首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 343 毫秒
1.
Previous studies have documented the capacity of European earthworms belonging to the family Lumbricidae to emit the greenhouse gas nitrous oxide (N2O), an activity attributed primarily to the activation of ingested soil denitrifiers. To extend the information base to earthworms in the Southern Hemisphere, four species of earthworms in New Zealand were examined for gut-associated denitrification. Lumbricus rubellus and Aporrectodea rosea (introduced species of Lumbricidae) emitted N2O, whereas emission of N2O by Octolasion cyaneum (an introduced species of Lumbricidae) and emission of N2O by Octochaetus multiporus (a native species of Megascolecidae) were variable and negligible, respectively. Exposing earthworms to nitrite or nitrate and acetylene significantly increased the amount of N2O emitted, implicating denitrification as the primary source of N2O and indicating that earthworms emitted dinitrogen (N2) in addition to N2O. The alimentary canal displayed a high capacity to produce N2O when it was supplemented with nitrite, and alimentary canal contents contained large amounts of carbohydrates and organic acids indicative of fermentation (e.g., succinate, acetate, and formate) that could serve as sources of reductant for denitrification. nosZ encodes a portion of the terminal oxidoreductase used in denitrification. The nosZ sequences detected in the alimentary canals of L. rubellus and O. multiporus were similar to those retrieved from soil and were distantly related to sequences of uncultured soil bacteria and genera common in soils (i.e., Bradyrhizobium, Azospirillum, Rhodopseudomonas, Rhodospirillum, Pseudomonas, Oligotropha, and Sinorhizobium). These findings (i) suggest that the capacity to emit N2O and N2 is a general trait of earthworms and not geographically restricted, (ii) indicate that species belonging to different earthworm families (i.e., Megascolecidae and Lumbricidae) may not have equal capacities to emit N2O, and (iii) also corroborate previous findings that link this capacity to denitrification in the alimentary canal.Earthworms are dominant members of the soil fauna and affect the structure and fertility of soils (5, 20, 22, 23). Various species of European earthworms belonging to the family Lumbricidae (e.g., Aporrectodea caliginosa, Lumbricus rubellus, and Octolasion lacteum) emit dinitrogen (N2) and the greenhouse gas nitrous oxide (N2O), and their burrowing activities and feeding habits in combination with in situ conditions can influence the emission of nitrogenous gases from soils that they inhabit (1, 2, 13, 17, 25, 27, 39).The microbiology of the earthworm alimentary canal has been addressed in numerous studies (3, 4, 6, 9, 14, 16, 32). The alimentary canal of the earthworm is anoxic, in marked contrast to the aerated material that earthworms ingest (14, 39). Anoxia and other in situ conditions of the alimentary canal appear to stimulate soil microbes capable of surviving under anaerobic conditions during passage through the gut (3, 4). Soils are rich in denitrifying bacteria (37), and the capacity of European earthworms to emit nitrogenous gases has been attributed primarily to the in situ activity of ingested denitrifying bacteria that appear to be highly active under the anoxic conditions of the earthworm alimentary canal (12, 15, 17, 25, 39). However, it is not known if the capacity to emit nitrogenous gases is a general trait of earthworms independent of their taxonomic family or geographic location. The main objectives of this study were to examine the capacity of Southern Hemisphere earthworms in New Zealand to emit N2O and to determine if this capacity was linked to denitrifying bacteria in the alimentary canal.  相似文献   

2.
Most studies of bacterial denitrification have used nitrate (NO3) as the first electron acceptor, whereas relatively less is understood about nitrite (NO2) denitrification. We isolated novel bacteria that proliferated in the presence of high levels of NO2 (72 mM). Strain YD50.2, among several isolates, was taxonomically positioned within the α subclass of Proteobacteria and identified as Ochrobactrum anthropi YD50.2. This strain denitrified NO2, as well as NO3. The gene clusters for denitrification (nar, nir, nor, and nos) were cloned from O. anthropi YD50.2, in which the nir and nor operons were linked. We confirmed that nirK in the nir-nor operon produced a functional NO2 reductase containing copper that was involved in bacterial NO2 reduction. The strain denitrified up to 40 mM NO2 to dinitrogen under anaerobic conditions in which other denitrifiers or NO3 reducers such as Pseudomonas aeruginosa and Ralstonia eutropha and nitrate-respiring Escherichia coli neither proliferated nor reduced NO2. Under nondenitrifying aerobic conditions, O. anthropi YD50.2 and its type strain ATCC 49188T proliferated even in the presence of higher levels of NO2 (100 mM), and both were considerably more resistant to acidic NO2 than were the other strains noted above. These results indicated that O. anthropi YD50.2 is a novel denitrifier that has evolved reactive nitrogen oxide tolerance mechanisms.Environmental bacteria maintain the global nitrogen cycle by metabolizing organic and inorganic nitrogen compounds. Denitrification is critical for maintenance of the global nitrogen cycle, through which nitrate (NO3) or nitrite (NO2) is reduced to gaseous nitrogen forms such as N2 and nitrous oxide (N2O) (19, 47). Decades of investigations into denitrifying bacteria have revealed their ecological impact (9), their molecular mechanisms of denitrification (13, 25, 47), and the industrial importance of removing nitrogenous contaminants from wastewater (31, 36). Bacterial denitrification is considered to comprise four successive reduction steps, each of which is catalyzed by NO3 reductase (Nar), NO2 reductase (Nir), nitric oxide (NO) reductase (Nor), and N2O reductase (Nos). The reaction of each enzyme is linked to the electron transport chain on the cellular membrane and accompanies oxidative phosphorylation, implying that bacterial denitrification is of as much physiological significance as anaerobic respiration (25, 47). Most denitrifying bacteria are facultative anaerobes and respire with oxygen under aerobic conditions. Because denitrification is induced in the absence of oxygen, it is considered an alternative mechanism of energy conservation that has evolved as an adaptation to anaerobic circumstances (13, 47).Nitrite and NO are hazardous to bacteria, since they generate highly reactive nitrogen species (RNS) under physiological conditions and damage cellular DNA, lipid, and proteins (28, 37). Denitrifying bacteria are thought to be threatened by RNS since they reduce NO3 to generate NO2 and NO as denitrifying intermediates. Furthermore, denitrifying bacteria often inhabit environments where