Skip to main content
Applied and Environmental Microbiology logoLink to Applied and Environmental Microbiology
. 2012 Nov;78(21):7501–7510. doi: 10.1128/AEM.01960-12

Diversity, Physiology, and Niche Differentiation of Ammonia-Oxidizing Archaea

Roland Hatzenpichler 1,
PMCID: PMC3485721  PMID: 22923400

Abstract

Nitrification, the aerobic oxidation of ammonia to nitrate via nitrite, has been suggested to have been a central part of the global biogeochemical nitrogen cycle since the oxygenation of Earth. The cultivation of several ammonia-oxidizing archaea (AOA) as well as the discovery that archaeal ammonia monooxygenase (amo)-like gene sequences are nearly ubiquitously distributed in the environment and outnumber their bacterial counterparts in many habitats fundamentally revised our understanding of nitrification. Surprising insights into the physiological distinctiveness of AOA are mirrored by the recognition of the phylogenetic uniqueness of these microbes, which fall within a novel archaeal phylum now known as Thaumarchaeota. The relative importance of AOA in nitrification, compared to ammonia-oxidizing bacteria (AOB), is still under debate. This minireview provides a synopsis of our current knowledge of the diversity and physiology of AOA, the factors controlling their ecology, and their role in carbon cycling as well as their potential involvement in the production of the greenhouse gas nitrous oxide. It emphasizes the importance of activity-based analyses in AOA studies and formulates priorities for future research.

INTRODUCTION

The discovery of archaeal ammonia oxidizers has radically challenged our bacteriocentric view of nitrification and stands as an example of the fascinating complexity of microbes involved in biogeochemical cycling. For over a century, ammonia-oxidizing bacteria (AOB) have been known to catalyze the first step of nitrification, the aerobic oxidation of ammonia (NH3) to nitrite (NO2). Their metabolic hallmark, ammonia monooxygenase (AMO), the enzyme catalyzing the initial oxidation of NH3, has become one of the most widely used molecular markers in environmental microbiology. The first indication of the involvement of archaea in ammonia oxidation (AO) came when amo-like genes associated with archaeal scaffolds were discovered in marine surface waters (137) and soil (118, 132).

DIVERSITY OF AMMONIA-OXIDIZING ARCHAEA

Shortly after these insights from metagenomics, the isolation of the first ammonia-oxidizing archaeon (AOA), Nitrosopumilus maritimus SMC1 of the until then enigmatic group I.1a archaea, was reported (65). Members of this lineage are ubiquitously distributed in open ocean and coastal waters and have been demonstrated to represent 20% to 30% of marine microbes (48, 59, 82, 144). Within the past few years, additional N. maritimus strains have been obtained in enrichment cultures (99, 144). Furthermore, the uncultivated marine sponge symbiont “Candidatus Cenarchaeum symbiosum” was shown to encode genes essential for the oxidation of NH3 and thus became regarded as an AOA (44, 45, 107). However, this organism has not yet been shown to catalyze the oxidation of NH3 and, until further data are available, should be considered an amoA-encoding archaeon (AEA; 25). Later, the group I.1a AOA “Ca. Nitrosoarchaeum limnia” SFB1 was enriched from a low-salinity sediment and its genome was nearly completely reconstructed via a combination of metagenomics and single-cell sequencing (15). The genomes of “Ca. Nitrosoarchaeum koreensis” and “Ca. Nitrosopumilus salaria” were obtained from enrichment cultures from agricultural soil and estuary sediment, respectively (58, 60, 88). Very recently, novel (as-yet-unnamed) AOA species were enriched from freshwater sediment (36), expanding our knowledge about the environmental distribution of this archaeal lineage. Besides these group I.1a archaea, two thermophilic AOA species, “Ca. Nitrososphaera gargensis” (46) and “Ca. Nitrosocaldus yellowstonii” (29), have been described. While the former was the first characterized member of group I.1b archaea, the latter represents a deep-branching lineage (thermophilic AOA [ThAOA] group; formerly hot water crenarchaeotal group III [HWCG-III]) with wide distribution in high-temperature habitats.

One of the milestones of research on mesophilic archaea was the isolation of the first representative of soil-inhabiting AOA, Nitrososphaera viennensis EN76 (131). This group I.1b archaeon could serve as a model organism for future studies, most importantly because it represents a population of archaea exhibiting global distribution in soils (42, 61, 72, 103). Until 2011, archaeal ammonia oxidizers capable of growth at low pH had not been cultivated. Thus, the discovery of the first obligately acidophilic AOA “Ca. Nitrosotalea devanaterra” of the group I.1a-associated lineage is of particular importance because acidic soils (pH < 5.5) can exhibit high nitrification rates and comprise ∼30% of the ice-free terrestrial surface (see reference 71 and references therein).

Table S1 in the supplemental material lists important features of characterized AOA species, and their phylogeny as determined on the basis of AmoA sequences is shown in Fig. S1 in the supplemental material. Besides the known lineages of AOA (group I.1a, group I.1a-associated, group I.1b, ThAOA), sequence data suggest that more, as-yet-unidentified amoA-encoding and potentially ammonia-oxidizing groups might exist (84, 102, 103, 108).

THAUMARCHAEOTA: BEYOND AMMONIA OXIDATION

In 2008, analysis of the “Ca. Cenarchaeum symbiosum” genome led to the proposal of a novel phylum of archaea, Thaumarchaeota, which was later confirmed when more genome sequences became available (19, 124). The idea of the uniqueness of Thaumarchaeota is supported by the phylum's distinct phylogenetic position, a characteristic set of archaeal marker genes, and specific insertion and deletion sites in certain tRNA synthetases and proteins engaged in ribosomal functioning and cell division, as well as by the presence of a phylum-specific membrane lipid, thaumarchaeol (see references 19, 102, and 124 and references therein).

Presently, chemolithoautotrophic growth on NH3 is the only physiology known for members of the Thaumarchaeota. It is, however, highly probable that their true metabolic diversity has as yet escaped our attention. Recently, it was reported that in some wastewater treatment plant sludges (wwtps), thaumarchaeotes that express putative amo genes (for their phylogenetic context, see Fig. S1 in the supplemental material) were highly enriched. Outnumbering AOB up to 10,000-fold, their cell numbers were 100 to 1,000 times higher than could be sustained by autotrophic AO alone. In highly dynamic systems such as wwtps, these high cell numbers can be explained only by active growth, suggesting the use of unknown heterotrophic substrates to conserve energy (92). This observation serves as an important reminder that for complex samples, the detection and enumeration of gene sequences are insufficient to propose a physiology. In addition, the discovery of giant thaumarchaeotes that may be involved in sulfur cycling in mangrove swamps demonstrated our lack of knowledge on the biology of members of this phylum (91). Despite its close relatedness to N. maritimus on the 16S rRNA sequence level (>97.7% identity), Giganthauma exhibits remarkable differences from this planktonic, sub-micrometer-sized AOA. If these data can be confirmed and extended in future experiments, this newly discovered microbe has the potential to shed new light on this still largely enigmatic archaeal phylum (91).

PHYSIOLOGY

AMO.

A key enzyme in the biochemistry of AOA and AOB is ammonia monooxygenase (AMO). AMO is a member of the AMO/pMMO/pBMO/pXMO enzyme group, a diverse family of copper-containing membrane-associated monoxygenases (CuMMOs) that engage in the oxidation of ammonia (AMO), methane (pMMO), butane (pBMO), and possibly propane and ethane, as well as other as-yet-unidentified substrates (pXMO) (see reference 129 and references therein). CuMMOs are heterotrimers and in bacteria are commonly genetically encoded in the subunit order CAB, while in AOA the arrangement of these genes differs between different lineages (6, 15, 60, 124, 140). In ammonia oxidizers, the genes amoA, amoB, and amoC encode the three subunits of AMO. Whereas multiple, nearly identical amoCAB copies occur in AOB (3), no AOA has yet been found to carry more than a single copy of amoA or amoB, while two amoC copies are regularly found (15, 45, 140). Based on data obtained from the soluble fraction of its homologue PmoB, AmoB is thought to harbor the active site of AMO (4, 74). In AOB, AmoC is thought to exhibit chaperonic activity with respect to the other subunits to assist in their integration into the membrane (62) and has been shown to be involved in recovery from NH3 starvation (13, 14). Unfortunately, AMO has so far defied all attempts for functional isolation and structural analyses. Thus, an unambiguous assignment of the roles of individual Amo proteins cannot yet be drawn.

With the exception of fosmid 19c08, all currently known genomes or genome fragments of AEA/AOA encode a conserved hypothetical protein directly upstream of amoA (6). Interestingly, a recent metatranscriptomic study demonstrated that homologues of this gene are among the most highly expressed mRNAs in estuarine bacterioplankton, amouting to ∼70% and ∼296% of amoA and amoB transcript numbers, respectively (49). It was proposed that this gene may be associated with the Amo proteins and referred to as “amoX” (6). In the same publication, the first evidence for the potential transposon-linked mobilization of archaeal amo genes was presented, which might also provide an explanation for the disconnection of amoA and “amoX” in fosmid 19c08 (6).

Proposed pathways.

