrite transporters. Mol Microbiol 44:143–155. doi: 10.1046/j.1365-2958.2002.02858.x [DOI] [PubMed] [Google Scholar]
  • 54. Ostrom NE, Pitt A, Sutka R, Ostrom PH, Grandy AS, Huizinga KM, Robertson GP. 2007. Isotopologue effects during N2O reduction in soils and in pure cultures of denitrifiers . J Geophys Res 112:G02005. doi: 10.1029/2006JG000287 [DOI] [Google Scholar]
  • 55. Harris E, Joss A, Emmenegger L, Kipf M, Wolf B, Mohn J, Wunderlin P. 2015. Isotopic evidence for nitrous oxide production pathways in a partial nitritation-anammox reactor. Water Res 83:258–270. doi: 10.1016/j.watres.2015.06.040 [DOI] [PubMed] [Google Scholar]
  • 56. Su X, Wen T, Wang Y, Xu J, Cui L, Zhang J, Xue X, Ding K, Tang Y, Zhu Y-G. 2021. Stimulation of N2O emission via bacterial denitrification driven by acidification in estuarine sediments. Glob Chang Biol 27:5564–5579. doi: 10.1111/gcb.15863 [DOI] [PubMed] [Google Scholar]
  • 57. Wankel SD, Ziebis W, Buchwald C, Charoenpong C, de Beer D, Dentinger J, Xu Z, Zengler K. 2017. Evidence for fungal and chemodenitrification based N2O flux from nitrogen impacted coastal sediments. Nat Commun 8:15595. doi: 10.1038/ncomms15595 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58. Sutka RL, Adams GC, Ostrom NE, Ostrom PH. 2008. Isotopologue fractionation during N2O production by fungal denitrification. Rapid Commun Mass Spectrom 22:3989–3996. doi: 10.1002/rcm.3820 [DOI] [PubMed] [Google Scholar]
  • 59. He Y, Cao X, Bao H. 2020. Ideas and perspectives: the same carbon behaves like different elements – an insight into position-specific isotope distributions. Biogeosciences 17:4785–4795. doi: 10.5194/bg-17-4785-2020 [DOI] [Google Scholar]
  • 60. Jiang Y, Chen B, Duan C, Sun B, Yang J, Yang S. 2015. Multigene editing in the Escherichia coli genome via the CRISPR-Cas9 system. Appl Environ Microbiol 81:2506–2514. doi: 10.1128/AEM.04023-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Kempers AJ. 1974. Determination of sub-microquantities of ammonium and nitrates in soils with phenol, sodiumnitroprusside and hypochlorite. Geoderma 12:201–206. doi: 10.1016/0016-7061(74)90068-8 [DOI] [Google Scholar]
  • 62. Bunsen R. 1855. XV. On the law of absorption of gases. Lond Edinb Dubl Phil Mag J Sci 9:116–130. doi: 10.1080/14786445508641836 [DOI] [Google Scholar]
  • 63. Liu S, Han P, Hink L, Prosser JI, Wagner M, Brüggemann N. 2017. Abiotic conversion of extracellular NH2OH contributes to N2O emission during ammonia oxidation. Environ Sci Technol 51:13122–13132. doi: 10.1021/acs.est.7b02360 [DOI] [PubMed] [Google Scholar]
  • 64. Hattori S, Savarino J, Kamezaki K, Ishino S, Dyckmans J, Fujinawa T, Caillon N, Barbero A, Mukotaka A, Toyoda S, Well R, Yoshida N. 2016. Automated system measuring triple oxygen and nitrogen isotope ratios in nitrate using the bacterial method and N2O decomposition by microwave discharge. Rapid Commun Mass Spectrom 30:2635–2644. doi: 10.1002/rcm.7747 [DOI] [PubMed] [Google Scholar]
  • 65. Toyoda S, Yoshida N. 1999. Determination of nitrogen isotopomers of nitrous oxide on a modified isotope ratio mass spectrometer. Anal Chem 71:4711–4718. doi: 10.1021/ac9904563 [DOI] [Google Scholar]
  • 66. Kobayashi K, Fukushima K, Onishi Y, Nishina K, Makabe A, Yano M, Wankel SD, Koba K, Okabe S. 2021. Influence of δ18O of water on measurements of δ18O of nitrite and nitrate. Rapid Commun Mass Spectrom 35:e8979. doi: 10.1002/rcm.8979 [DOI] [PubMed] [Google Scholar]
  • 67. Yamazaki T, Hozuki T, Arai K, Toyoda S, Koba K, Fujiwara T, Yoshida N. 2014. Isotopomeric characterization of nitrous oxide produced by reaction of enzymes extracted from nitrifying and denitrifying bacteria. Biogeosciences 11:2679–2689. doi: 10.5194/bg-11-2679-2014 [DOI] [Google Scholar]
  • 68. Kanehisa M, Sato Y, Morishima K. 2016. BlastKOALA and GhostKOALA: KEGG tools for functional characterization of genome and metagenome sequences. J Mol Biol 428:726–731. doi: 10.1016/j.jmb.2015.11.006 [DOI] [PubMed] [Google Scholar]
  • 69. Aziz RK, Bartels D, Best AA, DeJongh M, Disz T, Edwards RA, Formsma K, Gerdes S, Glass EM, Kubal M, et al. 2008. The RAST Server: rapid annotations using subsystems technology. BMC Genomics 9:75. doi: 10.1186/1471-2164-9-75 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70. Campeciño J, Lagishetty S, Wawrzak Z, Sosa Alfaro V, Lehnert N, Reguera G, Hu J, Hegg EL. 2020. Cytochrome c nitrite reductase from the bacterium Geobacter lovleyi represents a new NrfA subclass. J Biol Chem 295:11455–11465. doi: 10.1074/jbc.RA120.013981 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. 2018. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol 35:1547–1549. doi: 10.1093/molbev/msy096 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A, Lanfear R. 2020. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Mol Biol Evol 37:1530–1534. doi: 10.1093/molbev/msaa015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73. Mendler K, Chen H, Parks DH, Lobb B, Hug LA, Doxey AC. 2019. AnnoTree: visualization and exploration of a functionally annotated microbial tree of life. Nucleic Acids Res 47:4442–4448. doi: 10.1093/nar/gkz246 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74. Letunic I, Bork P. 2021. Interactive Tree Of Life (iTOL) v5: an online tool for phylogenetic tree display and annotation. Nucleic Acids Res 49:W293–W296. doi: 10.1093/nar/gkab301 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75. Langmead B, Salzberg SL. 2012. Fast gapped-read alignment with Bowtie 2. Nat Methods 9:357–359. doi: 10.1038/nmeth.1923 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76. Love MI, Huber W, Anders S. 2014. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol 15:550. doi: 10.1186/s13059-014-0550-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Shannon P, Markiel A, Ozier O, Baliga NS, Wang JT, Ramage D, Amin N, Schwikowski B, Ideker T. 2003. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res 13:2498–2504. doi: 10.1101/gr.1239303 [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental Material. mbio.02540-24-s0001.pdf.