they are exposed to NO2 and NO and hence high levels of RNS. Recent reports suggest that pathogenic bacteria invading animal tissues are attacked by NO generated by macrophages (12). Such bacteria involve denitrifiers, and some of them, for example, Neisseria meningitidis (1) and Pseudomonas aeruginosa, acquire resistance to NO by producing Nor (44). The utilization (reduction) of NO by Brucella increases the survival of infected mice (2). These examples suggest that production of a denitrifying mechanism affects bacterial survival of threats from both endogenous and extracellular RNS. However, the mechanism of RNS tolerance induced by denitrifying bacteria is not fully understood.Ubiquitous gram-negative Ochrobactrum strains are widely distributed in soils and aqueous environments, where they biodegrade aromatic compounds (11), organophosphorus pesticides (45), and other hydrocarbons (38) and remove heavy metal ions such as chromium and cadmium (24). Having been isolated from clinical specimens, Ochrobactrum anthropi is currently recognized as an emerging opportunistic pathogen, although relatively little is known about its pathogenesis and factors contributing to its virulence (7, 30). Manipulation systems have been developed to investigate these issues at the molecular genetic level (33). Some O. anthropi strains have been identified as denitrifiers (21), although the denitrifying properties of these strains have not been investigated in detail. This study was undertaken to examine the denitrifying properties of O. anthropi in more detail. O. anthropi YD50.2 was selected for this study and was isolated herein. The strain denitrified high levels of NO2 (up to 40 mM) to dinitrogen under anaerobic conditions. The strain was highly resistant to acidified NO2 under nondenitrifying aerobic conditions. These results indicate that O. anthropi YD50.2 has mechanisms that produce tolerance to RNS.  相似文献   

3.
4.
5.
6.
The objective of this study was to investigate how changes in soil pH affect the N2O and N2 emissions, denitrification activity, and size of a denitrifier community. We established a field experiment, situated in a grassland area, which consisted of three treatments which were repeatedly amended with a KOH solution (alkaline soil), an H2SO4 solution (acidic soil), or water (natural pH soil) over 10 months. At the site, we determined field N2O and N2 emissions using the 15N gas flux method and collected soil samples for the measurement of potential denitrification activity and quantification of the size of the denitrifying community by quantitative PCR of the narG, napA, nirS, nirK, and nosZ denitrification genes. Overall, our results indicate that soil pH is of importance in determining the nature of denitrification end products. Thus, we found that the N2O/(N2O + N2) ratio increased with decreasing pH due to changes in the total denitrification activity, while no changes in N2O production were observed. Denitrification activity and N2O emissions measured under laboratory conditions were correlated with N fluxes in situ and therefore reflected treatment differences in the field. The size of the denitrifying community was uncoupled from in situ N fluxes, but potential denitrification was correlated with the count of NirS denitrifiers. Significant relationships were observed between nirS, napA, and narG gene copy numbers and the N2O/(N2O + N2) ratio, which are difficult to explain. However, this highlights the need for further studies combining analysis of denitrifier ecology and quantification of denitrification end products for a comprehensive understanding of the regulation of N fluxes by denitrification.Denitrification is the microbial reduction of NO3 via NO2 to gaseous NO, N2O, and N2, which are then lost into the atmosphere (36). It therefore results in considerable loss of nitrogen, one of the most limiting nutrients for crop production in agriculture (20). Denitrification is also of environmental concern since, together with nitrification, it is the main biological process responsible for N2O emissions (7). N2O is a potent greenhouse gas which has a global warming potential about 320 times greater than that of CO2 and has a lifetime of approximately 120 years (32). In the stratosphere, N2O can also react with O2 to produce NO, which induces the destruction of stratospheric ozone (8). N2O can be released into the atmosphere by incomplete denitrification due to the effect of environmental conditions on the regulation of the different denitrification reductases (14, 41, 51), but it has recently been suggested that it could also be due to lack of nitrous oxide reductase in some denitrifiers (19, 41). Since N2O is an intermediate in the denitrification pathway, both the amount of N2O produced and the N2O/(N2O + N2) ratio are important in understanding and predicting N2O fluxes from soils.The main environmental factors known to influence the N2O/(N2O + N2) ratio are pH, organic carbon and NO3 availability, water content, and O2 partial pressure (50). Soil pH is one of the most important factors influencing both denitrification and N2O production (43). In general, the denitrification rate increases with increasing pH values (up to the optimum pH) while, in contrast, the N2O/(N2O + N2) ratio decreases (50). This relationship has been characterized in laboratory experiments (9, 45), but it is not clear whether the same relationships exist in the field because of methodological limitations of in situ measurement of N2 emissions (16). Nevertheless, 15N tracing experiments based on the addition of a labeled denitrification substrate to soil offer a useful tool to quantify emissions of both N2O and N2 in situ (47, 49). Soil pH is also an important factor influencing denitrifier community composition (35, 39), which can be an important driver of denitrification activity and N2O emissions (5, 21). A recent study reported a negative relationship between the proportion of bacteria genetically capable of reducing N2O within the total bacterial community and the N2O/(N2O + N2) ratio, with both being strongly correlated with soil pH (38).The objective of the present study was to explore the effect of changes in soil pH on in situ N2O and N2 emissions, denitrifying enzyme activity (DEA), and potential N2O production. In addition, we also investigated whether differences in N fluxes could be related to changes in the size of the microbial community possessing the different denitrification genes. A field experiment was conducted using replicated grassland plots in which the soil pH was modified by addition of either acid or hydroxide to the soil. A 15N tracer method was used to provide information on N emissions. In addition to measuring potential denitrification activity, the size of the denitrifier community was determined by real-time PCR quantification of the denitrification genes.  相似文献   

7.