In AOB, the membrane-associated AMO catalyzes the aerobic oxidation of NH3 to hydroxylamine (NH2OH) which is subsequently oxidized to NO2 by the periplasmic hydroxylamine oxidoreductase (HAO) (16). The downstream electron (e) flow has not yet been fully resolved (see references 64 and 123 and references therein). Under standard conditions (pH 7.0, 25°C), ∼0.7% of the total pool of ammonia plus ammonium (NH4+) is available as NH3. While it is generally accepted that NH3 and not NH4+ is the substrate for bacterial AMO (128), archaeal AMO has never been tested in that regard. It is unclear whether archaeal AMO catalyzes the same reaction as its bacterial counterpart, because no hao homologue, enzymes for the detoxification of NH2OH, or cytochrome c has been found in any AOA genome (see, e.g., references 123 and 140; for a list of sequenced AOA, see Table S1 in the supplemental material). In addition, the observation that archaeal AmoB and bacterial AmoB have significantly different predicted structures might indicate a different function of this protein in AOA (140). Thus, either an unidentified enzyme substitutes for HAO in AOA or the archaeal AMO reaction yields a different product (15, 45, 60, 124, 140). It was suggested that nitroxyl hydride (HNO) may be generated by a monooxygenase reaction of the archaeal AMO. Subsequently, HNO could be oxidized to NO2 via a nitroxyl oxidoreductase (NxOR) (119, 140). The activation of O2 for the monooxygenase reaction could also be achieved by nitric oxide (NO), the reaction product of nitrite reductase (NIR), which would result in the production of N2 gas (119). It was reported that archaeal nirK (encoding copper-dependent NIR) genes are expressed under aerobic conditions (see, e.g., references 5, 49, 58, 78, and 132), suggesting a different function of these enzymes in AOA compared to their bacterial counterparts. nirK homologues are present in all published AOA genomes but are absent in the genome of “Ca. Cenarchaeum symbiosum.” Thus, if this microbe is indeed an AOA, then the model described above might not be fully valid.

In all of the discussed physiological pathways, two of the released e participate in the generation of a proton motive force and lead to the conservation of energy via an ATPase. For both AOA and AOB, the exact amount of synthesized ATP per mole of NH3 is unknown, because assessment of the number of e available for energy conservation is dependent on the growth stage and the extent of reverse e flow, as well as other factors, and on the fact that their e transport process is yet not fully understood (16, 140). However, the lack of cytochrome c proteins and the existence of numerous genes encoding copper-containing proteins (multicopper oxidases and plastocyanin-like domain proteins) in AOA suggest a different e transport mechanism (15, 45, 140) from that of the highly iron-heme-dependent AOB (38, 39). A copper-based biochemistry would help to explain the ecological success of marine AOA (compared to AOB), because dissolved copper concentrations are generally an order of magnitude higher than those of iron in seawater (see reference 134 and references therein).

Specificity.

Bacterial AMO does not exhibit high substrate specificity: methane (CH4) and carbon monoxide (CO) as well as some aliphatic and aromatic compounds may be oxidized and act as competitive inhibitors of AMO (see references 51 and 102 and references therein) (109). These substrates, however, cannot be used for energy conservation and do not support the growth of AOB (16, 51). This low specificity is regarded as a reflection of the evolutionary history of AMO as a member of the CuMMO enzyme family (63). Given their substrate promiscuity, it has been hypothesized that not the type of CuMMO but the downstream biochemical machinery defines an organism's substrate specificity (129). The substrate spectrum of archaeal CuMMOs remains largely unknown, and growth of N. maritimus SCM1 on CH4 or other organic or inorganic e donors has not been observed (134). However, given the substrate promiscuity of known CuMMOs as well as the high abundance of AEA in the environment, the existence of other (potentially energy-yielding) substrates of archaeal CuMMOs seems plausible.

Inhibitors.

In contrast to AOB data (7, 16, 51), only limited information on potential inhibitors is available for AOA. While the metal chelator allylthiourea is observed to lead to a decrease, but not complete inhibition, of metabolic activity (46, 58, 61, 90), acetylene has been demonstrated to cause a total suppression of AO in cultivated AOA (71, 131) as well as in situ (56, 96, 97). Contrasting observations have been reported for dicyandiamide (30, 31, 58, 61, 151), and nitrapyrin has been successfully tested only on “Ca. Nitrosoarchaeum koreensis” and “Ca. Nitrososphaera sp.” JG1 (58, 61). Use of the antibiotic sulfadiazine offers the ability to differentiate between AOA- and AOB-catalyzed AO via the preferential inhibition of bacteria (117). In addition, both AOA (36, 83) and AOB (43, 52) exhibit species- and wavelength-specific photoinhibition, which in AOB is attributed to photo-oxidation damage of AMO (50, 53). Different light sensitivities of AOA and AOB support the idea of a potential mechanistic difference between their respective versions of AMO (36, 83).

EMISSION OF NITROUS OXIDE

As a result of classical and nitrifier denitrification—the reduction of NO2/NO3 in several steps to N2—nitrous oxide (N2O) is produced and partially escapes into the gas phase (18, 64). With its long atmospheric lifetime, N2O has a greenhouse warming potential 310 times higher than carbon dioxide (CO2) and is responsible for 5% to 7% of the observed greenhouse effect, making it the third most important greenhouse gas (after CO2 and CH4) (18, 55). Due to its reaction with atomic oxygen, N2O also has detrimental effects on Earth's ozone (O3) layer (110). Together with canonical denitrifiers, AOB (64, 75) and, potentially, AOA (58, 76, 90, 115, 116) are suspected to be responsible for ∼70% of the global N2O emissions (55).

Potential sources.

In AOB, two sources of this gas can be distinguished. Small amounts are produced via the chemical decomposition of NH2OH to NO and N2O (18). In addition, under O2-limiting conditions, AOB readily use NO2 as a terminal e acceptor, leading to the formation of nitrogen-containing gases via the activity of NIR and nitric oxide reductase (NOR) enzymes (18). For AOA, the enzymatic production of N2O has yet not been demonstrated, and observed N2O emissions may be due to spontaneous chemical reactions of metabolic intermediates. A NOR enzyme has not yet been identified in any fully sequenced AOA (for a list, see Table S1 in the supplemental material), and NH2OH is currently not regarded as an intermediate in their AO pathway. While the reactions leading to the formation of N2O apparently differ between AOA and AOB (76), 18O labeling studies suggest the same ratio of oxygen sources in NO2 (one atom originating from H2O and one from O2) (115). Moreover, the site preference for 15NN2O produced by enrichments (115) and pure cultures (76) of AOA is consistent with the values for AOB cultures seen under AO conditions. This suggests that in AOA (at least under the tested conditions), N2O originates from an intermediate in the AO process and not AOB-like nitrifier denitrification (76). Assuming that HNO is indeed an intermediate in the AOA AO pathway, it might be the source of archaeal N2O emission (M. G. Klotz, personal communication). In water, HNO molecules dimerize to form hyponitrous acid (H2N2O2), which is subsequently dehydrated to N2O—a process first proposed for denitrifying bacteria (see references 85 and 153 and references therein).

Yield.

While the amount of N2O emitted from an enrichment culture containing “Ca. Nitrosoarchaeum koreensis” is considerably lower than that seen with Nitrosomonas europaea (58), N2O production rates up to five times higher were reported for N. maritimus SCM1 versus Nitrosococcus oceani and Nitrosomonas marina (76). An explanation for this apparently conflicting result may be an overestimation of rates resulting from the use of high-density cultures that may not be representative of the environment (76). Other explanations may be that soil and marine ammonia oxidizers differ in their (relative) responses to low O2 conditions or the potential presence of N2O-removing microbes in the enrichment culture (58), leading to an underestimation of the rates. In N. maritimus SCM1 cultures, N2O emission inversely correlates with the O2 concentration, an observation that is in accordance with the idea that the gross amount of N2O generated in analyzed marine oxygen minimum zones (OMZ) potentially originates from group I.1a archaea (76).

NICHE DIFFERENTIATION

The high numbers of AEA/AOA in many natural and man-made systems raise questions on the exact nature of their cellular and biochemical adaptations that lead to this wide distribution, as well as on their relative significance in nitrification. This section discusses the most important factors known to shape the ecology of AOA compared to AOB. While it gives examples of the niche differentiation of specific AOA populations, this discussion is not intended to present an exhaustive list of habitats in which amoA-like gene sequences have been detected, and it emphasizes the importance of activity-based analyses in environmental studies.

Oligotrophy.

One of the principal factors controlling the relative distribution of AOA versus AOB is the substrate concentration. In contrast to several analyzed AOB, N. maritimus SCM1 has been demonstrated to exhibit exceptionally high affinities (Km = 133 nM total NH4+) and low (≤10 nM) thresholds with respect to its substrate (79, 80). Most importantly, these values correlate well with the in situ kinetics of the oligotrophic open oceans that have been shown to contain large amounts of AEA/AOA (see reference 79 and references therein). If most marine N. maritimus-related populations were indeed AOA, they would be responsible for the bulk of AO observed in these habitats, which would call for a reinvestigation of current biogeochemical models (79). A preference for low NH3 levels had initially been reported for “Ca. Nitrososphaera gargensis” (46) and was recently also demonstrated for “Ca. Nitrosotalea devanaterra” (71). In addition, three sedimentary/freshwater AOA enrichments were shown to extend their lag phases with increasing concentrations of NH3 (36; see also Table S1 in the supplemental material). This is in contrast to the soil AOA N. viennensis EN76 and “Ca. Nitrosoarchaeum koreensis,” which were found to be adapted to considerably higher substrate concentrations (58, 131). While the inhibitory NH4+ concentrations (10 to 20 mM) of these AOA (see Table S1 in the supplemental material) are comparable to that of the most oligotrophic AOB known (21.4 mM for N. oligotropha JL21), they are still low compared to the highest NH4+ tolerance of AOB (50 to 1,000 mM) (66). A preference of AEA/AOA for low NH4+ concentrations has also been reported from several environmental studies, especially for soils (see, e.g., references 30, 106, and 138). This might be explained by either higher affinities or higher densities of archaeal AMO or transporters for NH3/NH4+ compared to AOB (134). However, it is still unclear what the exact localization of the catalytic site of AMO is (i.e., whether it faces the periplasm or cytoplasm), whether the mentioned transporters are used for the accumulation of NH3, or whether archaeal AO is transporter dependent (134). Interestingly, reads assigned to group I.1a archaeal NH3 transporters and permeases are among the most commonly detected genes and transcripts in marine samples (49, 120, 125, 133).

