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 15 毫秒
1.
2.
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   

3.
4.
5.
In contrast to most denitrifiers studied so far, Pseudomonas stutzeri TR2 produces low levels of nitrous oxide (N2O) even under aerobic conditions. We compared the denitrification activity of strain TR2 with those of various denitrifiers in an artificial medium that was derived from piggery wastewater. Strain TR2 exhibited strong denitrification activity and produced little N2O under all conditions tested. Its growth rate under denitrifying conditions was near comparable to that under aerobic conditions, showing a sharp contrast to the lower growth rates of other denitrifiers under denitrifying conditions. Strain TR2 was tolerant to toxic nitrite, even utilizing it as a good denitrification substrate. When both nitrite and N2O were present, strain TR2 reduced N2O in preference to nitrite as the denitrification substrate. This bacterial strain was readily able to adapt to denitrifying conditions by expressing the denitrification genes for cytochrome cd1 nitrite reductase (NiR) (nirS) and nitrous oxide reductase (NoS) (nosZ). Interestingly, nosZ was constitutively expressed even under nondenitrifying, aerobic conditions, consistent with our finding that strain TR2 preferred N2O to nitrite. These properties of strain TR2 concerning denitrification are in sharp contrast to those of well-characterized denitrifiers. These results demonstrate that some bacterial species, such as strain TR2, have adopted a strategy for survival by preferring denitrification to oxygen respiration. The bacterium was also shown to contain the potential to reduce N2O emissions when applied to sewage disposal fields.Wastewater treatment processes produce one of the major greenhouse effect gases, nitrous oxide (N2O) (7, 25, 30). The global warming potential of N2O relative to that of carbon dioxide (CO2) is 298 for a 100-year time horizon, and its concentration in the atmosphere continues to increase by about 0.26% per year (9). Nitrogen removal in wastewater treatment plants is essentially based on the activities of nitrifying and denitrifying microorganisms, both of which are inhabitants of activated sludge. Nitrifying bacteria aerobically oxidize ammonium to nitrite (NO2) and nitrate (NO3), which are then reduced anaerobically by denitrifying bacteria to gaseous nitrogen forms, such as N2O and dinitrogen (N2). It has long been known that N2O can be produced during both nitrification and denitrification processes of wastewater treatment (3, 19, 23), but the cause of N2O emission during the nitrification process was not clear. We recently showed, however, using activated sludge grown under conditions that mimicked a piggery wastewater disposal, that N2O emission during the nitrification process depends on denitrification by ammonia-oxidizing bacteria (Nitrosomonas) (18). On the other hand, it is believed that denitrifying bacteria produce N2O as a by-product when anaerobiosis is insufficient during the denitrification process, because N2O reductase is the enzyme that is most sensitive to oxygen (6). Piggery wastewater, in particular, contains a high concentration of ammonia, and N2O emission tends to take place during the nitrogen removal process (5, 10). Experiments on the removal of ammonia and organic carbon by the aerobic denitrifier Pseudomonas stutzeri SU2 (24) and the heterotrophic nitrifier-aerobic denitrifier Alcaligenes faecalis no. 4 (16, 17) have been reported as examples of bioaugmentation in piggery wastewater treatment. Reduction of N2O emissions from pig manure compost by addition of nitrite-oxidizing bacteria has also been reported (11). However, there have been no reports of methods for reducing N2O emissions by bioaugmentation using aerobic denitrifying bacteria.Takaya et al. isolated the aerobic denitrifying bacterium Pseudomonas stutzeri TR2 (26). The denitrification activity of strain TR2 was monitored in batch and continuous cultures, using denitrification and artificial wastewater media, and the strain was found to keep a distinct activity (producing N2 from NO3) and to produce a very low level of N2O at a dissolved oxygen (O2) concentration of 1.25 mg liter−1. Therefore, strain TR2 should be useful in the future for reducing N2O emissions from wastewater treatment plants by bioaugmentation. To investigate the feasibility of using strain TR2 for future application to wastewater treatment processes, we examined its denitrification activity, N2O production, growth rate, and expression of denitrifying genes in batch cultures, using a medium that mimics the composition found in nitrogen removal wastewater plants. Comparison of the properties of strain TR2 with those of well-characterized denitrifying bacteria revealed characteristics of the strain that favor denitrification, although it can also respire oxygen.  相似文献   

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.
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   

8.
9.
10.
11.
Wetland mitigation is implemented to replace ecosystem functions provided by wetlands; however, restoration efforts frequently fail to establish equivalent levels of ecosystem services. Delivery of microbially mediated ecosystem functions, such as denitrification, is influenced by both the structure and activity of the microbial community. The objective of this study was to compare the relationship between soil and vegetation factors and microbial community structure and function in restored and reference wetlands within a mitigation bank. Microbial community composition was assessed using terminal restriction fragment length polymorphism targeting the 16S rRNA gene (total bacteria) and the nosZ gene (denitrifiers). Comparisons of microbial function were based on potential denitrification rates. Bacterial community structures differed significantly between restored and reference wetlands; denitrifier community assemblages were similar among reference sites but highly variable among restored sites throughout the mitigation bank. Potential denitrification was highest in the reference wetland sites. These data demonstrate that wetland restoration efforts in this mitigation bank have not successfully restored denitrification and that differences in potential denitrification rates may be due to distinct microbial assemblages observed in restored and reference (natural) wetlands. Further, we have identified gradients in soil moisture and soil fertility that were associated with differences in microbial community structure. Microbial function was influenced by bacterial community composition and soil fertility. Identifying soil factors that are primary ecological drivers of soil bacterial communities, especially denitrifying populations, can potentially aid the development of predictive models for restoration of biogeochemical transformations and enhance the success of wetland restoration efforts.Wetlands provide more ecosystem services (e.g., flood control, water purification, nutrient cycling, and habitat for wildlife) per hectare than any other ecosystem (16). Riparian wetlands, in particular, are sites of intense biogeochemical activity and play an important role in improving water quality, recycling nutrients, and detoxifying chemicals (41). Changing patterns of land use over the last century have resulted in the loss of over half of the wetlands in the contiguous United States (17) and about 60% of wetlands in the Midwestern United States (82). The loss of ecosystem services through conversion of wetlands to alternative (primarily agricultural) land uses exacerbates nutrient pollution and eutrophication of downstream ecosystems (57). Declines in wetland acreage have continued despite a federal policy goal of no-net-loss of wetland acreage and function adopted in 1990 (7, 55). Wetland mitigation projects provide compensation for impacted wetlands and aim to replace the critical functions provided by wetlands. Despite decades of wetland mitigation, however, restoration efforts frequently fail to reestablish desired levels of ecosystem services. Restoration outcomes remain uncertain, and more information is necessary in order to improve monitoring and assessment of wetland development (13, 18, 50, 80).One approach to wetland compensation is through mitigation banks. These sites are areas that are restored, established, enhanced, or preserved for replacement of wetlands that will be affected by future land use change. Mitigation banks are considered “third-party” compensatory mitigation, where the permittee (e.g., developer planning to destroy a wetland) is responsible for purchasing wetland credits in acreage, but the wetland bank is established and managed by another party (24). Wetland mitigation banks have unique characteristics that distinguish them from smaller individual restoration projects (7, 69, 81). Due to their size, wetland mitigation banks are especially heterogeneous and may have a great deal of within-site variability in hydrology and nutrient status, making it challenging to implement a single restoration design. Thus, wetland mitigation banks require intense management and monitoring for improved success (7, 69, 81).Restoration efforts such as mitigation