Rice paddy soil has been shown to have strong denitrifying activity. However, the microbial populations responsible for nitrate respiration and denitrification have not been well characterized. In this study, we performed a clone library analysis of >1,000 clones of the nearly full-length 16S rRNA gene to characterize bacterial community structure in rice paddy soil. We also identified potential key players in nitrate respiration and denitrification by comparing the community structures of soils with strong denitrifying activity to those of soils without denitrifying activity. Clone library analysis showed that bacteria belonging to the phylum Firmicutes, including a unique Symbiobacterium clade, dominated the clones obtained in this study. Using the template match method, several operational taxonomic units (OTUs), most belonging to the orders Burkholderiales and Rhodocyclales, were identified as OTUs that were specifically enriched in the sample with strong denitrifying activity. Almost one-half of these OTUs were classified in the genus Herbaspirillum and appeared >10-fold more frequently in the soils with strong denitrifying activity than in the soils without denitrifying activity. Therefore, OTUs related to Herbaspirillum are potential key players in nitrate respiration and denitrification under the conditions used.Rice is one of the most important agronomic plants in the world (20). More than 135 million ha are used for rice cultivation worldwide, 88% of which consists of paddy fields (i.e., flooded fields) (16). Since rice paddy soil has limited available oxygen, various anaerobic biochemical processes can occur, including methane production, Mn4+ and Fe3+ reduction, nitrate respiration, and denitrification.Denitrification is a microbial respiratory process during which soluble nitrogen oxides (NO3 and NO2) are reduced to gaseous products (NO, N2O, and N2) (14, 43). Reduction of nitrate (NO3) to nitrite (NO2) is part of the denitrification process; however, this reaction can also be performed by nondenitrifiers. Reduction of nitrate to nitrite as an end product is called nitrate respiration (43). The emission of N2O from rice paddy soils is less than that from upland crop fields (2), which is probably due to complete nitrate-nitrite reduction to N2, since rice paddy soil is known to have strong denitrifying activity (28). However, the microbes responsible for denitrification in rice paddy soil are not well known.Denitrifying ability is sporadically distributed among taxonomically diverse groups of bacteria, as well as some archaea and fungi (14, 33, 43). Therefore, it is difficult to identify denitrifying organisms based only on their 16S rRNA gene sequences (33). However, culture-independent 16S rRNA gene analysis can be used to identify microbial populations responsive to denitrification-inducing conditions if they are properly differentiated from background populations. The 16S rRNA gene can provide taxonomic information about organisms which cannot be obtained from analyses targeting nitrite reductase genes (nirS and nirK) alone (34).One approach to differentiate functionally active populations from background populations is to use stable-isotope probing (SIP) (35). SIP was previously used to identify succinate-assimilating bacterial populations under denitrifying conditions in rice paddy soil, using nitrate and succinate as the electron acceptor and donor, respectively (37). Although SIP analysis can provide solid evidence that links function with taxonomy, it requires assimilation of isotopically labeled substrates. This may limit the application of SIP in studies of dissimilatory processes, such as nitrate respiration and denitrification. For example, previous SIP studies targeted bacteria assimilating 13C-labeled acetate, methanol, or succinate under denitrifying conditions (13, 30, 37).Another approach is to detect specifically enriched microbial populations under certain conditions by comparative analysis of 16S rRNA gene sequences (9). This approach does not necessarily require addition of isotopically labeled substrates and therefore has the potential to identify microbes performing dissimilatory processes. Furthermore, the community structure of the total population can also be elucidated in this manner (10, 25, 36). However, the usefulness of comparative analysis of 16S rRNA gene sequences has not been thoroughly tested. In addition, this approach has not been used to study nitrate respirators and denitrifiers.Consequently, the objectives of this study were (i) to characterize the soil bacterial population in rice paddy soil by clone library analysis of >1,000 clones of the nearly full-length 16S rRNA gene and (ii) to identify active bacterial populations under denitrification-inducing conditions by comparing clone libraries.  相似文献   

8.
Wetlands are sources of denitrification-derived nitrous oxide (N2O). Thus, the denitrifier community of an N2O-emitting fen (pH 4.7 to 5.2) was investigated. N2O was produced and consumed to subatmospheric concentrations in unsupplemented anoxic soil microcosms. Total cell counts and most probable numbers of denitrifiers approximated 1011 cells·gDW−1 (where DW is dry weight) and 108 cells·gDW−1, respectively, in both 0- to 10-cm and 30- to 40-cm depths. Despite this uniformity, depth-related maximum reaction rate (vmax) values for denitrification in anoxic microcosms ranged from 1 to 24 and −19 to −105 nmol N2O h−1· gDW−1, with maximal values occurring in the upper soil layers. Denitrification was enhanced by substrates that might be formed via fermentation in anoxic microzones of soil. N2O approximated 40% of total nitrogenous gases produced at in situ pH, which was likewise the optimal pH for denitrification. Gene libraries of narG and nosZ (encoding nitrate reductase and nitrous oxide reductase, respectively) from fen soil DNA yielded 15 and 18 species-level operational taxonomic units, respectively, many of which displayed phylogenetic novelty and were not closely related to cultured organisms. Although statistical analyses of narG and nosZ sequences indicated that the upper 20 cm of soil contained the highest denitrifier diversity and species richness, terminal restriction fragment length polymorphism analyses of narG and nosZ revealed only minor differences in denitrifier community composition from a soil depth of 0 to 40 cm. The collective data indicate that the regional fen harbors novel, highly diverse, acid-tolerant denitrifier communities capable of complete denitrification and consumption of atmospheric N2O at in situ pH.Nitrous oxide (N2O) is a potent greenhouse gas with a global warming potential that is 300-fold higher than that of CO2, and its concentration increased from 270 ppb in 1750 to 319 ppb in 2005 (17). N2O can be produced in soils during denitrification, nitrification, the dissimilatory reduction of nitrate to nitrite and/or ammonium (hereafter referred to as dissimilatory nitrate reduction), or the chemical transformation of nitrite or hydroxylamine (5, 7, 49). The percentage of N2O produced in any of these processes is variable, depending mainly on the redox potential, pH, and C/N ratio (49). In anoxic ecosystems such as waterlogged soils, most of the N2O is considered to be denitrification derived (7, 9). Complete denitrification is the sequential reduction of nitrate to dinitrogen (N2) via nitrite, nitric oxide (NO), and N2O (75). The main product of denitrification varies with the organism and in situ conditions and is usually either N2O or N2 (68). N2O can occur as a by-product during dissimilatory nitrate reduction when accumulated nitrite interacts with nitrate reductase to form N2O (59). The production of N2O by dissimilatory nitrate reducers is favored in environments with large amounts of readily available organic carbon (65). Thus, their contribution to nitrate-dependent production of N2O in soils is likely insignificant compared to that of denitrifiers.The oxidoreductases involved in denitrification are termed dissimilatory nitrate reductase (Nar, encoded by narGHJI, or Nap, encoded by napEDABC), nitrite reductase (Nir, encoded by nirK and nirS), NO reductase (cNor and qNor, encoded by norBC and norB, respectively), and N2O reductase (Nos, encoded by nosZ) (75). Nitrate reductase is also found in dissimilatory nitrate reducers (60). narG can therefore be used as a molecular marker to assess both denitrifiers and dissimilatory nitrate reducers, whereas nosZ is specific for the assessment of denitrifiers (25, 43, 48).Denitrification in soils is regulated by temperature, pH, substrate (i.e., carbon) availability, and water content (10, 24, 66). Although denitrification increases with increasing temperature, it can still occur at temperatures below 0°C (10, 24). Low temperatures appear to limit the activity of N2O reductase more severely than other enzymes involved in denitrification and thus yield higher relative amounts of denitrification-derived N2O (24). Although denitrification activity usually decreases under acidic conditions, the relative percentage of N2O to total denitrification-derived nitrogenous gases increases with increasing acidity, a result attributed to the sensitivity of N2O reductase to low pH (27, 70). However, denitrifier communities can be adapted to the in situ pH of the system (40, 58, 73).Wetlands are ecosystems in which denitrification is likely a dominant source of emitted N2O (7, 44, 45). The identification and analysis of main drivers for N2O production (i.e., the microbiota catalyzing N2O production and consumption) is thus of major concern in such environments. Fens are specialized wetlands characterized by soil acidity (67). However, information on acid-tolerant denitrifier communities of such wetlands is scarce. It is hypothesized that fens harbor a diverse, hitherto unknown, denitrifier community that is adapted to in situ conditions and associated with N2O fluxes (i.e., fen denitrifiers are acid tolerant and have a high affinity for nitrate and N2O). Thus, the main objectives of the present study were to evaluate the capacities of denitrifier communities of an N2O-emitting fen (20) to produce or consume N2O and to determine if a novel and diverse denitrifier community was associated with these capacities.  相似文献   

9.
10.
Although denitrification or nitrate respiration has been found among a few eukaryotes, its phylogenetic relationship with the bacterial system remains unclear because orthologous genes involved in the bacterial denitrification system were not identified in these eukaryotes. In this study, we isolated a gene from the denitrifying fungus Fusarium oxysporum that is homologous to the bacterial nirK gene responsible for encoding copper-containing nitrite reductase (NirK). Characterization of the gene and its recombinant protein showed that the fungal nirK gene is the first eukaryotic ortholog of the bacterial counterpart involved in denitrification. Additionally, recent genome analyses have revealed the occurrence of nirK homologs in many fungi and protozoa, although the denitrifying activity of these eukaryotes has never been examined. These eukaryotic homolog genes, together with the fungal nirK gene of F. oxysporum, are grouped in the same branch of the phylogenetic tree as the nirK genes of bacteria, archaea, and eukaryotes, implying that eukaryotic nirK and its homologs evolved from a single ancestor (possibly the protomitochondrion). These results show that the fungal denitrifying system has the same origin as its bacterial counterpart.Denitrification plays an important role in the global nitrogen cycle and reduces nitrate (NO3) and/or nitrite (NO2) to a gaseous form of nitrogen, generally to dinitrogen (N2) or nitrous oxide (N2O) (27). It typically follows four reduction stages, NO3 → NO2 → NO → N2O → N2, each of which is catalyzed by a specific reductase: dissimilatory NO3 reductase (dNaR), dissimilatory NO2 reductase (dNiR), nitric oxide (NO) reductase (NoR), and N2O reductase, respectively. These enzymes receive electrons from a respiratory chain functioning as a “terminal reductase.” Thus, denitrification exhibits a physiological significance in its ability to anaerobically respire through the processes of nitrate respiration, nitrite respiration, and so forth. Denitrification was previously thought to be a characteristic of bacteria; however, similar reactions have been found to occur in a few eukaryotes and archaea (6, 27). Eukaryotic nitrate respiration was first found in protozoa that reside in an anaerobic freshwater habitat (8). The organism particularly reduces NO3 to NO2 in a single step, a process which recovers dNaR activity in the mitochondrial fraction but does not result in denitrification. Eukaryotic denitrification was first found to occur among fungi (19, 20), which generally form N2O from NO3 or NO2. Recently, eukaryotic denitrification was also found in a benthic foraminifer that forms N2 from NO3 (18). The fungal denitrification system localizes in the mitochondria and couples to the mitochondrial electron transport chain to produce ATP (12, 21), thus exhibiting properties similar to those of the bacterial systems in its ability to respire anaerobically. Moreover, the mechanism of anaerobic respiration in the “aerobic” organelle of eukaryotes (mitochondrion) evokes interest regarding the origin and evolution of the mitochondrion.The main components of the fungal denitrifying system, the dNaR, dNiR, and NoR proteins, were either completely or partially purified from Fusarium oxysporum. Fungal NoR of the cytochrome P450 (P450) type, referred to as P450nor (CYP55) (11, 16), is a distinct species of bacterial cytochrome cb-type NoR. By contrast, the previously isolated fungal dNiR protein is a copper-containing type (NirK) that closely resembles its bacterial counterpart (13). Furthermore, dNaR activity partially purified from the mitochondrial membrane fraction showed that fungal dNaR possibly resembles its bacterial counterpart, NarGHI (12, 23). Therefore, while a portion of the fungal system appears to resemble its bacterial counterpart, the phylogenetic relationship between the fungal and bacterial denitrification systems remained unclear because the genes of the fungal components (dNaR and dNiR) have not been sequenced.Recent genome analyses have revealed the presence of nirK homolog genes in many eukaryotes (fungi and protozoa), a finding consistent with our previous findings on the isolation of the fungal NirK protein (13). Therefore, whether these eukaryotes containing the nirK homolog gene exhibit denitrification activity and whether the denitrifying fungus F. oxysporum really contains a nirK gene deserve a great deal of attention. To address this issue, we used the suppression subtractive hybridization (SSH) technique (7) and succeeded in isolating the nirK gene from the denitrifying fungus F. oxysporum.  相似文献   

11.
12.
13.
14.
Bacterial anaerobic ammonium oxidation (anammox) is an important process in the marine nitrogen cycle. Because ongoing eutrophication of coastal bays contributes significantly to the formation of low-oxygen zones, monitoring of the anammox bacterial community offers a unique opportunity for assessment of anthropogenic perturbations in these environments. The current study used targeting of 16S rRNA and hzo genes to characterize the composition and structure of the anammox bacterial community in the sediments of the eutrophic Jiaozhou Bay, thereby unraveling their diversity, abundance, and distribution. Abundance and distribution of hzo genes revealed a greater taxonomic diversity in Jiaozhou Bay, including several novel clades of anammox bacteria. In contrast, the targeting of 16S rRNA genes verified the presence of only “Candidatus Scalindua,” albeit with a high microdiversity. The genus “Ca. Scalindua” comprised the apparent majority of active sediment anammox bacteria. Multivariate statistical analyses indicated a heterogeneous distribution of the anammox bacterial assemblages in Jiaozhou Bay. Of all environmental parameters investigated, sediment organic C/organic N (OrgC/OrgN), nitrite concentration, and sediment median grain size were found to impact the composition, structure, and distribution of the sediment anammox bacterial community. Analysis of Pearson correlations between environmental factors and abundance of 16S rRNA and hzo genes as determined by fluorescent real-time PCR suggests that the local nitrite concentration is the key regulator of the abundance of anammox bacteria in Jiaozhou Bay sediments.Anaerobic ammonium oxidation (anammox, NH4+ + NO2 → N2 + 2H2O) was proposed as a missing N transformation pathway decades ago. It was found 20 years later to be mediated by bacteria in artificial environments, such as anaerobic wastewater processing systems (see reference 32 and references therein). Anammox in natural environments was found even more recently, mainly in O2-limited environments such as marine sediments (28, 51, 54, 67, 69) and hypoxic or anoxic waters (10, 25, 39-42). Because anammox may remove as much as 30 to 70% of fixed N from the oceans (3, 9, 64), this process is potentially as important as denitrification for N loss and bioremediation (41, 42, 73). These findings have significantly changed our understanding of the budget of the marine and global N cycles as well as involved pathways and their evolution (24, 32, 35, 72). Studies indicate variable anammox contributions to local or regional N loss (41, 42, 73), probably due to distinct environmental conditions that may influence the composition, abundance, and distribution of the anammox bacteria. However, the interactions of anammox bacteria with their environment are still poorly understood.The chemolithoautotrophic anammox bacteria (64, 66) comprise the new Brocadiaceae family in the Planctomycetales, for which five Candidatus genera have been described (see references 32 and 37 and references therein): “Candidatus Kuenenia,” “Candidatus Brocadia,” “Candidatus Scalindua,” “Candidatus Anammoxoglobus,” and “Candidatus Jettenia. Due to the difficulty of cultivation and isolation, anammox bacteria are not yet in pure culture. Molecular detection by using DNA probes or PCR primers targeting the anammox bacterial 16S rRNA genes has thus been the main approach for the detection of anammox bacteria and community analyses (58). However, these studies revealed unexpected target sequence diversity and led to the realization that due to biased coverage and specificity of most of the PCR primers (2, 8), the in situ diversity of anammox bacteria was likely missed. Thus, the use of additional marker genes for phylogenetic analysis was suggested in hopes of better capturing the diversity of this environmentally important group of bacteria. By analogy to molecular ecological studies of aerobic ammonia oxidizers, most recent studies have attempted to include anammox bacterium-specific functional genes. All anammox bacteria employ hydrazine oxidoreductase (HZO) (= [Hzo]3) to oxidize hydrazine to N2 as the main source for a useable reductant, which enables them to generate proton-motive force for energy production (32, 36, 65). Phylogenetic analyses of Hzo protein sequences revealed three sequence clusters, of which the cladistic structure of cluster 1 is in agreement with the anammox bacterial 16S rRNA gene phylogeny (57). The hzo genes have emerged as an alternative phylogenetic and functional marker for characterization of anammox bacterial communities (43, 44, 57), allowing the 16S rRNA gene-based investigation methods to be corroborated and improved.The contribution of anammox to the removal of fixed N is highly variable in estuarine and coastal sediments (50). For instance, anammox may be an important pathway for the removal of excess N (23) or nearly negligible (48, 54, 67, 68). This difference may be attributable to a difference in the structure and composition of anammox bacterial communities, in particular how the abundance of individual cohorts depends on particular environmental conditions. Anthropogenic disturbance with variable source and intensity of eutrophication and pollution may further complicate the anammox bacterium-environment relationship.Jiaozhou Bay is a large semienclosed water body of the temperate Yellow Sea in China. Eutrophication has become its most serious environmental problem, along with red tides (harmful algal blooms), species loss, and contamination with toxic chemicals and harmful microbes (14, 15, 21, 61, 71). Due to different sources of pollution and various levels of eutrophication across Jiaozhou Bay (mariculture, municipal and industrial wastewater, crude oil shipyard, etc.), a wide spectrum of environmental conditions may contribute to a widely varying community structure of anammox bacteria. This study used both 16S rRNA and hzo genes as targets to measure their abundance, diversity, and spatial distribution and assess the response of the resident anammox bacterial community to different environmental conditions. Environmental factors with potential for regulating the sediment anammox microbiota are discussed.  相似文献   

15.
Dissimilatory NO3 reduction in sediments is often measured in bulk incubations that destroy in situ gradients of controlling factors such as sulfide and oxygen. Additionally, the use of unnaturally high NO3 concentrations yields potential rather than actual activities of dissimilatory NO3 reduction. We developed a technique to determine the vertical distribution of the net rates of dissimilatory nitrate reduction to ammonium (DNRA) with minimal physical disturbance in intact sediment cores at millimeter-level resolution. This allows DNRA activity to be directly linked to the microenvironmental conditions in the layer of NO3 consumption. The water column of the sediment core is amended with 15NO3 at the in situ 14NO3 concentration. A gel probe is deployed in the sediment and is retrieved after complete diffusive equilibration between the gel and the sediment pore water. The gel is then sliced and the NH4+ dissolved in the gel slices is chemically converted by hypobromite to N2 in reaction vials. The isotopic composition of N2 is determined by mass spectrometry. We used the combined gel probe and isotopic labeling technique with freshwater and marine sediment cores and with sterile quartz sand with artificial gradients of 15NH4+. The results were compared to the NH4+ microsensor profiles measured in freshwater sediment and quartz sand and to the N2O microsensor profiles measured in acetylene-amended sediments to trace denitrification.Nitrate accounts for the eutrophication of many human-affected aquatic ecosystems (19, 21). Sediment bacteria may mitigate NO3 pollution by denitrification and anaerobic ammonium oxidation (anammox), which produce N2 (13, 18). However, inorganic nitrogen is retained in aquatic ecosystems when sediment bacteria reduce NO3 to NH4+ by dissimilatory nitrate reduction to ammonium (DNRA) (5, 12, 16, 39). Hence, DNRA contributes to rather than counteracts eutrophication (23). DNRA may be the dominant pathway of dissimilatory NO3 reduction in sediments that are rich in electron donors, such as labile organic carbon and sulfide (4, 8, 17, 38, 55). High rates of DNRA are thus found in sediments affected by coastal aquaculture (8, 36) and settling algal blooms (16).DNRA, denitrification, and the chemical factors that control the partitioning between them (e.g., sulfide) should ideally be investigated in undisturbed sediments. The redox stratification of sediments involves vertical concentration gradients of pore water solutes. These gradients are often very steep, and their measurement requires high-resolution techniques, such as microsensors (26, 42) and gel probes (9, 54). If, for instance, the influence of sulfide on DNRA and denitrification is to be investigated, one wants to know exactly the sulfide concentration in the layers of DNRA and denitrification activity, as well as the flux of sulfide into these layers. This information can easily be obtained using H2S and pH microsensors (22, 43). It is less trivial to determine the vertical distribution of DNRA and denitrification activity in undisturbed sediments. Denitrification activity can be traced using a combination of the acetylene inhibition technique (51) and N2O microsensors (1). Acetylene inhibits the last step of denitrification, and therefore, N2O accumulates in the layer of denitrification activity (44). This method underestimates the denitrification activity in sediments with high rates of coupled nitrification-denitrification because acetylene also inhibits nitrification (50).The vertical distribution of DNRA activity in undisturbed sediment has, to the best of our knowledge, never been determined; thus, the microenvironmental conditions in the layer of DNRA activity remain unknown. Until now, the influence of chemical factors on DNRA and denitrification in sediments has been assessed by slurry incubations (4, 12, 30), by flux measurements with sealed sediment cores (7, 47) or flowthrough sediment cores (16, 27, 37), and in one case, in reconstituted sediment cores sliced at centimeter-level resolution (39). Here, we present a new method, the combined gel probe and isotope labeling technique, to determine the vertical distribution of the net rates of DNRA in sediments. The sediments remain largely undisturbed and the NO3 amendments are within the range of in situ concentrations. The DNRA measurements can be related to the microprofiles of potential influencing factors measured in close vicinity of the gel probe. This allows DNRA activity to be directly linked with the microenvironmental conditions in the sediment.  相似文献   

16.
17.
18.
19.
Knowing spatial patterns of functional microbial guilds can increase our understanding of the relationships between microbial community ecology and ecosystem functions. Using geostatistical modeling to map spatial patterns, we explored the distribution of the community structure, size, and activity of one functional group in N cycling, the denitrifiers, in relation to 23 soil parameters over a 44-ha farm divided into one organic and one integrated crop production system. The denitrifiers were targeted by the nirS and nirK genes that encode the two mutually exclusive types of nitrite reductases, the cd1 heme-type and copper reductases, respectively. The spatial pattern of the denitrification activity genes was reflected by the maps of the abundances of nir genes. For the community structure, only the maps of the nirS community were related to the activity. The activity was correlated with nitrate and dissolved organic nitrogen and carbon, whereas the gene pools for denitrification, in terms of size and composition, were influenced by the soil structure. For the nirS community, pH and soil nutrients were also important in shaping the community. The only unique parameter related to the nirK community was the soil Cu content. However, the spatial pattern of the nirK denitrifiers corresponded to the division of the farm into the two cropping systems. The different community patterns, together with the spatial distribution of the nirS/nirK abundance ratio, suggest habitat selection on the nirS- and nirK-type denitrifiers. Our findings constitute a first step in identifying niches for denitrifiers at scales relevant to land management.Soil microorganisms are abundant and diverse (46), drive key processes in biogeochemical cycles, and, thus, play crucial roles in ecosystem functioning (2). They are not randomly distributed but exhibit spatial patterns at different scales (26). Spatial patterns ranging from the micrometer up to the meter scale have been reported (19, 20, 32, 37), and an understanding of such patterns can give clues to how microbial communities are generated and maintained (17). Spatial patterns of microorganisms at the field and landscape scales warrant special attention, since they could be associated with land use and aid in creating knowledge-based management strategies for agricultural production (5, 42). However, our understanding of key habitat-selective factors is limited, and few studies have specified which factors influence the spatial patterns of soil microbial communities at larger scales. Lauber et al. (30) recently demonstrated that pH could predict the community composition of soil bacteria at the continental scale. The importance of pH as a key edaphic driver of bacterial community structure has also been shown in other studies (11, 47). Another major, but complex, factor pointed out in a few studies is the soil type (4, 5, 14). Most studies have included only a limited number of properties that are easy to measure; most often, carbon and nitrogen pools and soil physical factors have been neglected in microbial community ecology. Since these factors delineate soil oxygen and water content, they may exert a stronger impact on microbial communities than other soil resources.Reports on the field or landscape scale spatial distribution of soil bacteria have had a taxon-centered perspective at either the species or total-community level, but there is emerging interest in the biogeography of functional traits possessed by microorganisms (18). Bacterial species composition is likely important for soil ecosystem functions, but species affiliation rarely predicts in which way. In addition, the fuzzy species concept of bacteria makes it all the more difficult to link species to niches. Analysis of functional guilds, i.e., assemblages of populations sharing certain traits, can bridge this gap, and one guild of global concern that has been suggested and recently used as a model in functional ecology is the denitrifiers (39, 49). Denitrification is an anaerobic respiration pathway during which NO3 is reduced to N2 by a wide range of unrelated taxa. The process is an essential route for N loss from agricultural soil and a major source of the greenhouse gas N2O. It was recently shown that the spatial distribution of the relative abundances of denitrifiers with the genetic capacity to perform the last step in the denitrification pathway, reduction of N2O to N2, is linked to areas with high denitrification rates and low N2O emissions (38). Adding field scale predicted patterns of the denitrifier community structure to the abundance and activity would not only give insight into the mechanisms shaping the community, but also deepen our understanding of the relationships between the ecology of denitrifiers, N loss, and the agroecosystems'' impact on climate change.We hypothesize that spatial autocorrelations of the structures, sizes, and activities of communities of denitrifying bacteria is governed by soil-based resources at a scale compatible with land management. To test this, and to elucidate the effects of crop production systems and the importance of soil physical and chemical factors in the denitrifying community, we explored the spatial distribution of community structure, size, and activity in relation to 23 soil parameters at a 44-ha farm divided into one organic and one integrated crop production system. The denitrifier community was described in terms of the signature genes that encode the two different types of nitrite reductases in the denitrification pathway, the cd1 heme-type reductase (NirS), encoded by the nirS gene, and the copper oxidoreductase (NirK), encoded by nirK. The spatial patterns were mapped by geostatistical modeling, and correlation structures were explored.  相似文献   

20.
A molecular analysis of betaproteobacterial ammonia oxidizers and a N2O isotopomer analysis were conducted to study the sources of N2O emissions during the cow manure composting process. Much NO2-N and NO3-N and the Nitrosomonas europaea-like amoA gene were detected at the surface, especially at the top of the composting pile, suggesting that these ammonia-oxidizing bacteria (AOB) significantly contribute to the nitrification which occurs at the surface layer of compost piles. However, the 15N site preference within the asymmetric N2O molecule (SP = δ15Nα − δ15Nβ, where 15Nα and 15Nβ represent the 15N/14N ratios at the center and end sites of the nitrogen atoms, respectively) indicated that the source of N2O emissions just after the compost was turned originated mainly from the denitrification process. Based on these results, the reduction of accumulated NO2-N or NO3-N after turning was identified as the main source of N2O emissions. The site preference and bulk δ15N results also indicate that the rate of N2O reduction was relatively low, and an increased value for the site preference indicates that the nitrification which occurred mainly in the surface layer of the pile partially contributed to N2O emissions between the turnings.The very sensitive greenhouse gas nitrous oxide (N2O) has a 296 times higher impact than CO2 (39) and is also responsible for ozone depletion (10). Agricultural activities such as the use of nitrate fertilizers, livestock production, and manure management, including composting, are known to be important sources of N2O emissions (18). To devise a strategy to mitigate N2O emissions, it is essential to understand its sources in detail. However, the sources of N2O emissions during the composting process are still largely unclear.In the composting process, a part of NH4+-N is known to be processed through nitrification-denitrification and emitted as N2 and N2O. Nitrous oxide is known to be generated through both the nitrification and denitrification processes as intermediate products or by-products. Nitrous oxide emission is a very complex process because denitrifying bacteria are phylogenetically diverse (60), and nitrifiers are also known to utilize the denitrification process even under aerobic conditions (42). It is thus very difficult to estimate the relative contributions of nitrification and denitrification in actual N2O emissions from the environment. Until now, there has been insufficient knowledge about the relative contributions of these processes to N2O emissions during the animal manure composting process. Measurement of the actual contributions of N2O emissions from compost piles in the field is therefore critical to establishing a strategy of mitigating N2O emissions.Recently, a high-precision analytical technique for determining intramolecular 15N site preference in asymmetric molecules of N2O was developed (47). Since N2O has two N atoms within the molecule (central and outer N), distribution of a stable isotope, 15N, results in the distribution of three isotopomers, such as 15N15NO, 15N14NO, and 14N15NO. By using this newly developed innovative technique, the latter two types of molecules, which exist abundantly in the environment, can be individually measured. The difference in δ15N between δ15Nα and δ15Nβ is the so-called site preference (SP = δ15Nα − δ15Nβ, where 15Nα and 15Nβ represent the 15N/14N ratios at the center and end sites of the nitrogen atoms, respectively). The site preference enabled us to identify the source and sinks of N2O in the environment (48, 49, 50, 56). Using this technique, Sutka et al. (44) found that the site preference for N2O from hydroxylamine oxidation (∼33‰) and nitrite reduction (∼0‰) differs in a pure culture study and noted that this difference can be used to distinguish the relative contributions of nitrification and denitrification sources to N2O emissions. There have still been only several reported studies which applied this measurement technique to field N2O samples (48, 53) or referred to the relative contributions of nitrification and denitrification. To our knowledge, the present study is the first to apply this isotopomer analysis technique to the determination of N2O sources in the composting process. We specifically used this technique to understand the actual contributions of nitrification and denitrification to N2O emissions during the cow manure composting process.Ammonia oxidation, the conversion of ammonium to nitrite via hydroxylamine, is an initial step of the nitrification-denitrification process and is critical to the nitrogen cycle in the terrestrial environment (4, 24). In the nitrification process, N2O is generated as a by-product when ammonia oxidizers convert hydroxylamine to nitrite (35). Since NO2-N and NO3-N accumulate in the latter stages of the composting process (29, 30), it is obvious that nitrifiers are active in compost piles. Therefore, it is important to clarify the role and significance of ammonia oxidizers in N2O emissions during the composting process. However, since the pure culture isolation method is so difficult and time-consuming, little is known about these ammonia oxidizers. A molecular approach based on PCR has been recently developed and has to date been used to target the ammonia monooxygenase gene (amoA) or 16S rRNA gene of betaproteobacterial ammonia oxidizers in soil, wetlands, and marine sediments (2, 3, 6, 7, 13, 32, 52). Using these techniques, substantial information about uncultured ammonia-oxidizing bacteria (AOB) that are partially or wholly responsible for nitrification in the environment will become available. Since the microbial community drastically changes through the composting process (19, 29), and a high accumulation of nitrite or nitrate will occur, especially in the latter half of the process (30), we continuously sampled and analyzed the diversity and abundance of AOB throughout the process. Our objectives in this study were to elucidate the sources of N2O emissions during the cow manure composting process by combining the isotopomer analysis and molecular analysis of betaproteobacterial AOB.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号