Cell (ultra)structure.

For N. maritimus SCM1, the presence of a single genome copy per cell as well as a very low rate of replication (15 to 18 h for a 1.65-Mb genome) supports the idea of the organism's adaptation to oligotrophic habitats (101). According to cryoelectron tomography data, N. maritimus SCM1 cells in exponential growth harbor ∼1,000 ribosomes per ∼0.023-μm3 cell volume (134). Data on the stability of mRNAs, AMO, and ribosomal proteins of AOA are lacking but could be essential in understanding the ecological adaptations of AOA compared to AOB, which are known for the exceptionally high in vivo stabilities of their amo mRNA and proteins as well as the high level of their ribosomal contents (16, 66). While energetically expensive, consistently high numbers of ribosomes offer organisms the ability to respond quickly to changing environmental conditions, for example, fluctuating NH3 levels. The observations described above are consistent with the hypothesis that most archaea, in contrast to bacteria, are highly adapted to energy-stressed environments (135). The lower membrane permeability of AOA cells, a direct consequence of their preference of tetraether lipids, has been proposed to result in a reduction in ion cycling and, thus, lower levels of maintenance energy relative to AOB (135).

Furthermore, most AOA have cell volumes that are 10 to 100 times smaller than those of known AOB (for cell sizes and references, see Table S1 in the supplemental material). This has profound implications for their per-cell AO rates, which for N. maritimus SCM1 were reported to be ∼10-fold lower (0.53 fmol NH3 cell−1 h−1 at highest activity) than those of AOB (see references 79 and 134 and references therein). Thus, the high relative abundance of AOA compared to AOB in many environments does not per se implicate a major importance in net nitrification, necessitating the use of activity-correlated analyses that can differentiate between their individual contributions.

NH3 source.

Besides the substrate concentration, the form in which NH3 is supplied also governs niche adaptation. Due to the low nitrogen uptake efficiency of crops and leaching of NO2/NO3, ∼70% of fertilizer nitrogen is lost to the atmosphere or washed out from soils before assimilation into biomass can occur (see reference 21 and references therein). The preference of AOA versus AOB for different NH3 sources thus could have important consequences for agricultural fertilization strategies that aim to minimize nitrogen loss and maximize crop production. In all studies that reported growth of soil AOA, NH3 originated from mineralized organic material (31, 73, 96, 117, 150). On the other hand, when NH4+ fertilizer or urea (NH2CONH2) was provided as the source of substrate, AOB usually strongly outcompeted AOA (see, e.g., references 30 and 56). Also, the total nitrogen and organic carbon content of a range of different soils has been shown to negatively correlate with the species richness of putative AOA (103).

pH.

The observation of the preference of AOA for low NH3 concentrations is in accordance with studies that reported high numbers and activities of AEA/AOA in acidic soils (41, 42, 77, 93). A recent study demonstrated that, among seven physicochemical parameters measured (pH; carbon, nitrogen, and organic matter content; C:N ratio; soil moisture; and vegetation), the pH value was the major factor governing AEA community structure (42). Low-pH conditions decrease the availability of NH3 while increasing the toxicity of NO and N2O as well as gaseous nitrogen dioxide (NO2) (26, 27). While some AOB populations are adapted to coping with low pH, cultured representatives show no or only very limited activity at pH < 6.5 (see references 41, 67, and 93 and references therein). Many but not all species of AOB (105) and some AEA/AOA (45, 131, 133, 147) encode ureases, enzymes that catalyze the conversion of urea to CO2 and NH3. The reaction products can then be used as sources of carbon and energy, respectively, or potentially to regulate the pH in the vicinity of the cell. Compared to AOB, AEA/AOA are more transcriptionally active in acidic soils and both microbial groups harbor phylotypes that are specifically adapted to low-pH conditions (42, 93, 103). In a study targeting amoA-like gene diversity in a wide range of globally distributed soils, several lineages within group I.1b and the I.1a-associated group clearly exhibited adaptation to certain pH regimens, and these results were coherent at the global, regional, and local level of sampling sites. Most prominently, a strong correlation of the two sublineages of group I.1a-associated amoA-like sequences with acidic soils (pH < 5) was found (42). These findings are in accordance with the observation that in some acidic soils (pH 3.75 and 5.4) in which AOB could not be detected, AOA closely related to Nitrosotalea and Nitrososphaera grew when urea was provided as a substrate (77). Archaeal amoA-like gene abundance and diversity directly increase with soil pH (42, 103), and so far only one obligately acidophilic AOA has been described (71). However, the wide distribution of the Nitrososphaera cluster and other amoA lineages in acidic soils suggests the existence of other low-pH-adapted AOA (42, 103, 145).

Differential activities of soil AOA populations.

AEA/AOB amoA ratios increase with soil depth (31, 47, 56, 72, 93), and specific AEA phylotypes exist in different soil horizons, suggestive of populational adaptations to such microenvironments (72, 94). Only rarely, however, were such studies combined with activity tests (see, e.g., references 56, 106, 145, and 150), and available data are partially conflicting, which makes it hard to judge the relative importance of AOA in soil nitrification (aside from acidic habitats). A preferred tool in comparative soil studies has been stable isotope labeling (SIP) of nucleic acids. Using RNA SIP, it was demonstrated that, although AOA incorporate 13CO2-derived carbon into their amoA and carbon fixation gene transcripts (i.e., are metabolically active), their growth (i.e., replication of DNA) cannot be detected via DNA SIP (56, 106). This is in contrast to other studies that reported growth of AOA even after relatively short incubation times (4 compared to up to 12 weeks in other studies) (145, 150). Irrespective of their high numerical dominance in most soil habitats (42, 72, 103), only a limited number of studies reported the activity of Nitrosotalea- and Nitrososphaera-related archaea (77, 145). In contrast, several investigations observed the preferential activity and growth of the low-abundance group I.1a AOA (96, 130, 138, 150). These results are a warning that we still do not appreciate the functional diversity of the numerically dominant group of soil archaea (i.e., group I.1b). In addition, many soil AEA actually may not (always) be directly involved in AO because of either a potential for mixotrophic growth (see next chapter) or additional or different functions of their CuMMOs.

Elevated temperatures.

Specific soil-inhabiting AEA subpopulations (mostly group I.1a related), but not AOB, increase expression of their amoA-like genes and probably nitrify upon incubation at >30°C (96, 130). Due to the limited number of studies, we do not yet know whether this differential temperature response is a general trend for soil ammonia oxidizer communities. Interestingly, however, AOB have so far not been detected in environments experiencing constant temperatures of >40°C. Archaeal amoA-like genes (33, 111, 141, 149) and transcripts (57, 149), on the other hand, have been retrieved from geothermal habitats at up to 97°C. Many retrieved sequences are highly similar to those of amoA genes of the cultured thermophilic AOA “Ca. Nitrosocaldus yellowstonii” and “Ca. Nitrososphaera gargensis” (29, 46), but several other species still await characterization (57, 111, 149). Thermodynamic calculations of potential chemolithotrophic reactions in two geothermal springs demonstrated that AO is among the highest energy-yielding physiologies under such conditions (34). Consistently, in situ measurements of AO in several hot springs have indicated the essential role of heat-adapted AOA in these systems (33, 111). Recent data support the idea of a thermophilic ancestor of Thaumarchaeota and a geothermal origin of archaeal AO (20, 29, 40, 46, 95). According to this hypothesis, today's nearly ubiquitous mesophilic AEA/AOA are the result of secondary adaptations to the lower temperatures of terrestrial and marine systems (20, 29, 46).

Zonation with water depth.

Many studies have addressed the distribution of amoA-like gene sequences in the marine water column, but only a limited number have characterized the in situ activity of AOA (see, e.g., references 2, 9, 114, 125, and 147). Soon after the discovery of AOA, it was recognized that archaeal amoA-like gene numbers correlate with NO2/NO3 maxima in the oceans (see, e.g., references 9, 24, and 84). Besides the widely distributed N. maritimus-like sequences, many sequences that have been obtained from marine samples fall within two phylogenetically distinct clusters. Group A or “shallow” genes and transcripts are primarily derived from the shallow, euphotic zone (<200 m depth), while group B or “deep clade” sequences represent deep water (>200 m depth) ecotypes (see, e.g., references 9, 23, 84, 87, and 114). A similar zonation has been observed for genes encoding putative thaumarchaeotal carbon-fixing 4-hydroxybutyryl-coenzyme A (CoA) dehydratase (Hcd) and acetyl-CoA carboxylase (AccA/PccB), as well as nirK and urease enzymes (78, 147). In the photic zone, such patterns could be due to either differential photoinhibition or variable success of AOA populations in competition with phytoplankton for NH3 (9, 84). While these ideas have not been directly tested so far, it was recently reported that the sensitivity to light for three AOA species was greater than that for AOB (36, 83). While most studies recorded a strong correlation of AO with 16S rRNA and amoA-like gene sequences of “shallow” group A (9, 23), trends are not always clear for deep-water thaumarchaeotes, partly due to the presence and activity of AOB in the same waters (2, 114). In contrast to most other reports, two studies found stark discrepancies between thaumarchaeotal 16S rRNA and amoA-like gene copy numbers (2, 28), but these conflicting results are at least partly due to primer biases (114, 147).

Oxygen deprivation.

An intriguing feature of some marine AOA is their apparent preference for regions of low (<10 μM) levels of dissolved O2, where AO might be coupled to anaerobic ammonium oxidation (anammox) and/or denitrification (9, 69, 70, 86, 104, 125). It was thus suggested that AOA and AOB together provide 30% to 40% of the NO2 required by anammox in the OMZ of the Black Sea (69) and off Peru (70), but no indication for such coupling was found in the Arabian Sea (104). These results largely coincide with observations of pure and enrichment cultures which demonstrate a much higher affinity of N. maritimus SMC1 to O2 (79, 80) and likely also of other group I.1a AOA (58, 99) compared to AOB. In addition, three AOA enrichments that had been obtained from freshwater sediment were recently demonstrated to be active under conditions of reduced O2 concentrations (0.5% to 2% O2 in the headspace). The observation that other microbes, such as Escherichia coli, are able to grow aerobically at O2 concentrations ≤ 3 nM (126) raises the possibility that microbes with other physiologies could also be adapted to such minute O2 concentrations. It must be kept in mind that the nitrification rates observed in OMZ (20 to 150 nmol liter−1 day−1) are well within the range of the prevalent O2 concentrations (69, 70) and that the O2 levels required to sustain such rates are below the detection limit of currently used in-field O2 sensors (∼1 μM). In either case, AOA are expected to have evolved molecular adaptations to cope with periods of O2 deprivation, as known for AOB performing microaerophilic nitrification (37).

Investigations of the adaptivity and tolerance of AOA with respect to low-O2 regimens are of utmost importance, given the high abundance of archaeal amoA-like gene sequences in other habitats such as sediments, seasonally oxygen-deficient water zones, or certain agricultural soils (see, e.g., references 1, 10, 22, 68, 86, and 99). Unfortunately, the contribution of AOA to the in situ nitrification as well as N2O emission rates in these systems is largely unknown.

Other factors.

Additional variables that might be involved in shaping the ecology of AEA/AOA have been brought into discussion, but data are yet not conclusive. Some potential factors are the levels of sulfide, phosphate, salinity, soil moisture, and others (see references 17, 35, and 148 and references therein).

CARBON SOURCES

Autotrophy.

Although both AOA and AOB are usually regarded as autotrophic organisms, important differences exist in how they fix inorganic carbon and use organic carbon. While AOB rely on the Calvin-Benson-Bassham (Calvin) cycle for carbon fixation (3), a modified version of the 3-hydroxypropionate/4-hydroxybutyrate (3HP/4HB) cycle seems to operate in AOA (11, 15, 99, 131, 140). In addition, for N. maritimus SCM1 and “Ca. Cenarchaeum symbiosum” as well as for marine planktonic thaumarchaeotes, a reverse tricarboxylic acid (rTCA) cycle for carbon fixation has been suggested (45, 81, 140). The preference for different carbon-fixing pathways has important consequences for the ecological adaptation of microbes, as recently reviewed (11). Most importantly, in the 3HP/4HB cycle, bicarbonate (HCO3) is the fixed carbon species, while the Calvin cycle fixes CO2. This is of high importance, considering that at neutral and slightly alkaline pH, as found in marine waters, HCO3 is the predominant carbon species. In addition, a complete or even rudimentary 3HP/4HB cycle provides the organism the capability to coassimilate many different organic compounds, including fermentation products. For marine archaea, the ability to use 3HP, an intermediate in the metabolism of the ubiquitously distributed osmoprotectant dimethylsulfoniopropionate (DMSP), may be ecologically more important (11). It raises the question of whether these microbes are directly involved in the breakdown of DMSP or its products, e.g., acrylate, and whether this could potentially account for the observed heterotrophy and mixotrophy of marine thaumarchaeotal populations (2, 54, 136).

Mixotrophy?

A potential for mixotrophic growth has been reported from analyses of the genomes of N. maritimus SCM1 and “Ca. Cenarchaeum symbiosum” (44, 45, 140), but these hypotheses have not yet been experimentally supported. In contrast, the addition of organic substances (yeast extract, peptone, and acetate) has been shown to inhibit AO by N. maritimus SCM1 (65) and “Ca. Nitrosocaldus yellowstonii” (29). However, the tested compounds may not be representative of the substrates present in their respective niches. N. maritimus SCM1 and “Ca. Cenarchaeum symbiosum” harbor transporters for different amino acids, di- and oligopeptides, glycerol, and sulfonates or taurine, making these compounds ideal candidates for future experiments (44, 45, 140). A metatranscriptomic study did not report any expression of these transporters in coastal bacterioplankton, while transcripts assigned to the 3HP/4HB and tricarboxylic acid cycle were detected (49). However, the genomic analyses described above are consistent with studies that demonstrate the genetic potential (133) and metabolic ability of marine thaumarchaeotal populations to use amino acids (48, 98) or other organic matter (2, 54, 136). Interestingly, cultures of the soil AOA N. viennensis EN76 exhibit cell yields that are 12 times higher for growth under mixotrophic (pyruvate) rather than purely autotrophic conditions (131). However, less than 10% of cellular carbon seems to be derived from this substrate under the tested conditions (131). While key components of the 3HP/4HB cycle could be identified in the draft genome of N. viennensis EN76, no indications of its dependence on pyruvate compared to that of N. maritimus SCM1 could be found (131). Very recently, SIP experiments demonstrated the incorporation of 13C from labeled rice callus into the DNA of both AOA and AOB (142). While potential cross-feeding—via respiration of labeled substrates by other microbes and subsequent fixation of exhaled 13CO2—cannot be ruled out completely, the results of this study correspond to those of other reports of root-colonizing group I.1b thaumarchaeotes (121, 122, 146).

OUTLOOK

PCR-based studies.

Based on the pairwise comparison of thaumarchaeotal 16S rRNA and amoA genes from metagenomic fragments and cultured AEA/AOA, archaeal amoA-like sequences deposited in public databases as of June 2010 represented ∼113 to 120 AEA/AOA species (i.e., amoA identities < 85%; 103). This range of values might be a conservative estimate, given that the primers that have most frequently been used in past research do not effectively replicate the full diversity of the amoA-like genes now known (103). Primer sets which together target the whole range of archaeal amoA-like sequence diversity were recently published (32, 103, 130), and their use in future studies is recommended (primers are listed in Table S2 in the supplemental material). In the future, PCR-based analyses of AOA populations are expected to be expanded to genes encoding enzymes involved in pathways other than AO. Potential candidates include hcd and accA/pccB (11, 97) as well as nirK (5, 78, 132). Homologues of these genes are present in all genome-sequenced AOA (nirK is, however, absent from the genome of the AEA “Ca. Cenarchaeum symbiosum”), and their expression in enrichment cultures as well as in natural populations has recently been shown (see, e.g., references 5, 49, 58, 78, and 106). Another potential target is the uvrBAC cluster, proposed to encode an enzyme involved in DNA excision repair, which can be found in all published thaumarchaeotal genomes. The single unknown is “Ca. Nitrosoarchaeum limnia,” the genome of which harbors a partial uvrAB cluster at a contig end (6).

Correlation with activity.

While PCR-driven studies are expected to keep on being an essential part in AOA research, as discussed in this review, the detection—or even demonstration of expression—of amo-like genes is an insufficient basis for the proposal that the respective organism is indeed an obligate autotrophic ammonia oxidizer (92). Thus, experiments need to be carefully designed and combined with adequate controls (e.g., inhibitors) as well as tests of activity, most importantly, isotopic labeling techniques (targeting the level of single cells, DNA, RNA, lipids, etc.). With the successful cultivation of several species (see Table S1 in the supplemental material) as well as the more widespread application of combined “meta-omics” approaches—concomitantly targeting the genomic, transcriptomic, proteomic, metabolic, and/or geochemical level—and single-cell-resolving techniques within the last years, we are now able to directly approach fundamental problems in the AOA field.

Some open questions.

Currently, the most urgent issue is that of the biochemistry of archaeal AO and its implications for the niche separation from AOB. The determination of the substrate range of AOA as well as their involvement in the emission of N2O is expected to have profound consequences for our understanding of the impact of archaea on biogeochemical cycles. Substitution experiments performed with isotopically labeled nitrogen compounds as well as inhibition tests scavenging potential reaction intermediates should enable us to decipher AOA physiology. In addition, the use of meta-analyses in the comparative study of several AOA in parallel promises to yield important information on the underlying protein machinery and regulatory factors of AO. Similar analyses could also help us to shed light on potential additional functions of archaeal AMO.

It is surprising that, despite their importance for humans and the high diversity of archaeal amoA-like gene sequences detected in these habitats (see, e.g., references 100, 112, 113, 143, and 152), wwtps as well as ground- and freshwater systems have only very rarely been the focus of AOA-targeted enrichment and activity studies, with notable recent exceptions (36, 92). Thus, the available data make it hard to judge the contribution of AOA to nitrogen cycling in these habitats. However, the study of these environments is of particular importance not only for our understanding of nitrogen cycling but also for the optimization of detoxification and removal of municipal and industrial waste. Furthermore, the potential in situ nitrifying activity of sediment-dwelling thaumarchaeotes deserves more attention, considering the large inputs of anthropogenic nitrogen into many freshwater and estuary systems and the high abundance of AEA in the marine habitats (see, e.g., references 8, 10, 12, 89, and 99).

Another topic currently insufficiently studied is the relationship of AOA with nitrite-oxidizing bacteria (NOB). The mutual dependence of AOB and NOB is a typical textbook example of the metabolic coupling of two functional guilds of microbes. Hence, our lack of knowledge of a potential codependence of AOA and NOB is highly unfortunate. Besides the demonstration of a spatial-temporal co-occurrence of these microbes in marine and soil habitats (84, 114, 145), a deeper understanding of their relationship is so far lacking.

In our search for novel AOA, we should try to more directly address the biology of lineages with assumed ammonia-oxidizing activity, e.g., the pSL12, I.1c, and ALOHA groups (77, 84, 102, 127). This could be achieved by a combination of direct single-cell sorting from environmental samples with (meta)genomic sequencing to determine the genetic potential of these groups. After this screening phase, environmental samples could be incubated in the presence of stable or radioactively labeled substrates. They could then be subjected to activity tests using either molecular tools such as differently targeted SIP techniques or single-cell-resolving methods such as microautoradiography, Raman spectroscopy, or dynamic secondary-ion mass spectrometry (see reference 139 and references therein). Finally, given the potential physiological flexibility of some members of the Thaumarchaeota (2, 54, 56, 91, 92, 98, 136), we should extend our studies on these microbes beyond the single metabolism of ammonia oxidation.

Supplementary Material

Supplemental material

ACKNOWLEDGMENTS

I am thankful to Anne Dekas, Jennifer Glass, Martin Klotz, Shawn McGlynn, Victoria Orphan, and Patricia Tavormina for discussions and proofreading. The input by three anonymous reviewers strongly improved the manuscript. I am deeply indebted to Michael Wagner and Christa Schleper, as well as Holger Daims, Elena Lebedeva, Pierre Offre, Victoria Orphan, Thomas Rattei, Anja Spang, and Eva Spieck, for their support, collaboration, and stimulating discussions. Thanks to Christopher Francis, Annika Mosier, and Jose de la Torre for sharing unpublished data. I apologize to all authors whose work could not be discussed due to page limits.

This work was supported by the California Institute of Technology Division of Geological and Planetary Sciences O. K. Earl Postdoctoral Scholarship in Geobiology and an Erwin Schrödinger Fellowship from the Austrian Science Fund (FWF), J 3162-B20.

Biography

graphic file with name zam9991037910001.jpg

Roland Hatzenpichler studied biology with a focus on microbiology and genetics at the University of Vienna, Austria. In 2011, he obtained his Ph.D. at the Department of Microbial Ecology (University of Vienna) under the supervision of Michael Wagner, working on the identification and characterization of novel ammonia-oxidizing archaea and bacteria. He now is a postdoctoral scholar in geobiology in the group of Victoria Orphan at the California Institute of Technology (Caltech). He is currently a Marie Curie Fellow via the Erwin Schrödinger program of the Austrian Science Fund (FWF) at Caltech. Besides nitrifiers, his research interests include the anaerobic oxidation of methane, the diversity of archaea, and the microbiology of extreme habitats, as well as the development of techniques for in situ activity analyses of microbes on the single-cell and “meta” level.

Footnotes

Published ahead of print 24 August 2012

Supplemental material for this article may be found at http://aem.asm.org/.

REFERENCES

  • 1. Abell GC, et al. 2010. Archaeal ammonia oxidizers and nirS-type denitrifiers dominate sediment nitrifying and denitrifying populations in a subtropical macrotidal estuary. ISME J. 4:286–300 [DOI] [PubMed] [Google Scholar]
  • 2. Agogué H, Brink M, Dinasquet J, Herndl GJ. 2008. Major gradients in putatively nitrifying and non-nitrifying Archaea in the deep North Atlantic. Nature 456:788–791 [DOI] [PubMed] [Google Scholar]
  • 3. Arp DJ, Chain PS, Klotz MG. 2007. The impact of genome analyses on our understanding of ammonia-oxidizing bacteria. Annu. Rev. Microbiol. 61:503–528 [DOI] [PubMed] [Google Scholar]
  • 4. Balasubramanian R, et al. 2010. Oxidation of methane by a biological dicopper centre. Nature 465:115–119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Bartossek R, Nicol GW, Lanzen A, Klenk HP, Schleper C. 2010. Homologues of nitrite reductases in ammonia-oxidizing archaea: diversity and genomic context. Environ. Microbiol. 12:1075–1088 [DOI] [PubMed] [Google Scholar]
  • 6. Bartossek R, Spang A, Weidler G, Lanzen A, Schleper C. 2012. Metagenomic analysis of ammonia-oxidizing archaea affiliated with the soil group. Front. Microbiol. 3:208 doi:10.3389/fmicb.2012.00208 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Bédard C, Knowles R. 1989. Physiology, biochemistry, and specific inhibitors of CH4, NH4+, and CO oxidation by methanotrophs and nitrifiers. Microbiol. Rev. 53:68–84 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Beman JM, Francis CA. 2006. Diversity of ammonia-oxidizing archaea and bacteria in the sediments of a hypernutrified subtropical estuary: Bahia del Tobari, Mexico. Appl. Environ. Microbiol. 72:7767–7777 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Beman JM, Popp BN, Francis CA. 2008. Molecular and biogeochemical evidence for ammonia oxidation by marine Crenarchaeota in the Gulf of California. ISME J. 2:429–441 [DOI] [PubMed] [Google Scholar]
  • 10. Beman JM, Bertics VJ, Braunschweiler T, Wilson JM. 2012. Quantification of ammonia oxidation rates and the distribution of ammonia-oxidizing archaea and bacteria in marine sediment depth profiles from Catalina Island, California. Front. Microbiol. 3:263 doi:10.3389/fmicb.2012.00263 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. Berg IA. 2011. Ecological aspects of distribution of different autotrophic CO2 fixation pathways. Appl. Environ. Microbiol. 77:1925–1936 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12. Bernhard AE, Bollmann A. 2010. Estuarine nitrifiers: new players, patterns and processes. Estuar. Coast. Shelf Sci. 88:1–11 [Google Scholar]
  • 13. Berube PM, Samudrala R, Stahl DA. 2007. Transcription of all amoC copies is associated with recovery of Nitrosomonas europaea from ammonia starvation. J. Bacteriol. 189:3935–3944 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Berube PM, Stahl DA. 2012. The divergent amoC3 subunit of ammonia monooxygenase functions as part of a stress response system in Nitrosomonas europaea. J. Bacteriol. 194:3448–3456 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Blainey PC, Mosier AC, Potanina A, Francis CA, Quake SR. 2011. Genome of a low-salinity ammonia-oxidizing archaeon determined by single-cell and metagenomic analysis. PLoS One 6:e16626 doi:10.1371/journal.pone.0016626 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Bock E, Wagner M. 2006. Oxidation of inorganic nitrogen compounds as an energy source, p 457–495 In Dworkin M, Falkow S, Rosenberg E, Schleifer KH, Stackebrandt E. (ed), The prokaryotes, vol 2: ecophysiology and biochemistry. Springer, New York, NY [Google Scholar]
  • 17. Bouskill NJ, Eveillard D, Chien D, Jayakumar A, Ward BB. 2012. Environmental factors determining ammonia-oxidizing organism distribution and diversity in marine environments. Environ. Microbiol. 14:714–729 [DOI] [PubMed] [Google Scholar]
  • 18. Braker G, Conrad R. 2011. Diversity, structure, and size of N2O-producing microbial communities in soils—what matters for their functioning?, p 33–70 In Laskin AI, Sariaslani S, Gadd GM. (ed), Advances in microbiology, vol 75 Academic Press, Burlington, MA: [DOI] [PubMed] [Google Scholar]
  • 19. Brochier-Armanet C, Boussau B, Gribaldo S, Forterre P. 2008. Mesophilic Crenarchaeota: proposal for a third archaeal phylum, the Thaumarchaeota. Nat. Rev. Microbiol. 6:245–252 [DOI] [PubMed] [Google Scholar]
  • 20. Brochier-Armanet C, Gribaldo S, Forterre P. 2012. Spotlight on the Thaumarchaeota. ISME J. 6:227–230 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Canfield DE, Glazer AN, Falkowski PG. 2010. The evolution and future of Earth's nitrogen cycle. Science 330:192–196 [DOI] [PubMed] [Google Scholar]
  • 22. Chen XP, Zhu YG, Xia Y, Shen JP, He JZ. 2008. Ammonia-oxidizing archaea: important players in paddy rhizosphere soil? Environ. Microbiol. 10:1978–1987 [DOI] [PubMed] [Google Scholar]
  • 23. Church MJ, Wai B, Karl DM, Delong EF. 2010. Abundances of crenarchaeal amoA genes and transcripts in the Pacific Ocean. Environ. Microbiol. 12:679–688 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Coolen MJ, et al. 2007. Putative ammonia-oxidizing Crenarchaeota in suboxic waters of the Black Sea: a basin-wide ecological study using 16S ribosomal and functional genes and membrane lipids. Environ. Microbiol. 9:1001–1016 [DOI] [PubMed] [Google Scholar]
  • 25. Dang H, et al. 2009. Diversity and spatial distribution of amoA-encoding archaea in the deep-sea sediments of the tropical West Pacific continental margin. J. Appl. Microbiol. 106:1482–1493 [DOI] [PubMed] [Google Scholar]
  • 26. De Boer W, Gunnewiek PJ, Veenhuis M, Bock E, Laanbroek HJ. 1991. Nitrification at low pH by aggregated chemolithotrophic bacteria. Appl. Environ. Microbiol. 57:3600–3604 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. De Boer W, Kowalchuk GA. 2001. Nitrification in acid soils: micro-organisms and mechanisms. Soil Biol. Biochem. 33:853–866 [Google Scholar]
  • 28. De Corte D, Yokokawa T, Varela MM, Agogue H, Herndl GJ. 2009. Spatial distribution of Bacteria and Archaea and amoA gene copy numbers throughout the water column of the Eastern Mediterranean Sea. ISME J. 3:147–158 [DOI] [PubMed] [Google Scholar]
  • 29. de la Torre JR, Walker CB, Ingalls AE, Könneke M, Stahl DA. 2008. Cultivation of a thermophilic ammonia oxidizing archaeon synthesizing crenarchaeol. Environ. Microbiol. 10:810–818 [DOI] [PubMed] [Google Scholar]
  • 30. Di HJ, et al. 2009. Nitrification driven by bacteria and not archaea in nitrogen-rich grassland soils. Nat. Geosci. 2:621–624 [Google Scholar]
  • 31. Di HJ, et al. 2010. Ammonia-oxidizing bacteria and archaea grow under contrasting soil nitrogen conditions. FEMS Microbiol. Ecol. 72:386–394 [DOI] [PubMed] [Google Scholar]
  • 32. Dodsworth JA, Hungate B, de la Torre JR, Jiang H, Hedlund BP. 2011. Measuring nitrification, denitrification, and related biomarkers in terrestrial geothermal ecosystems. Methods Enzymol. 486:171–203 [DOI] [PubMed] [Google Scholar]
  • 33. Dodsworth JA, Hungate B, Hedlund BP. 2011. Ammonia oxidation, denitrification and dissimilatory nitrate reduction to ammonium in two US Great Basin hot springs with abundant ammonia-oxidizing archaea. Environ. Microbiol. 13:2371–2386 [DOI] [PubMed] [Google Scholar]
  • 34. Dodsworth JA, McDonald AI, Hedlund BP. 2012. Calculation of total free energy yield as an alternative approach for predicting the importance of potential chemolithotrophic reactions in geothermal springs. FEMS Microbiol. Ecol. 81:446–454 [DOI] [PubMed] [Google Scholar]
  • 35. Erguder TH, Boon N, Wittebolle L, Marzorati M, Verstraete W. 2009. Environmental factors shaping the ecological niches of ammonia-oxidizing archaea. FEMS Microbiol. Rev. 33:855–869 [DOI] [PubMed] [Google Scholar]
  • 36. French E, Kozlowski JA, Mukherjee M, Bullerjahn G, Bollmann A. 2012. Ecophysiological characterization of ammonia-oxidizing archaea and bacteria from freshwater. Appl. Environ. Microbiol. 78:5773–5780 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Geets J, Boon N, Verstraete W. 2006. Strategies of aerobic ammonia-oxidizing bacteria for coping with nutrient and oxygen fluctuations. FEMS Microbiol. Ecol. 58:1–13 [DOI] [PubMed] [Google Scholar]
  • 38. Glass JB, Orphan VJ. 2012. Trace metal requirements for microbial enzymes involved in the production and consumption of methane and nitrous oxide. Front. Microbiol. 3:61 doi:10.3389/fmicb.2012.00061 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Godfrey LV, Glass JB. 2011. The geochemical record of the ancient nitrogen cycle, nitrogen isotopes, and metal cofactors. Methods Enzymol. 486:483–506 [DOI] [PubMed] [Google Scholar]
  • 40. Groussin M, Gouy M. 2011. Adaptation to environmental temperature is a major determinant of molecular evolutionary rates in archaea. Mol. Biol. Evol. 28:2661–2674 [DOI] [PubMed] [Google Scholar]
  • 41. Gubry-Rangin C, Nicol GW, Prosser JI. 2010. Archaea rather than bacteria control nitrification in two agricultural acidic soils. FEMS Microbiol. Ecol. 74:566–574 [DOI] [PubMed] [Google Scholar]
  • 42. Gubry-Rangin C, et al. 2011. Niche specialization of terrestrial archaeal ammonia oxidizers. Proc. Natl. Acad. Sci. U. S. A. 108:21206–21211 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43. Guerrero M, Jones R. 1996. Photoinhibition of marine nitrifying bacteria. I. Wavelength dependent response. Mar. Ecol. Progr. Ser. 141:183–192 [Google Scholar]
  • 44. Hallam SJ, et al. 2006. Genomic analysis of the uncultivated marine crenarchaeote Cenarchaeum symbiosum. Proc. Natl. Acad. Sci. U. S. A. 103:18296–18301 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Hallam SJ, et al. 2006. Pathways of carbon assimilation and ammonia oxidation suggested by environmental genomic analyses of marine Crenarchaeota. PLoS Biol. 4:e95 doi:10.1371/journal.pbio.0040095 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Hatzenpichler R, et al. 2008. A moderately thermophilic ammonia-oxidizing crenarchaeote from a hot spring. Proc. Natl. Acad. Sci. U. S. A. 105:2134–2139 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. He J, et al. 2007. Quantitative analyses of the abundance and composition of ammonia-oxidizing bacteria and ammonia-oxidizing archaea of a Chinese upland red soil under long-term fertilization practices. Environ. Microbiol. 9:2364–2374 [DOI] [PubMed] [Google Scholar]
  • 48. Herndl GJ, et al. 2005. Contribution of Archaea to total prokaryotic production in the deep Atlantic Ocean. Appl. Environ. Microbiol. 71:2303–2309 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Hollibaugh JT, Gifford S, Sharma S, Bano N, Moran MA. 2011. Metatranscriptomic analysis of ammonia-oxidizing organisms in an estuarine bacterioplankton assemblage. ISME J. 5:866–878 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Hooper AB, Terry KR. 1974. Photoinactivation of ammonia oxidation in Nitrosomonas. J. Bacteriol. 119:899–906 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Hooper AB, Vannelli T, Bergmann DJ, Arciero DM. 1997. Enzymology of the oxidation of ammonia to nitrite by bacteria. Antonie Van Leeuwenhoek 71:59–67 [DOI] [PubMed] [Google Scholar]
  • 52. Horrigan SG, Springer AL. 1990. Oceanic and estuarine ammonium oxidation: effects of light. Limnol. Oceanogr. 35:479–482 [Google Scholar]
  • 53. Hyman MR, Arp DJ. 1992. 14C2H2- and 14CO2-labeling studies of the de novo synthesis of polypeptides by Nitrosomonas europaea during recovery from acetylene and light inactivation of ammonia monooxygenase. J. Biol. Chem. 267:1534–1545 [PubMed] [Google Scholar]
  • 54. Ingalls AE, et al. 2006. Quantifying archaeal community autotrophy in the mesopelagic ocean using natural radiocarbon. Proc. Natl. Acad. Sci. U. S. A. 103:6442–6447 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55. Intergovernmental Panel on Climate Change 2007. The physical science basis: contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom [Google Scholar]
  • 56. Jia Z, Conrad R. 2009. Bacteria rather than Archaea dominate microbial ammonia oxidation in an agricultural soil. Environ. Microbiol. 11:1658–1671 [DOI] [PubMed] [Google Scholar]
  • 57. Jiang H, et al. 2010. RNA-based investigation of ammonia-oxidizing archaea in hot springs of Yunnan Province, China. Appl. Environ. Microbiol. 76:4538–4541 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58. Jung M-Y, et al. 2011. Enrichment and characterization of an autotrophic ammonia-oxidizing archaeon of mesophilic crenarchaeal group I.1a from an agricultural soil. Appl. Environ. Microbiol. 77:8635–8647 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59. Karner MB, DeLong EF, Karl DM. 2001. Archaeal dominance in the mesopelagic zone of the Pacific Ocean. Nature 409:507–510 [DOI] [PubMed] [Google Scholar]
  • 60. Kim BK, et al. 2011. Genome sequence of an ammonia-oxidizing soil archaeon, “Candidatus Nitrosoarchaeum koreensis” MY1. J. Bacteriol. 193:5539–5540 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Kim JG, et al. 2012. Cultivation of a highly enriched ammonia-oxidizing archaeon of thaumarchaeotal group I.1b from an agricultural soil. Environ. Microbiol. 14:1528–1543 [DOI] [PubMed] [Google Scholar]
  • 62. Klotz MG, Alzerreca J, Norton JM. 1997. A gene encoding a membrane protein exists upstream of the amoA/amoB genes in ammonia oxidizing bacteria: a third member of the amo operon? FEMS Microbiol. Lett. 150:65–73 [DOI] [PubMed] [Google Scholar]
  • 63. Klotz MG, Norton JM. 1998. Ammonia monooxygenase has evolved by operon duplication under biased AT/GC mutational pressure in ammonia-oxidizing autotrophic bacteria. FEMS Microbiol. Lett. 168:303–311 [DOI] [PubMed] [Google Scholar]
  • 64. Klotz MG, Stein LY. 2008. Nitrifier genomics and evolution of the nitrogen cycle. FEMS Microbiol. Lett. 278:146–156 [DOI] [PubMed] [Google Scholar]
  • 65. Könneke M, et al. 2005. Isolation of an autotrophic ammonia-oxidizing marine archaeon. Nature 437:543–546 [DOI] [PubMed] [Google Scholar]
  • 66. Koops HP, Purkhold U, Pommerening-Röser A, Timmermann G, Wagner M. 2003. The lithoautotrophic ammonia-oxidizing bacteria, p 778–811 In Dworkin M. (ed), The procaryotes: an evolving electronic resource for the microbiological community, 3rd ed Springer-Verlag, New York, NY [Google Scholar]
  • 67. Koper TE, El-Sheikh AF, Norton JM, Klotz MG. 2004. Urease-encoding genes in ammonia-oxidizing bacteria. Appl. Environ. Microbiol. 70:2342–2348 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68. Labrenz M, et al. 2010. Relevance of a crenarchaeotal subcluster related to Candidatus Nitrosopumilus maritimus to ammonia oxidation in the suboxic zone of the central Baltic Sea. ISME J. 4:1496–1508 [DOI] [PubMed] [Google Scholar]
  • 69. Lam P, et al. 2007. Linking crenarchaeal and bacterial nitrification to anammox in the Black Sea. Proc. Natl. Acad. Sci. U. S. A. 104:7104–7109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70. Lam P, et al. 2009. Revising the nitrogen cycle in the Peruvian oxygen minimum zone. Proc. Natl. Acad. Sci. U. S. A. 106:4752–4757 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Lehtovirta-Morley LE, Stoecker K, Vilcinskas A, Prosser JI, Nicol GW. 2011. Cultivation of an obligate acidophilic ammonia oxidizer from a nitrifying acid soil. Proc. Natl. Acad. Sci. U. S. A. 108:15892–15897 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Leininger S, et al. 2006. Archaea predominate among ammonia-oxidizing prokaryotes in soils. Nature 442:806–809 [DOI] [PubMed] [Google Scholar]
  • 73. Levičnik-Höfferle S, Nicol GW, Ausec L, Mandic-Mulec I, Prosser JI. 2012. Stimulation of thaumarchaeal ammonia oxidation by ammonia derived from organic nitrogen but not added inorganic nitrogen. FEMS Microbiol. Ecol. 80:114–123 [DOI] [PubMed] [Google Scholar]
  • 74. Lieberman RL, Rosenzweig AC. 2005. Crystal structure of a membrane-bound metalloenzyme that catalyses the biological oxidation of methane. Nature 434:177–182 [DOI] [PubMed] [Google Scholar]
  • 75. Lipschultz F, et al. 1981. Production of NO and N2O by soil nitrifying bacteria. Nature 294:641–643 [Google Scholar]
  • 76. Loescher CR, et al. 2012. Production of oceanic nitrous oxide by ammonia-oxidizing archaea. Biogeosci. Discuss. 9:2095–2122 [Google Scholar]
  • 77. Lu L, et al. 17 May 2012, posting date Nitrification of archaeal ammonia oxidizers in acid soils is supported by hydrolysis of urea. ISME J. [Epub ahead of print.] doi:10.1038/ismej.2012.45 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78. Lund MB, Smith JM, Francis CA. 17 May 2012, posting date Diversity, abundance and expression of nitrite reductase (nirK)-like genes in marine thaumarchaea. ISME J. [Epub ahead of print.] doi:10.1038/ismej.2012.40 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79. Martens-Habbena W, Berube PM, Urakawa H, de la Torre JR, Stahl DA. 2009. Ammonia oxidation kinetics determine niche separation of nitrifying Archaea and Bacteria. Nature 461:976–979 [DOI] [PubMed] [Google Scholar]
  • 80. Martens-Habbena W, Stahl DA. 2011. Nitrogen metabolism and kinetics of ammonia-oxidizing archaea. Methods Enzymol. 496:465–487 [DOI] [PubMed] [Google Scholar]
  • 81. Martin-Cuadrado AB, et al. 2008. Hindsight in the relative abundance, metabolic potential and genome dynamics of uncultivated marine archaea from comparative metagenomic analyses of bathypelagic plankton of different oceanic regions. ISME J. 2:865–886 [DOI] [PubMed] [Google Scholar]
  • 82. Massana R, DeLong EF, Pedros-Alio C. 2000. A few cosmopolitan phylotypes dominate planktonic archaeal assemblages in widely different oceanic provinces. Appl. Environ. Microbiol. 66:1777–1787 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83. Merbt SN, et al. 2012. Differential photoinhibition of bacterial and archaeal ammonia oxidation. FEMS Microbiol. Lett. 327:41–46 [DOI] [PubMed] [Google Scholar]
  • 84. Mincer TJ, et al. 2007. Quantitative distribution of presumptive archaeal and bacterial nitrifiers in Monterey Bay and the North Pacific subtropical gyre. Environ. Microbiol. 9:1162–1175 [DOI] [PubMed] [Google Scholar]
  • 85. Miranda KM, et al. 2005. Comparison of the chemical biology of NO and HNO: an inorganic perspective, p 349–384 In Karlin KD. (ed), Progress in inorganic chemistry, vol 54 John Wiley & Sons, Inc., Hoboken, NJ [Google Scholar]
  • 86. Molina V, Belmar L, Ulloa O. 2010. High diversity of ammonia-oxidizing archaea in permanent and seasonal oxygen-deficient waters of the eastern South Pacific. Environ. Microbiol. 12:2450–2465 [DOI] [PubMed] [Google Scholar]
  • 87. Mosier AC, Francis CA. 2011. Determining the distribution of marine and coastal ammonia-oxidizing archaea and bacteria using a quantitative approach. Methods Enzymol. 486:205–221 [DOI] [PubMed] [Google Scholar]
  • 88. Mosier AC, Allen EE, Kim M, Ferriera S, Francis CA. 2012. Genome sequence of “Candidatus Nitrosopumilus salaria” BD31, an ammonia-oxidizing archaeon from the San Francisco Bay estuary. J. Bacteriol. 194:2121–2122 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Mosier AC, Allen EE, Kim M, Ferriera S, Francis CA. 2012. Genome sequence of “Candidatus Nitrosoarchaeum limnia” BG20, a low-salinity ammonia-oxidizing archaeon from the San Francisco Bay estuary. J. Bacteriol. 194:2119–2120 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90. Mosier AC, Lund MB, Francis CA. 30 May 2012, posting date Ecophysiology of an ammonia-oxidizing archaeon adapted to low-salinity habitats. Microb. Ecol. [Epub ahead of print.] doi:10.1007/s00248-012-0075-1 [DOI] [PubMed] [Google Scholar]
  • 91. Muller F, Brissac T, Le Bris N, Felbeck H, Gros O. 2010. First description of giant Archaea (Thaumarchaeota) associated with putative bacterial ectosymbionts in a sulfidic marine habitat. Environ. Microbiol. 12:2371–2383 [DOI] [PubMed] [Google Scholar]
  • 92. Mussmann M, et al. 2011. Thaumarchaeotes abundant in refinery nitrifying sludges express amoA but are not obligate autotrophic ammonia oxidizers. Proc. Natl. Acad. Sci. U. S. A. 108:16771–16776 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93. Nicol GW, Leininger S, Schleper C, Prosser JI. 2008. The influence of soil pH on the diversity, abundance and transcriptional activity of ammonia oxidizing archaea and bacteria. Environ. Microbiol. 10:2966–2978 [DOI] [PubMed] [Google Scholar]
  • 94. Nicol GW, Leininger S, Schleper C. 2011. Distribution and activity of ammonia-oxidizing archaea in natural environments, p 157–180 In Ward BB, Arp DJ, Klotz MG. (ed), Nitrification. ASM Press, Washington, DC [Google Scholar]
  • 95. Nunoura T, et al. 2011. Insights into the evolution of Archaea and eukaryotic protein modifier systems revealed by the genome of a novel archaeal group. Nucleic Acids Res. 39:3204–3223 doi:10.1093/nar/gkq1228 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Offre P, Prosser JI, Nicol GW. 2009. Growth of ammonia-oxidizing archaea in soil microcosms is inhibited by acetylene. FEMS Microbiol. Ecol. 70:99–108 [DOI] [PubMed] [Google Scholar]
  • 97. Offre P, Nicol GW, Prosser JI. 2011. Community profiling and quantification of putative autotrophic thaumarchaeal communities in environmental samples. Environ. Microbiol. Rep. 3:245–253 [DOI] [PubMed] [Google Scholar]
  • 98. Ouverney CC, Fuhrman JA. 2000. Marine planktonic archaea take up amino acids. Appl. Environ. Microbiol. 66:4829–4833 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99. Park B-J, et al. 2010. Cultivation of autotrophic ammonia-oxidizing archaea from marine sediments in coculture with sulfur-oxidizing bacteria. Appl. Environ. Microbiol. 76:7575–7587 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Park HD, Wells GF, Bae H, Criddle CS, Francis CA. 2006. Occurrence of ammonia-oxidizing archaea in wastewater treatment plant bioreactors. Appl. Environ. Microbiol. 72:5643–5647 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Pelve EA, et al. 2011. Cdv-based cell division and cell cycle organization in the thaumarchaeon Nitrosopumilus maritimus. Mol. Microbiol. 82:555–566 [DOI] [PubMed] [Google Scholar]
  • 102. Pester M, Schleper C, Wagner M. 2011. The Thaumarchaeota: an emerging view of their phylogeny and ecophysiology. Curr. Opin. Microbiol. 14:300–306 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103. Pester M, et al. 2012. amoA-based consensus phylogeny of ammonia-oxidizing archaea and deep sequencing of amoA genes from soils of four different geographic regions. Environ. Microbiol. 14:525–539 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104. Pitcher A, et al. 2011. Niche segregation of ammonia-oxidizing archaea and anammox bacteria in the Arabian Sea oxygen minimum zone. ISME J. 5:1896–1904 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105. Pommerening-Röser A, Koops HP. 2005. Environmental pH as an important factor for the distribution of urease positive ammonia-oxidizing bacteria. Microbiol. Res. 160:27–35 [DOI] [PubMed] [Google Scholar]
  • 106. Pratscher J, Dumont MG, Conrad R. 2011. Ammonia oxidation coupled to CO2 fixation by archaea and bacteria in an agricultural soil. Proc. Natl. Acad. Sci. U. S. A. 108:4170–4175 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107. Preston CM, Wu KY, Molinski TF, DeLong EF. 1996. A psychrophilic crenarchaeon inhabits a marine sponge: Cenarchaeum symbiosum gen. nov., sp. nov. Proc. Natl. Acad. Sci. U. S. A. 93:6241–6246 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108. Prosser JI, Nicol GW. 2008. Relative contributions of archaea and bacteria to aerobic ammonia oxidation in the environment. Environ. Microbiol. 10:2931–2941 [DOI] [PubMed] [Google Scholar]
  • 109. Rasche ME, Hyman MR, Arp DJ. 1990. Biodegradation of halogenated hydrocarbon fumigants by nitrifying bacteria. Appl. Environ. Microbiol. 56:2568–2571 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110. 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] [PubMed] [Google Scholar]
  • 111. Reigstad LJ, et al. 2008. Nitrification in terrestrial hot springs of Iceland and Kamchatka. FEMS Microbiol. Ecol. 64:167–174 [DOI] [PubMed] [Google Scholar]
  • 112. Rogers DR, Casciotti KL. 2010. Abundance and diversity of archaeal ammonia oxidizers in a coastal groundwater system. Appl. Environ. Microbiol. 76:7938–7948 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113. Santoro AE, Francis CA, de Sieyes NR, Boehm AB. 2008. Shifts in the relative abundance of ammonia-oxidizing bacteria and archaea across physicochemical gradients in a subterranean estuary. Environ. Microbiol. 10:1068–1079 [DOI] [PubMed] [Google Scholar]
  • 114. Santoro AE, Casciotti KL, Francis CA. 2010. Activity, abundance and diversity of nitrifying archaea and bacteria in the central California Current. Environ. Microbiol. 12:1989–2006 [DOI] [PubMed] [Google Scholar]
  • 115. Santoro AE, Buchwald C, McIlvin MR, Casciotti KL. 2011. Isotopic signature of N2O produced by marine ammonia-oxidizing archaea. Science 333:1282–1285 [DOI] [PubMed] [Google Scholar]
  • 116. Santoro AE, Casciotti KL. 2011. Enrichment and characterization of ammonia-oxidizing archaea from the open ocean: phylogeny, physiology and stable isotope fractionation. ISME J. 5:1796–1808 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117. Schauss K, et al. 2009. Dynamics and functional relevance of ammonia-oxidizing archaea in two agricultural soils. Environ. Microbiol. 11:446–456 [DOI] [PubMed] [Google Scholar]
  • 118. Schleper C, Jurgens G, Jonuscheit M. 2005. Genomic studies of uncultivated archaea. Nat. Rev. Microbiol. 3:479–488 [DOI] [PubMed] [Google Scholar]
  • 119. Schleper C, Nicol GW. 2010. Ammonia-oxidising archaea—physiology, ecology and evolution. Adv. Microb. Physiol. 57:1–41 [DOI] [PubMed] [Google Scholar]
  • 120. Shi Y, Tyson GW, Eppley JM, DeLong EF. 2011. Integrated metatranscriptomic and metagenomic analyses of stratified microbial assemblages in the open ocean. ISME J. 5:999–1013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121. Simon HM, Dodsworth JA, Goodman RM. 2000. Crenarchaeota colonize terrestrial plant roots. Environ. Microbiol. 2:495–505 [DOI] [PubMed] [Google Scholar]
  • 122. Simon HM, et al. 2005. Cultivation of mesophilic soil crenarchaeotes in enrichment cultures from plant roots. Appl. Environ. Microbiol. 71:4751–4760 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123. Simon J, Klotz MG. 25 July 2012, posting date Diversity and evolution of bioenergetic systems involved in microbial nitrogen compound transformations. Biochim. Biophys. Acta accepted. [Epub ahead of print.] http://dx.doi.org/10.1016/j.bbabio.2012.07.005 [DOI] [PubMed]
  • 124. Spang A, et al. 2010. Distinct gene set in two different lineages of ammonia-oxidizing archaea supports the phylum Thaumarchaeota. Trends Microbiol. 18:331–340 [DOI] [PubMed] [Google Scholar]
  • 125. Stewart FJ, Ulloa O, DeLong EF. 2012. Microbial metatranscriptomics in a permanent marine oxygen minimum zone. Environ. Microbiol. 14:23–40 [DOI] [PubMed] [Google Scholar]
  • 126. Stolper DA, Revsbech NP, Canfield DE. 2010. Aerobic growth at nanomolar oxygen concentrations. Proc. Natl. Acad. Sci. U. S. A. 107:18755–18760 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127. Stopnisek N, et al. 2010. Thaumarchaeal ammonia oxidation in an acidic forest peat soil is not influenced by ammonium amendment. Appl. Environ. Microbiol. 76:7626–7634 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128. Suzuki I, Dular U, Kwok SC. 1974. Ammonia or ammonium ion as substrate for oxidation by Nitrosomonas europaea cells and extracts. J. Bacteriol. 120:556–558 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129. Tavormina PL, Orphan VJ, Kalyuzhnaya MG, Jetten MSM, Klotz MG. 2011. A novel family of functional operones encoding methane/ammonia monooxygenase-related proteins in gammaproteobacterial methanotrophs. Environ. Microbiol. Rep. 3:91–100 [DOI] [PubMed] [Google Scholar]
  • 130. Tourna M, Freitag TE, Nicol GW, Prosser JI. 2008. Growth, activity and temperature responses of ammonia-oxidizing archaea and bacteria in soil microcosms. Environ. Microbiol. 10:1357–1364 [DOI] [PubMed] [Google Scholar]
  • 131. Tourna M, et al. 2011. Nitrososphaera viennensis, an ammonia oxidizing archaeon from soil. Proc. Natl. Acad. Sci. U. S. A. 108:8420–8425 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132. Treusch AH, et al. 2005. Novel genes for nitrite reductase and Amo-related proteins indicate a role of uncultivated mesophilic crenarchaeota in nitrogen cycling. Environ. Microbiol. 7:1985–1995 [DOI] [PubMed] [Google Scholar]
  • 133. Tully BJ, Nelson WC, Heidelberg JF. 2012. Metagenomic analysis of a complex marine planktonic thaumarchaeal community from the Gulf of Maine. Environ. Microbiol. 14:254–267 [DOI] [PubMed] [Google Scholar]
  • 134. Urakawa H, Martens-Habbena W, Stahl DA. 2011. Physiology and genomics of ammonia-oxidizing archaea, p 117–155 In Ward BB, Arp DJ, Klotz MG. (ed), Nitrification. ASM Press, Washington, DC [Google Scholar]
  • 135. Valentine DL. 2007. Adaptations to energy stress dictate the ecology and evolution of the Archaea. Nat. Rev. Microbiol. 5:316–323 [DOI] [PubMed] [Google Scholar]
  • 136. Varela MM, van Aken HM, Sintes E, Reinthaler T, Herndl GJ. 2011. Contribution of Crenarchaeota and Bacteria to autotrophy in the North Atlantic interior. Environ. Microbiol. 13:1524–1533 [DOI] [PubMed] [Google Scholar]
  • 137. Venter JC, et al. 2004. Environmental genome shotgun sequencing of the Sargasso Sea. Science 304:66–74 [DOI] [PubMed] [Google Scholar]
  • 138. Verhamme DT, Prosser JI, Nicol GW. 2011. Ammonia concentration determines differential growth of ammonia-oxidising archaea and bacteria in soil microcosms. ISME J. 5:1067–1071 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139. Wagner M. 2009. Single-cell ecophysiology of microbes as revealed by Raman microspectroscopy or secondary ion mass spectrometry imaging. Annu. Rev. Microbiol. 63:411–429 [DOI] [PubMed] [Google Scholar]
  • 140. Walker CB, et al. 2010. Nitrosopumilus maritimus genome reveals unique mechanisms for nitrification and autotrophy in globally distributed marine crenarchaea. Proc. Natl. Acad. Sci. U. S. A. 107:8818–8823 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141. Wang S, et al. 2009. Diversity and abundance of ammonia-oxidizing archaea in hydrothermal vent chimneys of the Juan de Fuca Ridge. Appl. Environ. Microbiol. 75:4216–4220 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142. Watanabe T, Lee CG, Murase J, Asakawa S, Kimura M. 2011. Carbon flow into ammonia-oxidizing bacteria and archaea during decomposition of 13C-labeled plant residues in soil. Soil Sci. Plant Nutr. 57:775–785 [Google Scholar]
  • 143. Wells GF, et al. 2009. Ammonia-oxidizing communities in a highly aerated full-scale activated sludge bioreactor: betaproteobacterial dynamics and low relative abundance of Crenarchaea. Environ. Microbiol. 11:2310–2328 [DOI] [PubMed] [Google Scholar]
  • 144. Wuchter C, et al. 2006. Archaeal nitrification in the ocean. Proc. Natl. Acad. Sci. U. S. A. 103:12317–12322 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145. Xia W, et al. 2011. Autotrophic growth of nitrifying community in an agricultural soil. ISME J. 5:1226–1236 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146. Xu M, Schnorr J, Keibler B, Simon HM. 2012. Comparative analysis of 16S rRNA and amoA genes from archaea selected with organic and inorganic amendments in enrichment culture. Appl. Environ. Microbiol. 78:2137–2146 doi:10.1128/AEM.06845-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147. Yakimov MM, et al. 2011. Contribution of crenarchaeal autotrophic ammonia oxidizers to the dark primary production in Tyrrhenian deep waters (Central Mediterranean Sea). ISME J. 5:945–961 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148. Zhalnina K, de Quadros PD, Camargo FA, Triplett EW. 2012. Drivers of archaeal ammonia-oxidizing communities in soil. Front. Microbiol. 3:210 doi:10.3389/fmicb.2012.00210 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149. Zhang CL, et al. 2008. Global occurrence of archaeal amoA genes in terrestrial hot springs. Appl. Environ. Microbiol. 74:6417–6426 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150. Zhang LM, et al. 2010. Autotrophic ammonia oxidation by soil thaumarchaea. Proc. Natl. Acad. Sci. U. S. A. 107:17240–17245 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151. Zhang LM, Hu HW, Shen JP, He JZ. 2012. Ammonia-oxidizing archaea have more important role than ammonia-oxidizing bacteria in ammonia oxidation of strongly acidic soils. ISME J. 6:1032–1045 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152. Zhang T, et al. 2009. Occurrence of ammonia-oxidizing Archaea in activated sludges of a laboratory scale reactor and two wastewater treatment plants. J. Appl. Microbiol. 107:970–977 [DOI] [PubMed] [Google Scholar]
  • 153. Zumft WG. 1997. Cell biology and molecular basis of denitrification. Microbiol. Mol. Biol. Rev. 61:533–616 [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

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

RESOURCES