banks aim to replace chemical, physical, and biological ecosystem functions of wetlands that have been lost through anthropogenic disturbance (24). Monitoring of wetland mitigation sites has largely focused on measures of macro-scale community structure (e.g., vegetation surveys) (52) along with measures of hydrology and soil type (24). Measurement of vegetation is a common proxy for wetland performance but does not provide an accurate assessment of wetland function (6, 52). Quantitative assessment is achievable, however, for ecosystem services such as water quality improvement through nitrate removal, where well-characterized microbial mechanisms underlie denitrification processes.The link between microbial community structure and function in a restoration context is a topic of current interest (33). Relating microbial community composition and dynamics to chemical, physical, and biological variables can help to reveal important ecological drivers of microbial communities and their activities (26, 35, 42). Conserved bacterial functional genes related to specific biogeochemical transformations allow evaluation of the community structure of microbial populations directly involved in these processes (49, 60, 63, 77, 79). Assessing the diversity of microorganisms that are specifically involved in denitrification is possible through amplification of the nosZ gene, which encodes the catalytic subunit of nitrous oxide reductase, the enzyme responsible for the final step of denitrification (60, 63, 66). Phylogenetically diverse microorganisms can carry out denitrification though the majority of previously described denitrifiers belong to subphyla within the Proteobacteria (53, 56, 60, 61). Denitrification is a facultative process that occurs only under anaerobic conditions (53, 75). Complete denitrification to N2 is more prevalent in anaerobic, saturated wetland ecosystems (14, 76), and incomplete denitrification to N2O is the less desirable, more common endpoint of denitrification under more aerobic, drier conditions (14, 62). While the environmental factors (e.g., oxygen, carbon, nitrate, and pH) that influence bulk denitrification rates have been well characterized (31, 72), the influence of these factors on the composition of denitrifier communities, particularly in a restoration context, is unclear. Understanding the relationship between the microbial populations responsible for nitrogen transformations and easily measured environmental parameters (e.g., soil chemical and physical measures) could lead to assessment metrics that are linked directly to ecosystem functions such as denitrification and bridge the current gap in functional assessment methods (36, 60, 70).The objectives of this study were (i) to compare the microbial and plant community composition in restored wetlands to the composition in adjacent reference floodplain forest wetlands; (ii) to assess the relationship between microbial community composition (based on terminal restriction fragment length polymorphism [T-RFLP]) and potential denitrification activity throughout the mitigation bank; and (iii) to examine soil factors correlated with microbial community composition using both phylogenetic and functional gene markers. As soil environmental conditions affect microbial community structure and activity, we expected that sites where wetland hydrology and soil chemistry have been successfully restored would harbor microbial assemblages that are similar in composition and denitrification function to those observed in reference wetlands within this mitigation bank.  相似文献   

12.
13.
14.
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   

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.
Since enterohemorrhagic Escherichia coli (EHEC) isolates of serogroup O156 have been obtained from human diarrhea patients and asymptomatic carriers, we studied cattle as a potential reservoir for these bacteria. E. coli isolates serotyped by agglutination as O156:H25/H−/Hnt strains (n = 32) were isolated from three cattle farms during a period of 21 months and characterized by rapid microarray-based genotyping. The serotyping by agglutination of the O156 isolates was not confirmed in some cases by the results of DNA-based serotyping as only 25 of the 32 isolates were conclusively identified as O156:H25. In the multilocus sequence typing (MLST) analysis, all EHEC O156:H25 isolates were characterized as sequence type 300 (ST300) and ST688, which differ by a single-nucleotide exchange in the purA gene. Oligonucleotide microarrays allow simultaneous detection of a wider range of EHEC-associated and other E. coli virulence markers than other methods. All O156:H25 isolates showed a wide spectrum of virulence factors typical for EHEC. The stx1 genes combined with the EHEC hlyA (hlyAEHEC) gene, the eae gene of the ζ subtype, as well as numerous other virulence markers were present in all EHEC O156:H25 strains. The behavior of eight different cluster groups, including four that were EHEC O156:H25, was monitored in space and time. Variations in the O156 cluster groups were detected. The results of the cluster analysis suggest that some O156:H25 strains had the genetic potential for a long persistence in the host and on the farm, while other strains did not. As judged by their pattern of virulence markers, E. coli O156:H25 isolates of bovine origin may represent a considerable risk for human infection. Our results showed that the miniaturized E. coli oligonucleotide arrays are an excellent tool for the rapid detection of a large number of virulence markers.Shiga toxin-producing Escherichia coli (STEC) strains comprise a group of zoonotic enteric pathogens (45). In humans, infections with some STEC serotypes may result in hemorrhagic or nonhemorrhagic diarrhea, which can be complicated by the hemolytic uremic syndrome (HUS) (32). These STEC strains are also designated enterohemorrhagic Escherichia coli (EHEC). Consequently, EHEC strains represent a subgroup of STEC with high pathogenic potential for humans. Although E. coli O157:H7 is the most frequent EHEC serotype implicated in HUS, other serotypes can also cause this complication. Non-O157:H7 EHEC strains including serotypes O26:H11/H−, O103:H2/H−, O111:H8/H10/H−, and O145:H28/H25/H− and sorbitol-fermenting E. coli O157:H− isolates are present in about 50% of stool cultures from German HUS patients (10, 42). However, STEC strains that cause human infection belong to a large number of E. coli serotypes, although a small number of STEC isolates of serogroup O156 were associated with human disease (7). Strains of the serotypes O156:H1/H8/H21/H25 were found in human cases of diarrhea or asymptomatic infections (9, 22, 25, 26). The detection of STEC of serogroup O156 from healthy and diseased ruminants such as cattle, sheep, and goats was reported by several authors (1, 11-13, 21, 39, 46, 50, 52). Additional EHEC-associated virulence genes such as stx, eae, hlyAEHEC, or nlaA were found preferentially in the serotypes O156:H25 and O156:H− (11-13, 21, 22, 50, 52).Numerous methods exist for the detection of pathogenic E. coli, including genotypic and phenotypic marker assays for the detection of virulence genes and their products (19, 47, 55, 57). All of these methods have the common drawback of screening a relatively small number of determinants simultaneously. A diagnostic DNA microarray based on the ArrayTube format of CLONDIAG GmbH was developed as a viable alternative due to its ability to screen multiple virulence markers simultaneously (2). Further microarray layouts working with the same principle but different gene targets were developed for the rapid identification of antimicrobial resistance genes in Gram-negative bacteria (5) and for the rapid DNA-based serotyping of E. coli (4). In addition, a protein microarray for E. coli O serotyping based on the ArrayTube format was described by Anjum et al. (3).The aim of our study was the molecular genotyping of bovine E. coli field isolates of serogroup O156 based on miniaturized E. coli oligonucleotide arrays in the ArrayStrip format and to combine the screening of E. coli virulence markers, antimicrobial resistance genes, and DNA serotyping targets, some of which were partially described previously for separate arrays (2, 4, 5). The epidemiological situation in the beef herds from which the isolates were obtained and the spatial and temporal behavior of the clonal distribution of E. coli serogroup O156 were analyzed during the observation period. The potential risk of the isolates inducing disease in humans was assessed.  相似文献   

19.
20.
Human cytomegalovirus (HCMV) UL37 proteins traffic sequentially from the endoplasmic reticulum (ER) to the mitochondria. In transiently transfected cells, UL37 proteins traffic into the mitochondrion-associated membranes (MAM), the site of contact between the ER and mitochondria. In HCMV-infected cells, the predominant UL37 exon 1 protein, pUL37x1, trafficked into the ER, the MAM, and the mitochondria. Surprisingly, a component of the MAM calcium signaling junction complex, cytosolic Grp75, was increasingly enriched in heavy MAM from HCMV-infected cells. These studies show the first documented case of a herpesvirus protein, HCMV pUL37x1, trafficking into the MAM during permissive infection and HCMV-induced alteration of the MAM protein composition.The human cytomegalovirus (HCMV) UL37 immediate early (IE) locus expresses multiple products, including the predominant UL37 exon 1 protein, pUL37x1, also known as viral mitochondrion-localized inhibitor of apoptosis (vMIA), during lytic infection (16, 22, 24, 39, 44). The UL37 glycoprotein (gpUL37) shares UL37x1 sequences and is internally cleaved, generating pUL37NH2 and gpUL37COOH (2, 22, 25, 26). pUL37x1 is essential for the growth of HCMV in humans (17) and for the growth of primary HCMV strains (20) and strain AD169 (14, 35, 39, 49) but not strain TownevarATCC in permissive human fibroblasts (HFFs) (27).pUL37x1 induces calcium (Ca2+) efflux from the endoplasmic reticulum (ER) (39), regulates viral early gene expression (5, 10), disrupts F-actin (34, 39), recruits and inactivates Bax at the mitochondrial outer membrane (MOM) (4, 31-33), and inhibits mitochondrial serine protease at late times of infection (28).Intriguingly, HCMV UL37 proteins localize dually in the ER and in the mitochondria (2, 9, 16, 17, 24-26). In contrast to other characterized, similarly localized proteins (3, 6, 11, 23, 30, 38), dual-trafficking UL37 proteins are noncompetitive and sequential, as an uncleaved gpUL37 mutant protein is ER translocated, N-glycosylated, and then imported into the mitochondria (24, 26).Ninety-nine percent of ∼1,000 mitochondrial proteins are synthesized in the cytosol and directly imported into the mitochondria (13). However, the mitochondrial import of ER-synthesized proteins is poorly understood. One potential pathway is the use of the mitochondrion-associated membrane (MAM) as a transfer waypoint. The MAM is a specialized ER subdomain enriched in lipid-synthetic enzymes, lipid-associated proteins, such as sigma-1 receptor, and chaperones (18, 45). The MAM, the site of contact between the ER and the mitochondria, permits the translocation of membrane-bound lipids, including ceramide, between the two organelles (40). The MAM also provides enriched Ca2+ microdomains for mitochondrial signaling (15, 36, 37, 43, 48). One macromolecular MAM complex involved in efficient ER-to-mitochondrion Ca2+ transfer is comprised of ER-bound inositol 1,4,5-triphosphate receptor 3 (IP3R3), cytosolic Grp75, and a MOM-localized voltage-dependent anion channel (VDAC) (42). Another MAM-stabilizing protein complex utilizes mitofusin 2 (Mfn2) to tether ER and mitochondrial organelles together (12).HCMV UL37 proteins traffic into the MAM of transiently transfected HFFs and HeLa cells, directed by their NH2-terminal leaders (8, 47). To determine whether the MAM is targeted by UL37 proteins during infection, we fractionated HCMV-infected cells and examined pUL37x1 trafficking in microsomes, mitochondria, and the MAM throughout all temporal phases of infection. Because MAM domains physically bridge two organelles, multiple markers were employed to verify the purity and identity of the fractions (7, 8, 19, 46, 47).(These studies were performed in part by Chad Williamson in partial fulfillment of his doctoral studies in the Biochemistry and Molecular Genetics Program at George Washington Institute of Biomedical Sciences.)HFFs and life-extended (LE)-HFFs were grown and not infected or infected with HCMV (strain AD169) at a multiplicity of 3 PFU/cell as previously described (8, 26, 47). Heavy (6,300 × g) and light (100,000 × g) MAM fractions, mitochondria, and microsomes were isolated at various times of infection and quantified as described previously (7, 8, 47). Ten- or 20-μg amounts of total lysate or of subcellular fractions were resolved by SDS-PAGE in 4 to 12% Bis-Tris NuPage gels (Invitrogen) and examined by Western analyses (7, 8, 26). Twenty-microgram amounts of the fractions were not treated or treated with proteinase K (3 μg) for 20 min on ice, resolved by SDS-PAGE, and probed by Western analysis. The blots were probed with rabbit anti-UL37x1 antiserum (DC35), goat anti-dolichyl phosphate mannose synthase 1 (DPM1), goat anti-COX2 (both from Santa Cruz Biotechnology), mouse anti-Grp75 (StressGen Biotechnologies), and the corresponding horseradish peroxidase-conjugated secondary antibodies (8, 47). Reactive proteins were detected by enhanced chemiluminescence (ECL) reagents (Pierce), and images were digitized as described previously (26, 47).  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