Supplemental methods; Figures S1 to S21.

DOI: 10.1128/mbio.02540-24.SuF1
Table S1. mbio.02540-24-s0002.xlsx.

Detailed isotopocule signatures of N2O produced from NO3- and NO2- reduction processes by Geobacterales strains and abiotic reactions.

mbio.02540-24-s0002.xlsx (23.4KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF2
Table S2. mbio.02540-24-s0003.xlsx.

Summary statistics of transcriptomic data and mapping results.

mbio.02540-24-s0003.xlsx (10.3KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF3
Table S3. mbio.02540-24-s0004.xlsx.

Detailed transcriptomic data focus on the nitrogen metabolism process of strain Red32 under NO3- and NO2- treatments.

mbio.02540-24-s0004.xlsx (22.9KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF4
Table S4. mbio.02540-24-s0005.xlsx.

Detailed information of species in the family Geobacteraceae.

mbio.02540-24-s0005.xlsx (16.6KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF5
Table S5. mbio.02540-24-s0006.xlsx.

Hit numbers and hit ratios of narG-hcp and nrfA-hcp gene combinations in the Bacteria domain.

mbio.02540-24-s0006.xlsx (41.2KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF6
Table S6. mbio.02540-24-s0007.xlsx.

Primers used in this study.

mbio.02540-24-s0007.xlsx (12.5KB, xlsx)
DOI: 10.1128/mbio.02540-24.SuF7

Data Availability Statement

The transcriptomic data generated in this study are publicly available in the NCBI under the BioProject accession number PRJNA801072.


Articles from mBio are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES