Skip to main content
Applied and Environmental Microbiology logoLink to Applied and Environmental Microbiology
. 2018 Jan 31;84(4):e01985-17. doi: 10.1128/AEM.01985-17

Denitrification by Anaeromyxobacter dehalogenans, a Common Soil Bacterium Lacking the Nitrite Reductase Genes nirS and nirK

Jenny R Onley a, Samiha Ahsan b, Robert A Sanford c, Frank E Löffler a,d,e,f,g,
Editor: Harold L Drakeh
PMCID: PMC5795083  PMID: 29196287

ABSTRACT

The versatile soil bacterium Anaeromyxobacter dehalogenans lacks the hallmark denitrification genes nirS and nirK (encoding NO2→NO reductases) and couples growth to NO3 reduction to NH4+ (respiratory ammonification) and to N2O reduction to N2. A. dehalogenans also grows by reducing Fe(III) to Fe(II), which chemically reacts with NO2 to form N2O (i.e., chemodenitrification). Following the addition of 100 μmol of NO3 or NO2 to Fe(III)-grown axenic cultures of A. dehalogenans, 54 (±7) μmol and 113 (±2) μmol N2O-N, respectively, were produced and subsequently consumed. The conversion of NO3 to N2 in the presence of Fe(II) through linked biotic-abiotic reactions represents an unrecognized ecophysiology of A. dehalogenans. The new findings demonstrate that the assessment of gene content alone is insufficient to predict microbial denitrification potential and N loss (i.e., the formation of gaseous N products). A survey of complete bacterial genomes in the NCBI Reference Sequence database coupled with available physiological information revealed that organisms lacking nirS or nirK but with Fe(III) reduction potential and genes for NO3 and N2O reduction are not rare, indicating that NO3 reduction to N2 through linked biotic-abiotic reactions is not limited to A. dehalogenans. Considering the ubiquity of iron in soils and sediments and the broad distribution of dissimilatory Fe(III) and NO3 reducers, denitrification independent of NO-forming NO2 reductases (through combined biotic-abiotic reactions) may have substantial contributions to N loss and N2O flux.

IMPORTANCE Current attempts to gauge N loss from soils rely on the quantitative measurement of nirK and nirS genes and/or transcripts. In the presence of iron, the common soil bacterium Anaeromyxobacter dehalogenans is capable of denitrification and the production of N2 without the key denitrification genes nirK and nirS. Such chemodenitrifiers denitrify through combined biotic and abiotic reactions and have potentially large contributions to N loss to the atmosphere and fill a heretofore unrecognized ecological niche in soil ecosystems. The findings emphasize that the comprehensive understanding of N flux and the accurate assessment of denitrification potential can be achieved only when integrated studies of interlinked biogeochemical cycles are performed.

KEYWORDS: Anaeromyxobacter, chemodenitrification, coupled Fe- and N-redox processes, denitrification, nitrogen loss

INTRODUCTION

Fixed nitrogen (N) availability limits primary production in most natural and managed soil ecosystems, and to meet the demands of a growing human population for food, feed, and biofuel crops, global fertilizer usage continues to increase. The Haber-Bosch process (the chemical conversion of N2 to ammonia [NH3]) introduces about 100 Tg of fixed N into the environment each year worldwide as fertilizer (1). As a consequence, the mass of fixed N has doubled within the last century and is at levels never seen over the history of life on Earth. A substantial amount of the fixed N is lost from the soil due to microbial nitrification (ammonium oxidation to nitrate; NH4+→NO3) and subsequent microbial denitrification (nitrate reduction to nitrous oxide/nitrogen gas; NO3→N2O/N2). Both nitrification and denitrification generate N2O (2), a potent greenhouse gas with ozone depletion potential (3). To assess the role of microorganisms in N loss from soils to the atmosphere (i.e., emissions of the gaseous products N2O and N2), the measurement of the denitrification genes nirS and nirK (encoding nitrite→nitric oxide [NO2→NO] reductases) and nosZ (encoding N2O→N2 reductase) has been applied. Several studies demonstrated a positive correlation between the abundance of these denitrification genes (i.e., nirK, nirS, and nosZ) and NO3/NO2 conversion to gaseous products (for examples, see references 47); however, other studies failed to establish this relationship (811). Subsequent efforts showed that the commonly used nirK and nirS quantitative PCR (qPCR) primers yield poor coverage of nirK and nirS sequences recovered from metagenomes and underestimate the true nirK and nirS gene abundance in environmental samples (12, 13). Similar issues apply to nosZ, and recent efforts discovered novel clade II nosZ genes, which are not amplified by previously designed primers targeting the clade I nosZ sequences (14, 15). The analysis of metagenomic data sets revealed that the previously unrecognized clade II nosZ genes are more prevalent in many soil ecosystems, suggesting that their hosts are potentially relevant contributors to N2O consumption (1417). Additionally, a multitude of environmental parameters, such as soil moisture (i.e., water-filled pore space), soil type, total N, redox potential (Eh), and reactive iron content, may also influence N2O emissions (9, 18, 19). Chemodenitrification, a process in which NO2 is converted to N2O in a chemical (abiotic) redox reaction with ferrous iron [Fe(II)] (20), has not been attributed a major role in soil N loss. A few studies have shown that NO3-reducing microorganisms couple chemodenitrification (NO2→N2O) and biotic reactions (NO3→NO2) to reduce NO3/NO2 and oxidize Fe(II) (2123). Recent efforts have demonstrated rapid chemodenitrification under a range of NO2 and Fe(II) concentrations (2426) and provided evidence for in situ chemodenitrification in soils and sediments (26, 27). These observations indicate that monitoring the nirS and nirK genes is insufficient to gauge denitrification potential, and abiotic factors should be considered for predicting N loss.

The soil bacterium Anaeromyxobacter dehalogenans strain 2CP-C reduces ferric iron [Fe(III)] to Fe(II) (28) and reduces NO3 to NH4+ with NO2 as the intermediate via the respiratory ammonification pathway (15, 29). A. dehalogenans possesses the narG and napA genes (encoding NO3 reductases) and the nrfA gene (encoding NO2→NH4+ reductase) and the nosZ gene (encoding clade II N2O→N2 reductase), but the organism is not classified as a denitrifier due to the observed NO3 to NH4+ reduction activity and the absence of nirK and nirS. Recent studies demonstrated that nosZ genes belonging to members of the Anaeromyxobacter genus compose a large proportion of the total nosZ genes in Illinois soil metagenomes (16) and that Anaeromyxobacter transcripts compose a sizeable proportion of nar, nor, nos, and nrf transcripts in rice paddy soils (30). Since little is known how organisms with genes encoding NO3 and N2O reductases but lacking nirK and nirS might bridge the Fe and N cycles, the effect of Fe(II) on NO3 and NO2 reduction was explored in axenic A. dehalogenans cultures. These experiments demonstrated that A. dehalogenans facilitates chemodenitrification through Fe(III) and NO3 reduction and ultimately denitrifies to N2 through coupled biotic-abiotic reactions.

RESULTS

A. dehalogenans reduces NO3 to N2 and NH4+ in the presence of Fe(II).

In the absence of Fe(III) as the electron acceptor, stoichiometric conversion of NO3 to NH4+ was observed in pure cultures of A. dehalogenans, consistent with the organism's classification as a respiratory ammonifier (29). In contrast, the addition of 100 μmol of NO3 to A. dehalogenans strain 2CP-C cultures that had reduced 480 (±20) μmol of Fe(III) [as Fe(III) citrate] resulted in the rapid formation of 54 (±7) μmol N2O-N and 63 (±5) μmol NH4+ (Fig. 1A). The N2O was subsequently consumed (Fig. 1A) but was stable in replicate cultures containing acetylene, which inhibits N2O reduction (Fig. 1B). qPCR enumeration of 16S rRNA genes revealed cell increases indicative of growth during the Fe(III) and NO3/NO2 reduction phases (Table 1; see Fig. S1 in the supplemental material). Although A. dehalogenans couples N2O reduction with energy conservation and growth (15, 17), qPCR could not reveal additional cell increases (expected yield of 1.3 × 107 cells/ml from 54 μmol N2O-N reduction after day 12) due to prior growth with Fe(III) and NO3 as electron acceptors (resulting in 6.0 × 107 cells/ml) and interference of Fe(III) oxide precipitate in DNA extraction. A bright orange precipitate, presumably Fe(III) oxides, formed in the DNA extraction tubes, suggesting interference and reduced DNA extraction efficiency following Fe(III) reduction activity (31). In the cultures lacking Fe(II) [i.e., containing only the 750 nmol of Fe(II) introduced with the trace metal solution], A. dehalogenans reduced 200 (±10) μmol NO3 to 190 (±12) μmol NH4+ (Fig. S2), which is consistent with the gene content of A. dehalogenans and the organism's characterization as a respiratory ammonifier (15, 29). The amount of N2O-N did not exceed 2 μmol in any of the incubation vessels not amended with Fe(III) as the electron acceptor. Thus, the major product of NO3 reduction was NH4+ in the absence of Fe(II) and N2 in the presence of Fe(II).

FIG 1.

FIG 1

Abiotic production of N2O after NO3 addition to A. dehalogenans cultures grown with Fe(III). (A) A. dehalogenans grown in defined mineral salts medium with 1,000 μmol of acetate and 700 μmol of Fe(III) citrate. NO3 (100 μmol) was added on day 10 (indicated by an arrow in each panel). (B) Same growth conditions as those described for panel A except that acetylene (10% of headspace) was included to prevent NosZ activity. The error bars represent the standard deviation of results of three replicate cultures. Symbols: black squares, Fe(II); open circles, N2O-N; open triangles, NH4+. No error bars are shown if the standard deviation is smaller than the symbol.

TABLE 1.

qPCR data showing cell growth associated with Fe(III), NO3, and N2O reduction, corresponding to Fig. 1 and 2a

Corresponding figure A. dehalogenans (cells ml−1 [SD]b)
Initialc After Fe(III) reduction phased After NO3/NO2 reduction phasee After N2O reduction phasef
1A 4.8 × 103 (1.5 × 103) 5.6 × 106 (2.0 × 106) 6.0 × 107 (2.5 × 107) 3.3 × 107 (1.4 × 107)
1B 7.8 × 103 (3.2 × 103) 8.4 × 106 (3.9 × 106) 4.4 × 107 (6.5 × 106) NAg
2A 2.2 × 104 (1.4 × 104) 4.3 × 106 (2.0 × 106) 2.1 × 106 (6.1 ×105) 1.4 × 107 (6.5 × 106)
2B 2.1 × 104 (1.4 × 104) 1.2 × 107 (9.4 × 106) 5.3 × 102 (6.6 × 102) NA
a

See Fig. S1 and S3 in the supplemental material for more detailed qPCR data.

b

Standard deviations of triplicate cultures are indicated in parentheses.

c

Cells introduced with the inoculum.

d

Day 10 for Fig. 1A and B; day 13 for Fig. 2A and B.

e

Day 12 for Fig. 1A and B; day 18 for Fig. 2A and B.

f

Day 21 for Fig. 1A; day 41 for Fig. 1B.

g

NA, not applicable.

A. dehalogenans reduces NO2 to N2 in the presence of Fe(II).

Similar to NO3 amendment, the addition of 100 μmol of NO2 to A. dehalogenans strain 2CP-C cultures that had reduced 590 (±60) μmol of Fe(III) resulted in rapid N2O formation (Fig. 2A). N2O formation also occurred in bottles that were heat treated prior to NO2 addition, indicating that the reduction of NO2 to N2O was an abiotic process (Fig. 2B). In live and killed control incubations, 113 (±2) and 109 (±9) μmol N2O-N were produced, respectively, indicating stoichiometric conversion of NO2 to N2O.

FIG 2.

FIG 2

Abiotic production of N2O after NO2 addition to A. dehalogenans cultures grown with Fe(III). (A) A. dehalogenans grown in defined mineral salts medium with 100 μmol of acetate and 900 μmol of Fe(III) citrate. On day 14, 100 μmol of NO2 was added (indicated by the left downward arrow in each panel). On day 18, an additional 200 μmol of acetate was added (indicated by the right downward arrow in each panel). (B) Cultures treated the same as described for panel A except that they were autoclaved prior to NO2 addition. Error bars represent the standard deviation of results of three replicate cultures. Symbols: black squares, Fe(II); open circles, N2O-N. No error bars are shown if the standard deviation is smaller than the symbol.

Live cultures had depleted the initial amount of 100 μmol acetate during the initial Fe(III) reduction phase, and N2O was not consumed. Following the addition of 200 μmol of acetate on day 18, Fe(III) reduction resumed and N2O consumption started (Fig. 2A). During the initial Fe(III) reduction phase, 4.3 × 106 cells/ml were produced, and an increase by approximately 1.2 × 107 cells/ml was measured during the consumption of N2O as the electron acceptor (Fig. S3). The cell yield is slightly lower than the expected value of 2.7 × 107 cells/ml (for 113 μmol N2O-N) (Table 1), which is likely due to interferences of iron precipitates with DNA extraction and subsequent qPCR enumeration (see above). In controls that were heat killed following the initial Fe(III) reduction phase, the biotically produced Fe(II) reacted with NO2, leading to the formation of stoichiometric amounts of N2O, but N2O was stable even after acetate addition, indicating that live cells were required for N2O reduction to occur (Fig. 2B). Live and heat-killed cultures produced N2O at similar rates of 1.9 (±0.03) and 1.9 (±0.13) μmol h−1, respectively (Fig. 2A and B), whereas lower N2O formation rates of 1.3 (±0.04) μmol h−1 were measured in abiotic (i.e., no cells) control incubations with 800 μmol of ferrous chloride and 100 μmol of NO2 (Fig. S4). These findings suggest that equimolar concentrations of biotically produced Fe(II) were more reactive than Fe(II) added as ferrous chloride.

N2O formation in cultures with insoluble Fe(III) and effect of sulfide on N2O reduction.

NO2 reduction was also monitored in A. dehalogenans cultures that had reduced 250 (±7) μmol poorly crystalline Fe(III) oxyhydroxide (Fig. 3) or 960 (±50) μmol soluble Fe(III) (as ferric citrate; data not shown) in medium containing 0.2 mM sulfide as a reducing agent. With insoluble and soluble forms of Fe(III), 35 (±4) and 47 (±8) μmol N2O-N were produced, respectively, during the reduction of 100 μmol NO2, but N2O was not subsequently consumed by A. dehalogenans (Fig. 3). Cultures of A. dehalogenans consistently used N2O as a respiratory electron acceptor in defined, anoxic medium without sulfide (Fig. S5), and further tests were performed to determine if the presence of sulfide could explain why N2O was not consumed. In medium amended with 0.2 mM sulfide, A. dehalogenans cultures did not reduce N2O (Fig. S5), despite the ability of A. dehalogenans to grow with other electron acceptors [e.g., NO3 and Fe(III)] in the presence of 0.2 mM sulfide.

FIG 3.

FIG 3

Cultures of A. dehalogenans grown with 1,000 μmol of acetate and Fe(III) oxyhydroxide (500 acid-extractable μmol). NO2 (100 μmol) was added on day 9 (indicated by an arrow). Fe(II) (black squares) and N2O-N (open circles) were measured over time. Error bars represent the standard deviation of results of duplicate cultures, and no error bars are shown if the standard deviation is smaller than the symbol.

Denitrification potential of bacteria lacking NO-forming NO2 reductases.

The NCBI Reference Sequence (RefSeq) database was searched for organisms that have at least one nitrate reductase gene (napA and/or narG) and nosZ genes and lack nirK and nirS genes. If also capable of Fe(III) reduction, these organisms could potentially occupy a previously unrecognized ecological niche as complete denitrifiers (NO3→N2) without the need for NO-forming NO2 reductases. Among 4,739 translated RefSeq genomes examined, 336 genomes had at least one nitrate reductase gene (napA or narG) and one N2O reductase gene (nosZ), of which 51 lacked nitrite reductase genes (nirK or nirS) (Fig. 4). The 51 genomes lacking nirK or nirS represented organisms from diverse bacterial taxa and from diverse habitats, including freshwater, saltwater, soil, contaminated sites, and hot springs (Fig. 4; Table S1). Sixteen genomes contained clade I nosZ genes, while 35 genomes contained clade II nosZ genes, supporting prior observations that clade II nosZ genes have relevant contributions to the fate of N2O (14, 15, 32). Strains of the species Aeromonas media and Rhodoferax ferrireducens, as well as members of the genera Anaeromyxobacter and Desulfitobacterium, are known Fe(III) reducers (Fig. 4; Table S1). Since many organisms have not been tested for the ability to reduce Fe(III), the genomes were further searched for c-type cytochromes with multiple heme-binding domains, a shared trait among many Fe(III)-reducing bacteria (33). All 51 genomes that fit the above-mentioned criteria contain 15 or more putative c-type cytochromes, and in 27 genomes, at least six of the c-type cytochromes have four or more predicted heme-binding domains (Fig. 4). Of course, these criteria are insufficient to prove that these organisms are indeed Fe(III) reducers, and physiological studies are required. Still, these observations suggest that a much broader organismal group not possessing nirS or nirK has the potential to perform denitrification by coupling biotic and abiotic reactions. The term chemodenitrifier is proposed to describe organisms that combine the chemical chemodenitrification reactions [i.e., Fe(II)-mediated NO2→N2O conversion] and an enzymatic reaction(s) to reduce NO3 to N2O or N2 (i.e., denitrification that is independent of NO-forming NO3 reductases). The organisms uncovered by the above analysis represent potential chemodenitrifiers, as they lack the keystone denitrification genes nirK and nirS yet have the potential to reduce NO3 to N2. Of note, not included in this analysis are potential chemodenitrifiers that lack nosZ and potential chemodenitrifiers that possess nirK or nirS but reduce NO2 to N2O via Fe(II)-mediated rather than enzymatic reactions.

FIG 4.

FIG 4

Maximum likelihood phylogenetic tree of the 16S rRNA genes from RefSeq bacterial genomes with NO3 reductase (narG and napA) and N2O reductase (nosZ) genes but lacking NO2 reductase (nirK or nirS) genes. All intergeneric nodes with bootstrap values of <50% are marked with red circles. Four archaeal sequences were included as outgroups, and the branch coloring reflects class taxonomy. For all bacteria, the estimated number of c-type cytochromes (with the number of c-type cytochromes with ≥4 predicted heme-binding domains in parentheses), demonstrated ability to reduce Fe(III) (indicated by a plus sign), the type of nosZ gene (i.e., clade I or clade II), and the source of the isolate are indicated. The absence of a plus sign for Fe(III) reduction indicates that this activity is unknown (most frequently, untested), and the absence of a symbol in the source column indicates that the source is unknown. The scale bar corresponds to the number of nucleotide substitutions per site. References for individual isolates are provided in Table S1 in the supplemental material. The Venn diagram shows the proportions of the analyzed RefSeq genomes containing nirK or nirS, napA or narG, and nosZ. Out of the 4,739 genomes analyzed, 2,567 had either a napA or a narG gene, 634 had a nirK or a nirS gene, and 394 had a nosZ gene. *, isolated from an alkaline lake.

DISCUSSION

This study reveals a heretofore unrecognized ecological niche for A. dehalogenans and organisms with shared physiological features that directly link the N and Fe cycles. Previous growth studies characterized A. dehalogenans as a NO3-reducing respiratory ammonifier that is also able to reduce N2O to N2 by utilizing a clade II NosZ (15, 29). Genome analysis corroborated these observations, and no genes responsible for NO2 reduction to NO (i.e., nirS or nirK) occur on the genome (34). Prior pure culture studies were performed with single electron acceptors, and potential synergistic effects were not recognized. The results presented here demonstrate that Fe(III) reduction profoundly affects NO3 and NO2 metabolism in A. dehalogenans (Fig. 5). Complete NO3 reduction to N2 was the result of a combination of biotic and abiotic reactions in A. dehalogenans cultures, indicating that cultivation on single electron acceptors and genome analysis failed to show that A. dehalogenans effectively acts as a denitrifier when Fe(III) and N oxides are present.

FIG 5.

FIG 5

Schematic illustration depicting NO3/NO2 reduction pathways observed in A. dehalogenans cultures in the absence versus the presence of bioavailable iron. A. dehalogenans performs respiratory ammonification (i.e., NO3→NH4, depicted by the red arrows in the top right portion of the figure) in the absence of reduced Fe(II). In the presence of Fe(II), denitrification is the major pathway facilitated by the abiotic reaction of Fe(II) with NO2 to form N2O. A. dehalogenans reduces Fe(III) generated from this reaction and is able to grow using oxidized Fe(III) as the electron acceptor. A. dehalogenans possesses a clade II NosZ to reduce N2O to N2. NosZ is a copper-dependent enzyme system and may not be functional in sulfidic environments with very low concentrations of bioavailable copper (lightning arrow). Straight arrows indicate reactions catalyzed by A. dehalogenans and linked to energy conservation; curved arrows depict abiotic reactions.

Combined biotic-abiotic reactions have been demonstrated in other NO3-reducing cultures, including cultures of Escherichia coli, Shewanella putrefaciens strain 200, Wolinella succinogenes, and Klebsiella mobilis. NO3-reducing cultures of E. coli were shown to abiotically produce N2O through chemodenitrification when Fe(II) was present in the medium (35). Cultures of S. putrefaciens were also shown to produce N2O during NO3 and Fe(III) reduction (21). N2O was the end product in experiments with E. coli and S. putrefaciens strain 200, in contrast to the experiments with A. dehalogenans, in which NO3 was completely reduced to N2. The formation and consumption of small amounts of N2O (0.15% of the NO3-N) were observed in NO3-replete cultures of W. succinogenes, a respiratory ammonifier possessing a clade II nos operon (36). Like Anaeromyxobacter, W. succinogenes lacks nirK and nirS, and it was speculated N2O was produced through reactions of NO2 with medium components and NO detoxification (36). In the presence of Fe(III) and NO3, A. dehalogenans produced substantial amounts of N2O (i.e., greater than 50%) from available NO3 or NO2 and has the enzymatic capacity of subsequently reducing the N2O to N2. In contrast to studies with E. coli, S. putrefaciens, and W. succinogenes, the work with Anaeromyxobacter demonstrates that a bacterium lacking NO-forming NO2 reductases can effectively reduce NO3/NO2 to N2 via coupled biotic-abiotic reactions. It should be noted that studies with K. mobilis demonstrated abiotic NH4+ production via NO2 reaction with green rust-associated Fe(II) (23); however, evidence for abiotic NH4+ production was not observed in any of the incubation studies with A. dehalogenans.

Due to a versatile metabolism, A. dehalogenans is an ecologically competitive bacterium in various ecosystems (15, 29, 37). Previous studies uncovered the novel clade II nosZ genes and demonstrated efficient growth of A. dehalogenans coupled to N2O reduction (15, 17). The clade II nos operons provide greater growth yields (i.e., more efficient energy capture), and bacteria expressing clade II NosZ exhibit greater affinity to N2O than bacteria with clade I NosZ, suggesting that organisms with clade II NosZ outcompete organisms with clade I NosZ at low (i.e., environmentally relevant) N2O concentrations (17). In the presence of Fe(II), chemodenitrification prevents A. dehalogenans from effectively coupling growth with NO2 reduction to NH4+; however, the organism has the potential to conserve energy from N2O to N2 reduction (15, 17). Respiratory ammonification and complete NO3 reduction to N2 (via combined biotic reactions and chemodenitrification) are both associated with similar negative Gibbs free energy changes (ΔG′) (Table 2), suggesting that both of these pathways are equally favorable from an energetic perspective. A. dehalogenans captures more energy on a per electron basis from N2O to N2 reduction than from NO2 to NH4+ reduction (i.e., the reduction of 0.5 mol of N2O yields 2.4 × 1013 cells compared to a yield of 2.0 × 1013 cells from the reduction of 0.16 mol of NO2) (Table 3, equations 3 and 1, respectively). Additionally, the reaction between NO2 and Fe(II) yields Fe(III), which serves as an electron acceptor for A. dehalogenans. By coupling respiratory [NO3→NO2, N2O→N2, and Fe(III)→Fe(II)] and abiotic [NO2→N2O, Fe(II)→Fe(III)] reactions (Fig. 5), A. dehalogenans has the potential to produce more biomass than it does via respiratory ammonification (NO3→NH4+), demonstrating an ecological strategy to optimize energy capture and competitiveness. Collectively, the experiments presented here and past N2O growth experiments (15, 17) demonstrate that the nondenitrifier A. dehalogenans can reduce NO3 to N2 in the presence of Fe(III)/Fe(II), and the organism potentially benefits from this coupled biotic-abiotic process by conserving more energy than it does via the respiratory ammonification pathway.

TABLE 2.

Free energy changes (ΔG′) associated with different N and Fe cycling processes at pH 7.0 and 30°C

Equation Reaction Free energy changes (kJ/reaction)a
Free energy boundaries for total reactions (kJ/reaction)b
Initial Final Maximum Minimum
1 NO3 + 0.25 CH3COO → NO2 + 0.25 H+ + 0.5 HCO3 −152.5 (AC) −137.1 (BD)
2 2 Fe3+(aq) + H2O + 0.25 CH3COO → 2.25 H+ + 2 Fe2+(aq) + 0.5 HCO3 −247.8 (GC) −206.2 (HD)
3 NO2 + 2 Fe2+(aq) + 4.5 H2O → 0.5 N2O(g) + 3 H+ + 2 Fe(OH)3(amorph) −169.6 (BHE) −108.4 (AGF)
4 2 Fe(OH)3(amorph) + 3.75 H+ + 0.25 CH3COO → 5 H2O + 2 Fe2+(aq) + 0.5 HCO3 −52.3 (GC) −52.3 (HD)
5 0.5 N2O(g) + 0.125 CH3COO → 0.5 N2(g) + 0.125 H+ + 0.25 HCO3 −136.8 (ED) −154.2 (FD)
6 NO2 + 0.75 CH3COO + 1.25 H+ + H2O → NH4+ + 1.5 HCO3 −386.1 (BC) −363.2 (AD)
7 NO3 + 0.875 CH3COO + 2 Fe3+(aq) + 0.5 H2O → 0.5 N2(g) + 2 Fe2+ + 1.75 HCO3 + 1.875 H+ −606.8 −532.4
8 NO3 + 0.875 CH3COO + 2 Fe(OH)3(amorph) + 4.125 H+ → 0.5 N2(g) + 2 Fe2+ + 1.75 HCO3 + 5.5 H2O −411.3 −378.5
9 NO3 + CH3COO + 1.0 H+ + H2O → NH4+ + 2 HCO3 −538.6 −500.3
a

Initial and final ΔG′ values were calculated in Geochemists Workbench (Aqueous Solutions, LLC). Initial free energy change values were calculated using the concentrations of reactants/products present at the beginning of the reactions. Final free energy change values were calculated using the concentrations of reactants/products measured or estimated at the end of the reactions. Bold values indicate an abiotic reaction not directly coupled to microbial energy metabolism. Concentrations used for calculations reflecting the experimental conditions were as follows: NO3, 1 mM; NO2, 0.01 mM (A) or 1.0 mM (B); Fe3+ (ferric citrate), 8 mM; CH3COO, 10 mM; HCO3, 0.1 mM (C) or 2.0 mM (D); N2O(g), 0.01 ppmv (E) or 9,600 ppmv (F); Fe2+, 0.01 mM (G) or 6.0 mM (H); NH4+, 3 mM. Letters in parentheses refer to the concentrations noted for the reactant or the product noted. amorph, amorphous; aq, aqueous.

b

The total free energy changes associated with complete NO3 conversion to N2 (equations 7 and 8) or NH4+ (equation 9) were calculated by adding the ΔG′ values for the individual reactions (equations 1 to 6). The values reported as maximums represent the highest possible free energy change (i.e., most-negative ΔG′ values) associated with the respective reactions, while the values reported as minimums represent the lower threshold for free energy changes (i.e., the most-positive ΔG′ values). To calculate the maximum free energy values, the largest (i.e., most-negative) free energy change values from the individual reactions (equations 1 to 6) were added together. To calculate the minimum free energy values, the smallest (i.e., most-positive) free energy change values were added. Note that the ΔG′ value for Fe3+ was added twice in equation 8 to account for amorphous Fe3+ produced during chemodenitrification. Equation 7 was calculated by adding equations 1, 2, 4, and 5. Equation 8 was calculated by adding equations 1, 4 (twice), and 5. Equation 9 was calculated by adding equations 1 and 6.

TABLE 3.

Theoretical growth yields per mole of electrons

Equation Process Reduction half reaction Growth yield (no. of cells/mol e) Reference
1 NO2 reduction to NH4+ 0.16 NO2 + 1 e + 4.33 H+ → 0.16 NH4+ + 0.33 H2O 2 × 1013a R. A. Sanford, unpublished data
2 Fe(III) reduction Fe3+ + 1 e → Fe2+ 6.0 × 1012b 28
3 N2O reduction 0.5 N2O + 1 e + 1 H+ → 1/2 N2 + H2O 2.4 × 1013b 17
4 Chemodenitrification 1.5 × 1013 This studyc
a

Based on Anaeromyxobacter dehalogenans strain R cell enumeration with flow cytometry.

b

Based on published numbers [1.4 mg biomass/mmol Fe(III) and 11.2 mg biomass/mmol N2O] and a weight of 2.4 × 10−13 g per bacterial cell.

c

Theoretical calculation based on the reduction of 0.5 mol Fe(III) and 0.25 mol N2O [produced by chemical reaction between NO2 and Fe(II)].

An initially perplexing observation was the inability of A. dehalogenans to grow with N2O produced by abiotic reactions (see Fig. S5 in the supplemental material) when the medium received sulfide as a reducing agent. No such growth inhibition was observed with NO3 and ferric iron as electron acceptors, suggesting that the observed growth inhibition was due to the sequestration of copper (38). The requirement of copper for N2O reductases suggests that chemodenitrifiers such as A. dehalogenans contribute to N2O production in sulfidic (that is, copper-limited) environments (Fig. 5).

The coupled biotic-abiotic reactions in NO3- and Fe(III)-reducing A. dehalogenans cultures reveal an important ecological role for organisms that possess nos operons. Fe(II) and NO3 cooccur in many natural environments, allowing interactions between N and Fe cycling (for an example, see reference 65) and suggesting that microbially mediated chemodenitrification occurs broadly. Soils contain 100 to 100,000 ppm iron, and in addition to Fe(III) serving as an electron acceptor in anoxic soils, Fe(III) and Fe(II) cycling commonly occurs near oxic-anoxic interfaces, thus replenishing the Fe(III) pool for dissimilatory Fe(III) reduction (4043). Dissimilatory Fe(III)-reducing bacteria such as A. dehalogenans and Geobacter spp. are ubiquitous in soils and subsurface environments (44, 45). During NO3 reduction and ammonia oxidation, NO2 is released temporarily (Fig. S2) (46, 47), and NO2 can accumulate for weeks to months in both alkaline and neutral soils (48, 49). Additionally, the enzyme kinetics of NO3 and NO2 reductases vary by organism, and for organisms that reduce NO3 at higher rates than NO2, intermediate NO2 forms prior to further NO2 reduction (46). Fe(III) reduction occurs outside the cell or in the periplasm (50), and NO2 reduction occurs in the periplasm, suggesting that reactions with NO2 can occur even under conditions where NO2 does not accumulate extracellularly. When NO2 and Fe(II) cooccur outside the cytoplasm, enzymatic NO2 reduction must compete with abiotic chemistry (i.e., chemodenitrification). The competition between abiotic and biotic reduction of NO2 is evidenced by the mixed end products (i.e., N2O and NH4+) produced during NO3 reduction by A. dehalogenans (Fig. 1). The occurrence of Fe(II) and NO2 outside the cell suggests that microbially mediated chemodenitrification is not limited to organisms that can reduce both Fe(III) and NO3. A microbial community with Fe(III)- and NO3-reducing microorganisms could facilitate chemodenitrification, especially under conditions where NO2 is not immediately consumed following NO3 reduction. As NO3-reducing microorganisms generate NO2, NO2 can react with Fe(II) produced by Fe(III) reducers. Chemodenitrification results in the loss of NO2 as an electron acceptor for microbial metabolism; however, organisms with N2O reductases, in particular those that express high-affinity clade II NosZ (17), would be able to utilize the resulting N2O as an electron acceptor. The possession of the nosZ gene would therefore confer an ecological advantage in anoxic environments with dissimilatory Fe(III)-reducing activity.

Chemodenitrifiers such as A. dehalogenans fill a previously unrecognized ecological niche. Chemodenitrification merely describes the abiotic decomposition of NO2 coupled to Fe(II) oxidation, whereas the term chemodenitrifier describes a microorganism that reduces NO3 to N2O or N2 via combined enzymatic and abiotic steps. These organisms may be (i) true denitrifiers that possess nirS or nirK but reduce NO2 to N2O abiotically in the presence of Fe(II), or (ii) nondenitrifiers (i.e., organisms that lack nirS or nirK genes) that can denitrify only in the presence of Fe(II). Past studies have examined the cooccurrence of nirK, nirS, and nosZ in sequenced genomes, demonstrating that denitrification is a modular pathway and many genomes contain nosZ but lack nirK and nirS (15, 32). The observation of chemodenitrification in A. dehalogenans cultures reveals a possible ecophysiological explanation for nondenitrifiers that possess nosZ. The cooccurrence study of nirK, nirS, nosZ, and napA and narG (the last two of which were not included in the above-mentioned studies) revealed that nondenitrifiers (i.e., genomes lacking nirS and nirK) possess nosZ, napA and/or narG, and multiple genes encoding multiheme proteins [i.e., may have the Fe(III)-reducing phenotype] (Fig. 4; Table S1). Although a general set of biomarkers for Fe(III) reduction has not yet emerged, a suite of c-type cytochromes (typically with four or more heme binding sites) are indicated in Fe(III) reduction activity by at least some Fe(III)-reducing bacteria (33, 50). The identified potential chemodenitrifiers span multiple bacterial classes occurring in diverse habitats, suggesting that chemodenitrifiers are dispersed across the phylogenetic tree and represent previously unrecognized contributors to NO3 reduction to N2. The chemodenitrifier potential is a prediction based on genome sequence analysis, and experimental efforts are needed to confirm this physiology [that is, NO3/Fe(III) reduction activity] for the organisms included in Fig. 4. Additionally, not all Fe(III)-reducing bacteria require c-type cytochromes (39), suggesting that the prediction of potential iron-reducing bacteria by searching for c-type cytochromes is a conservative approach and likely an underestimation of potential chemodenitrifiers. Physiological studies and measurements of the NO3 reduction products of potential chemodenitrifiers will reveal if these organisms indeed utilize the chemodenitrifier ecological strategy demonstrated for A. dehalogenans.

Collectively, our findings show the reduction of NO3 to N2 mediated by a single organism via combined abiotic and biotic mechanisms in the absence of nirS and nirK (Fig. 5). The findings emphasize that gene content is not sufficient to predict the final products of NO3 reduction, putting into question attempts to link denitrification activity with the abundance and expression of nirS and nirK. Chemodenitrifiers represent understudied contributors to the formation of gaseous products from N-oxyanions, and the study of chemodenitrifiers may impact our current understanding of N loss to the atmosphere. Nondenitrifiers lacking nirS or nirK but possessing napA, narG, and/or nosZ contribute to N2O flux via coupled biotic-abiotic processes, and a holistic understanding of N flux can be achieved only when integrated studies of interlinked biogeochemical cycles are performed.

MATERIALS AND METHODS

Chemicals.

HEPES (4-[2-hydroxyethyl]-1-piperazineethanesulfonic acid) (≥99%), l-cysteine hydrochloride monohydrate (98.5 to 101.0%), anhydrous sodium acetate (≥99.2%), sodium hydroxide pellets (≥99.0%), sodium nitrite (≥99.6%), ammonium chloride, ferric chloride hexahydrate (≥99.7%), and ferrous ammonium sulfate hexahydrate (101.2%) were purchased from Fisher Scientific (Fair Lawn, NJ). N2O (99%), Fe(III) citrate, and ferrous chloride tetrahydrate (≥99.0%) were purchased from Sigma-Aldrich (St. Louis, MO). Sodium phosphate dibasic (≥99.6%) and sodium nitrate (100.8%) were purchased from J. T. Baker (Phillipsburg, NJ). Ferrozine iron reagent (≥98%) and sodium sulfide nonahydrate were purchased from Acros Organics (Fair Lawn, NJ). Hydrochloric acid (36.5 to 38% [wt/wt]) was purchased from Aqua Solutions, Inc. (Deer Park, TX). Acetylene (99.6%) and compressed nitrogen (ultrahigh purity) were purchased from AirGas (Knoxville, TN).

Culture conditions.

Anaeromyxobacter dehalogenans strain 2CP-C was grown in 100 ml anoxic defined mineral salts medium in 160-ml glass serum bottles sealed with butyl rubber stoppers (Geo-Microbial Technologies, Ochelata, OK) as described previously (51) with the following modifications: bicarbonate was replaced with 50 mM HEPES, sodium sulfide was omitted following the recognition that sulfide inhibited NosZ activity, phosphate was reduced from 1.47 mM to 0.29 mM, l-cysteine was decreased from 0.2 mM to 0.1 mM, acetate (1 or 10 mM) was included as the electron donor, and the headspace was 100% N2. The medium pH was adjusted to 7.2 by adding 10 M NaOH. Fe(III) citrate and NO3 were added from anoxic, autoclaved stock solutions, and N2O gas and vitamins were added via filter sterilization with Nalgene polyethersulfone syringe filter units (0.2 μm) (Thermo Scientific, Rochester, NY). For growth with NO3 as the electron acceptor, the NH4+ concentrations in the medium were decreased from 5.6 mM to 3 mM to allow more accurate quantification. The amount of NH4+ produced is reported as total micromoles of NH4+ minus the 300 μmol NH4+ present in each vessel. In control vessels, 10% of the 60-ml N2 headspace was replaced with acetylene to inhibit N2O conversion to N2 (52). In abiotic (i.e., no inoculum) control vessels, ferrous chloride and NO2 were added to final concentrations of 8 mM and 1 mM, respectively. All vessels were incubated without shaking at 30°C in the dark with the stoppers facing upward. Cultures grown with the same electron donors and acceptors as the experimental cultures were used as seed cultures, with the exception of cultures amended with Fe(III) oxyhydroxide, which were inoculated from fumarate-grown cultures (29). Sulfide (0.2 mM) was initially included as a reducing agent in cultures that received Fe(III) oxyhydroxide as the electron acceptor. All experiments used triplicate culture vessels, unless indicated otherwise, and were repeated independently at least twice.

The Fe(III) citrate stock solution was prepared by adding 24.5 g of ferric citrate to approximately 90 ml of a 1.7 M NaOH solution. The solution was boiled until it turned dark, translucent brown. After cooling, the resulting acidic solution was neutralized (pH 7.0) by slowly adding approximately 10 ml of 10 M NaOH. The final stock solution was autoclaved and contained approximately 0.6 M Fe(III) and 0.1 to 0.2 M Fe(II), as determined by the ferrozine assay (53). Poorly crystalline Fe(III) oxide was prepared from ferric chloride hexahydrate as described previously (45). Fe(III) oxide stocks were sterilized by daily heating in a 90°C water bath and overnight cooling at room temperature over a 5-day period. A ferrous chloride stock solution was prepared inside an anoxic chamber (Coy Laboratory Products, Grass Lake, MI) by adding 9.9 g of ferrous chloride tetrahydrate to 100 ml of ultrapure water degassed with N2, and the green solution was stored in N2-gassed, sealed glass serum bottles after filter sterilization.

Analytical methods.

N2O in the headspace of the incubation vessels was quantified with a gas chromatograph as described previously (15, 54). To account for cooling during sample extraction (due to removal from the 30°C incubator), the temperature of an uninoculated medium bottle was measured with a glass thermometer and used to select Ostwald coefficients of 0.6788, 0.5937, and 0.5241 for temperatures of 20°C, 25°C, and 30°C, respectively (55). Total N2O-N (i.e., the total N2O mass multiplied by a factor of two) in culture vessels is reported as the sum of the aqueous phase N2O-N, the headspace N2O-N, and any N2O-N that was removed during liquid and headspace sampling. N2O production rates were calculated by graphing N2O-N over time to determine the linear range of N2O production and are reported in micromoles per hour. Acetate was quantified by high-performance liquid chromatography (1200 series; Agilent Technologies, Santa Clara, CA) using an organic acid analysis column (Aminex HPX-87H ion exclusion column; Bio-Rad, Hercules, CA) heated to 30°C. The eluent was 4 mM sulfuric acid prepared in Milli-Q water, and the flow rate was 0.6 ml min−1.

NH4+ was quantified with an ion chromatograph (Dionex ICS-1100) and a Dionex IonPac cation exchange CS16 column heated to 30°C. The eluent was 20 mM sulfuric acid, and the flow rate was 1 ml min−1. NO3 and NO2 were quantified by ion chromatography using a reagent-free eluent regeneration system (ICS-2100; Dionex, Sunnyvale, CA) and a Dionex IonPac AS18 4- by 250-mm analytical column heated to 30°C. The eluent was 10 mM KOH, and the flow rate was 1 ml min−1. Acid-extractable Fe(III)/Fe(II) were measured using a modification of the ferrozine assay (53). Briefly, ferrozine buffer was prepared with 1 g/liter Acros Organics FerroZine iron reagent dissolved in 50 mM HEPES buffer (56). Cultures were shaken vigorously by hand, suspensions were withdrawn by syringe, and 100 or 500 μl was incubated overnight in 10-ml glass vials containing 5 ml of 0.04 M sulfamic acid, which was used to prevent a chemical reaction of NO2 with Fe(II) (22). After the vials were shaken, 20-μl aliquots were withdrawn and mixed with 1 ml of ferrozine buffer. Absorbance was measured using a Thermo Scientific Spectronic 20D+ spectrophotometer at a wavelength of 562 nm. Samples were measured quickly after mixing with the ferrozine buffer to avoid interference with Fe(III) (56). Standards were prepared with ferrous ammonium sulfate hexahydrate spanning a concentration range of 1.0 mM to 20 mM Fe(II). For experiments in which NO2 was added to Fe(III) citrate-grown cultures, 0.5 M HCl was used in lieu of sulfamic acid, and Fe(II) concentrations are not reported when NO2 was present in the culture due to potential reactions between Fe(II) and NO2 after sampling.

Cell enumeration.

Genomic DNA was extracted using the protocol provided in the MoBio PowerSoil DNA isolation kit (Carlsbad, CA). Cells were enumerated with an established TaqMan qPCR assay (37) using the Applied Biosystems ViiA 7 real-time PCR system (Foster City, CA). Results were analyzed with the ViiA 7 software (version 1.2.3).

Genomic and phylogenetic analyses.

Prokaryotic genomes (4,739) were searched for genes encoding N cycle enzymes (i.e., NarG, NapA, NirK, NirS, and NosZ) by performing a Hidden Markov Model (HMM) search with HMMER 3.1b2 (http://www.hmmer.org). HMMs for NarG, NapA, NirK, NirS, and NosZ were downloaded from FunGene (57). The sequences of annotated proteins of available complete bacterial genomes (i.e., labeled as “Latest” and “Complete Genome”) from the National Center for Biotechnology Information (NCBI) Reference Sequence (RefSeq) database were downloaded in April 2016 and used as the target files. HMMER was run locally with a target E value cutoff of ≤1 × 10−10 using a BASH script. Results were parsed and analyzed using Python (https://www.python.org/). False positives were removed by searching for the absence of conserved domains visually in Geneious software suite version 8.1.9 (58) and with the Reverse Position-Specific Basic Local Alignment Tool (RPS-BLAST) (59). The NCBI Conserved Domains Database (CDD) was downloaded in September 2016 from the NCBI FTP website (ftp://ftp.ncbi.nih.gov/pub/mmdb/cdd/little_endian/Cdd_LE.tar.gz) for a local RPS-BLAST. Proteins that did not match conserved domains with an E value of <1 × 10−30 were excluded from further analysis.

Available bacterial 16S rRNA gene sequences of organisms containing narG, napA, and nosZ genes, but lacking nirK and nirS genes, and four outgroup archaeal 16S rRNA genes were downloaded from NCBI and used to construct a phylogenetic tree. Full-length 16S rRNA gene sequences were aligned with the SILVA Incremental Aligner version 1.2.11 (60) using default settings with unaligned bases removed from the alignment ends. Aligned positions containing only gaps were manually removed in Geneious. JModelTest 2 (61) was used to predict the optimal evolutionary model for tree reconstruction. A maximum likelihood tree was constructed with 100 bootstrap replicates with PhyML 3.0 (62) using a generalized time reversible (GTR) model (63) with gamma distributed rate heterogeneity and a proportion of invariable sites.

The organisms represented in the phylogenetic tree were searched in the literature for the ability to reduce Fe(III). The numbers of c-type cytochrome genes per genome were approximated by using a Python script modeled after a published script (64), which identified the heme binding motifs CxxCH, CxxxCH, and CxxxxCH in the annotated proteins of RefSeq genomes. Candidate c-type cytochromes were further screened by running PSI-BLAST against the nr NCBI database (downloaded in December 2016) and eliminating any sequences that did not match annotated c-type cytochromes with an E value of ≤1 × 10−11.

Supplementary Material

Supplemental material

ACKNOWLEDGMENTS

This work was supported through funding by the U.S. Department of Energy, Office of Biological and Environmental Research, Genomic Science Program, award DE-SC0006662.

This article was authored by UT-Battelle, LLC, under contract DE-AC05-00OR22725 with the U.S. Department of Energy (DOE).

The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.

We acknowledge Steven Higgins and Luis Orellana for helpful discussions.

We declare no conflicts of interest with the contents of this article.

Footnotes

Supplemental material for this article may be found at https://doi.org/10.1128/AEM.01985-17.

REFERENCES

  • 1.Fields S. 2004. Global nitrogen: cycling out of control. Environ Health Perspect 112:A556–A563. doi: 10.1289/ehp.112-a556. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Stein LY. 2011. Surveying N2O-producing pathways in bacteria. Methods Enzymol 486:131–152. doi: 10.1016/B978-0-12-381294-0.00006-7. [DOI] [PubMed] [Google Scholar]
  • 3.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]
  • 4.Philippot L, Čuhel J, Saby NPA, Chèneby D, Chroňáková A, Bru D, Arrouays D, Martin-Laurent F, Šimek M. 2009. Mapping field-scale spatial patterns of size and activity of the denitrifier community. Environ Microbiol 11:1518–1526. doi: 10.1111/j.1462-2920.2009.01879.x. [DOI] [PubMed] [Google Scholar]
  • 5.Enwall K, Throbäck IN, Stenberg M, Söderström M, Hallin S. 2010. Soil resources influence spatial patterns of denitrifying communities at scales compatible with land management. Appl Environ Microbiol 76:2243–2250. doi: 10.1128/AEM.02197-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Morales SE, Cosart T, Holben WE. 2010. Bacterial gene abundances as indicators of greenhouse gas emission in soils. ISME J 4:799–808. doi: 10.1038/ismej.2010.8. [DOI] [PubMed] [Google Scholar]
  • 7.Lee S-H, Kang H. 2016. The activity and community structure of total bacteria and denitrifying bacteria across soil depths and biological gradients in estuary ecosystem. Appl Microbiol Biotechnol 100:1999–2010. doi: 10.1007/s00253-015-7111-2. [DOI] [PubMed] [Google Scholar]
  • 8.Wallenstein MD, Myrold DD, Firestone M, Voytek M. 2006. Environmental controls on denitrifying communities and denitrification rates: insights from molecular methods. Ecol Appl 16:2143–2152. doi: 10.1890/1051-0761(2006)016[2143:ECODCA]2.0.CO;2. [DOI] [PubMed] [Google Scholar]
  • 9.Dandie CE, Burton DL, Zebarth BJ, Henderson SL, Trevors JT, Goyer C. 2008. Changes in bacterial denitrifier community abundance over time in an agricultural field and their relationship with denitrification activity. Appl Environ Microbiol 74:5997–6005. doi: 10.1128/AEM.00441-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Ma WK, Bedard-Haughn A, Siciliano SD, Farrell RE. 2008. Relationship between nitrifier and denitrifier community composition and abundance in predicting nitrous oxide emissions from ephemeral wetland soils. Soil Biol Biochem 40:1114–1123. doi: 10.1016/j.soilbio.2007.12.004. [DOI] [Google Scholar]
  • 11.Henderson SL, Dandie CE, Patten CL, Zebarth BJ, Burton DL, Trevors JT, Goyer C. 2010. Changes in denitrifier abundance, denitrification gene mRNA levels, nitrous oxide emissions, and denitrification in anoxic soil microcosms amended with glucose and plant residues. Appl Environ Microbiol 76:2155–2164. doi: 10.1128/AEM.02993-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Wei W, Isobe K, Nishizawa T, Zhu L, Shiratori Y, Ohte N, Koba K, Otsuka S, Senoo K. 2015. Higher diversity and abundance of denitrifying microorganisms in environments than considered previously. ISME J 9:1954–1965. doi: 10.1038/ismej.2015.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Bonilla-Rosso G, Wittorf L, Jones CM, Hallin S. 2016. Design and evaluation of primers targeting genes encoding NO-forming nitrite reductases: implications for ecological inference of denitrifying communities. Sci Rep 6:39208. doi: 10.1038/srep39208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Jones CM, Graf DRH, Bru D, Philippot L, Hallin S. 2013. The unaccounted yet abundant nitrous oxide-reducing microbial community: a potential nitrous oxide sink. ISME J 7:417–426. doi: 10.1038/ismej.2012.125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Sanford RA, Wagner DD, Wu Q, Chee-Sanford JC, Thomas SH, Cruz-García C, Rodríguez G, Massol-Deyá A, Krishnani KK, Ritalahti KM, Nissen S, Konstantinidis KT, Löffler FE. 2012. Unexpected nondenitrifier nitrous oxide reductase gene diversity and abundance in soils. Proc Natl Acad Sci U S A 109:19709–19714. doi: 10.1073/pnas.1211238109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Orellana LH, Rodriguez-R LM, Higgins S, Chee-Sanford JC, Sanford RA, Ritalahti KM, Löffler FE, Konstantinidis KT. 2014. Detecting nitrous oxide reductase (nosZ) genes in soil metagenomes: method development and implications for the nitrogen cycle. mBio 5:e01193-14. doi: 10.1128/mBio.01193-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Yoon S, Nissen S, Park D, Sanford RA, Löffler FE. 2016. Nitrous oxide reduction kinetics distinguish bacteria harboring clade I NosZ from those harboring clade II NosZ. Appl Environ Microbiol 82:3793–3800. doi: 10.1128/AEM.00409-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Zhu X, Silva LCR, Doane TA, Horwath WR. 2013. Iron: the forgotten driver of nitrous oxide production in agricultural soil. PLoS One 8:e60146. doi: 10.1371/journal.pone.0060146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Fan X, Yu H, Wu Q, Ma J, Xu H, Yang J, Zhuang Y. 2016. Effects of fertilization on microbial abundance and emissions of greenhouse gases (CH4 and N2O) in rice paddy fields. Ecol Evol 6:1054–1063. doi: 10.1002/ece3.1879. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Moraghan JT, Buresh RJ. 1977. Chemical reduction of nitrite and nitrous oxide by ferrous iron. Soil Sci Soc Am J 41:47–50. doi: 10.2136/sssaj1977.03615995004100010017x. [DOI] [Google Scholar]
  • 21.Cooper DC, Picardal FW, Schimmelmann A, Coby AJ. 2003. Chemical and biological interactions during nitrate and goethite reduction by Shewanella putrefaciens 200. Appl Environ Microbiol 69:3517–3525. doi: 10.1128/AEM.69.6.3517-3525.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Klueglein N, Kappler A. 2013. Abiotic oxidation of Fe(II) by reactive nitrogen species in cultures of the nitrate-reducing Fe(II) oxidizer Acidovorax sp. BoFeN1—questioning the existence of enzymatic Fe(II) oxidation. Geobiology 11:180–190. [DOI] [PubMed] [Google Scholar]
  • 23.Etique M, Jorand FPA, Zegeye A, Grégoire B, Despas C, Ruby C. 2014. Abiotic process for Fe(II) oxidation and green rust mineralization driven by a heterotrophic nitrate reducing bacteria (Klebsiella mobilis). Environ Sci Technol 48:3742–3751. doi: 10.1021/es403358v. [DOI] [PubMed] [Google Scholar]
  • 24.Jones LC, Peters B, Lezama Pacheco JS, Casciotti KL, Fendorf S. 2015. Stable isotopes and iron oxide mineral products as markers of chemodenitrification. Environ Sci Technol 49:3444–3452. doi: 10.1021/es504862x. [DOI] [PubMed] [Google Scholar]
  • 25.Buchwald C, Grabb K, Hansel CM, Wankel SD. 2016. Constraining the role of iron in environmental nitrogen transformations: dual stable isotope systematics of abiotic NO2 reduction by Fe(II) and its production of N2O. Geochim Cosmochim Acta 186:1–12. doi: 10.1016/j.gca.2016.04.041. [DOI] [Google Scholar]
  • 26.Zhu-Barker X, Cavazos AR, Ostrom NE, Horwath WR, Glass JB. 2015. The importance of abiotic reactions for nitrous oxide production. Biogeochemistry 126:251–267. doi: 10.1007/s10533-015-0166-4. [DOI] [Google Scholar]
  • 27.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]
  • 28.He Q, Sanford RA. 2003. Characterization of Fe(III) reduction by chlororespiring Anaeromxyobacter dehalogenans. Appl Environ Microbiol 69:2712–2718. doi: 10.1128/AEM.69.5.2712-2718.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Sanford RA, Cole JR, Tiedje JM. 2002. Characterization and description of Anaeromyxobacter dehalogenans gen. nov., sp. nov., an aryl-halorespiring facultative anaerobic myxobacterium. Appl Environ Microbiol 68:893–900. doi: 10.1128/AEM.68.2.893-900.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.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]
  • 31.Hurt RA, Robeson MS II, Shakya M, Moberly JG, Vishnivetskaya TA, Gu B, Elias DA. 2014. Improved yield of high molecular weight DNA coincides with increased microbial diversity access from iron oxide cemented sub-surface clay environments. PLoS One 9:e102826. doi: 10.1371/journal.pone.0102826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Graf DRH, Jones CM, Hallin S. 2014. Intergenomic comparisons highlight modularity of the denitrification pathway and underpin the importance of community structure for N2O emissions. PLoS One 9:e114118. doi: 10.1371/journal.pone.0114118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Shi L, Squier TC, Zachara JM, Fredrickson JK. 2007. Respiration of metal (hydr)oxides by Shewanella and Geobacter: a key role for multihaem c-type cytochromes. Mol Microbiol 65:12–20. doi: 10.1111/j.1365-2958.2007.05783.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Thomas SH, Wagner RD, Arakaki AK, Skolnick J, Kirby JR, Shimkets LJ, Sanford RA, Löffler FE. 2008. The mosaic genome of Anaeromyxobacter dehalogenans strain 2CP-C suggests an aerobic common ancestor to the delta-proteobacteria. PLoS One 3:e2103. doi: 10.1371/journal.pone.0002103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Brons HJ, Hagen WR, Zehnder AJB. 1991. Ferrous iron dependent nitric oxide production in nitrate reducing cultures of Escherichia coli. Arch Microbiol 155:341–347. doi: 10.1007/BF00243453. [DOI] [PubMed] [Google Scholar]
  • 36.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]
  • 37.Thomas SH, Sanford RA, Amos BK, Leigh MB, Cardenas E, Löffler FE. 2010. Unique ecophysiology among U(VI)-reducing bacteria as revealed by evaluation of oxygen metabolism in Anaeromyxobacter dehalogenans strain 2CP-C. Appl Environ Microbiol 76:176–183. doi: 10.1128/AEM.01854-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Manconi I, van der Maas P, Lens P. 2006. Effect of copper dosing on sulfide inhibited reduction of nitric and nitrous oxide. Nitric Oxide 15:400–407. doi: 10.1016/j.niox.2006.04.262. [DOI] [PubMed] [Google Scholar]
  • 39.Dong Y, Sanford RA, Boyanov MI, Kemner KM, Flynn TM, O'Loughlin EJ, Chang Y, Locke RA Jr, Weber JR, Egan SM, Mackie RI, Cann I, Fouke BW. 2016. Orenia metallireducens sp. nov. strain Z6, a novel metal-reducing member of the phylum Firmicutes from the deep subsurface. Appl Environ Microbiol 82:6440–6453. doi: 10.1128/AEM.02382-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Shacklette HT, Boerngen J. 1984. Element concentrations in soils and other surficial materials of the conterminous United States. U.S. Geological Survey Professional Paper 1270. U.S. Government Printing Office, Washington, DC. [Google Scholar]
  • 41.Thamdrup B. 2000. Bacterial manganese and iron reduction in aquatic sediments, p 41–83. In Schink B. (ed), Advances in microbial ecology, vol 16. Kluwer Academic/Plenum Publishers, New York, NY. [Google Scholar]
  • 42.Cornell RM, Schwertmann U. 2003. The iron oxides: structure, properties, reactions, occurrences and uses, 2nd ed Wiley-VCH, Weinheim, Germany. [Google Scholar]
  • 43.Roden EE, McBeth JM, Blöthe M, Percak-Dennett EM, Fleming EJ, Holyoke RR, Luther GW III, Emerson D, Schieber J. 2012. The microbial ferrous wheel in a neutral pH groundwater seep. Front Microbiol 3:172. doi: 10.3389/fmicb.2012.00172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Petrie L, North NN, Dollhopf SL, Balkwill DL, Kostka JE. 2003. Enumeration and characterization of iron(III)-reducing microbial communities from acidic subsurface sediments contaminated with uranium(VI). Appl Environ Microbiol 69:7467–7479. doi: 10.1128/AEM.69.12.7467-7479.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Lovley D. 2013. Dissimilatory Fe(III)- and Mn(IV)-reducing prokaryotes, p 287–308. In Rosenberg E, DeLong EF, Lory S, Stackebrandt E, Thompson F (ed), The prokaryotes. Springer, Berlin, Germany. [Google Scholar]
  • 46.Betlach MR, Tiedje JM. 1981. Kinetic explanation for accumulation of nitrite, nitric oxide, and nitrous oxide during bacterial denitrification. Appl Environ Microbiol 42:1074–1084. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Matocha CJ, Dhakal P, Pyzola SM. 2012. The role of abiotic and coupled biotic/abiotic mineral controlled redox processes in nitrate reduction, p 181–214. In Advances in agronomy. Elsevier, Inc., San Diego, CA. [Google Scholar]
  • 48.Cleemput O Van, Samater AH. 1996. Nitrite in soils: accumulation and role in the formation of gaseous N compounds. Fertil Res 45:81–89. doi: 10.1007/BF00749884. [DOI] [Google Scholar]
  • 49.Gelfand I, Yakir D. 2008. Influence of nitrite accumulation in association with seasonal patterns and mineralization of soil nitrogen in a semi-arid pine forest. Soil Biol Biochem 40:415–424. doi: 10.1016/j.soilbio.2007.09.005. [DOI] [Google Scholar]
  • 50.Lovley DR, Holmes DE, Nevin KP. 2004. Dissimilatory Fe(III) and Mn(IV) reduction, p 219–286. In Poole RK. (ed), Advances in microbial physiology. Academic Press, London, England. [DOI] [PubMed] [Google Scholar]
  • 51.Löffler FE, Sanford RA, Tiedje JM. 1996. Initial characterization of a reductive dehalogenase from Desulfitobacterium chlororespirans Co23. Appl Environ Microbiol 62:3809–3813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Yoshinari T, Hynes R, Knowles R. 1977. Acetylene inhibition of nitrous oxide reduction and measurement of denitrification and nitrogen fixation in soil. Soil Biol Biochem 9:177–183. doi: 10.1016/0038-0717(77)90072-4. [DOI] [Google Scholar]
  • 53.Stookey LL. 1970. Ferrozine—a new spectrophotometric reagent for iron. Anal Chem 42:779–781. doi: 10.1021/ac60289a016. [DOI] [Google Scholar]
  • 54.Higgins SA, Welsh A, Orellana LH, Konstantinidis KT, Chee-Sanford JC, Sanford RA, Schadt CW, Löffler FE. 2016. Detection and diversity of fungal nitric oxide reductase genes (p450nor) in agricultural soils. Appl Environ Microbiol 82:2919–2928. doi: 10.1128/AEM.00243-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Wilhelm E, Battino R, Wilcock RJ. 1977. Low-pressure solubility of gases in liquid water. Chem Rev 77:219–262. doi: 10.1021/cr60306a003. [DOI] [Google Scholar]
  • 56.Im J, Lee J, Löffler FE. 2013. Interference of ferric ions with ferrous iron quantification using the ferrozine assay. J Microbiol Methods 95:366–367. doi: 10.1016/j.mimet.2013.10.005. [DOI] [PubMed] [Google Scholar]
  • 57.Fish JA, Chai B, Wang Q, Sun Y, Brown CT, Tiedje JM, Cole JR. 2013. FunGene: the functional gene pipeline and repository. Front Microbiol 4:291. doi: 10.3389/fmicb.2013.00291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, Buxton S, Cooper A, Markowitz S, Duran C, Thierer T, Ashton B, Meintjes P, Drummond A. 2012. Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 28:1647–1649. doi: 10.1093/bioinformatics/bts199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Marchler-Bauer A, Panchenko AR, Shoemaker BA, Thiessen PA, Geer LY, Bryant SH. 2002. CDD: a database of conserved domain alignments with links to domain three-dimensional structure. Nucleic Acids Res 30:281–283. doi: 10.1093/nar/30.1.281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Yilmaz P, Parfrey LW, Yarza P, Gerken J, Pruesse E, Quast C, Schweer T, Peplies J, Ludwig W, Glöckner FO. 2014. The SILVA and All-Species Living Tree Project (LTP) taxonomic frameworks. Nucleic Acids Res 42:D643–D648. doi: 10.1093/nar/gkt1209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Darriba D, Taboada GL, Doallo R, Posada D. 2012. jModelTest 2: more models, new heuristics and high-performance computing. Nat Methods 9:1–4. doi: 10.1038/nchembio.1155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Guindon S, Dufayard J-F, Lefort V, Anisimova M, Hordijk W, Gascuel O. 2010. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol 59:307–321. doi: 10.1093/sysbio/syq010. [DOI] [PubMed] [Google Scholar]
  • 63.Tavaré S. 1986. Some probabilistic and statistical problems in the analysis of DNA sequences, p 57–86. In Miura RM. (ed), Lectures on mathematics in the life sciences, 17th ed American Mathematical Society, Providence, RI. [Google Scholar]
  • 64.Wagner DD, Hug LA, Hatt JK, Spitzmiller MR, Padilla-Crespo E, Ritalahti KM, Edwards EA, Konstantinidis KT, Löffler FE. 2012. Genomic determinants of organohalide-respiration in Geobacter lovleyi, an unusual member of the Geobacteraceae. BMC Genomics 13:200. doi: 10.1186/1471-2164-13-200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Melton ED, Stief P, Behrens S, Kappler A, Schmidt C. 2014. High spatial resolution of distribution and interconnections between Fe- and N-redox processes in profundal lake sediments. Environ Microbiol 16:3287–3303. doi: 10.1111/1462-2920.12566. [DOI] [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

Articles from Applied and Environmental Microbiology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES