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.
In a laboratory incubation study, methane (CH4) production in two saline soils and a nonsaline soil sample was investigated under flooded conditions. Mean CH4 production was remarkable (630.86 ng CH4/g) in nonsaline alluvial soil, but low (12.97 ng CH4/g) in acid sulfate saline (Pokkali) soil which was attributed to the high sulfate content of the later. CH4 production was also low in the coastal saline (Canning) soil (142.36 ng CH4/g) but increased upon leaching the soil of its salt content. Addition of salts to the nonsaline alluvial soil at 4, 8, 16 and 20 dS/m progressively decreased CH4 production. The inhibition of CH4 production was related to low microbial activities as reflected by decreased microbial biomass C and low soil microbial population including that of methanogens.  相似文献   

3.
Abstract Contamination of ground water with halogenated aliphatic hydrocarbons threatens this source of drinking water. In order to study microbial processes that may enhance the removal of these compounds, Lincoln fine sand was exposed to an atmosphere containing methane (4%) to enrich microorganisms capable of growth on this gaseous hydrocarbon. The methane-enriched soil was then tested to determine whether the enriched microbes could remove seven halocarbons from aqueous solution. Removal of dichloromethane. trans -1,2-dichloroethylene, chloroform, 1,2-dichloroethane, trichloroethylene, and 1,1,1-trichloroethane was significantly different in methane-enriched soil compared to non-enriched soil (ANOVA, 95% significance level). Tetrachloroethylene was not removed. Autoclaving the methane-enriched soil inhibited completely the removal of all the compounds. Once the soil was enriched with methane, its presence in the headspace was not required for removal of several of the compounds but methane was required for their complete removal. These results suggest that methane stimulation of microbial communities may be an alternative treatment technology for bioremediation of contaminated subsurface soils and ground water.  相似文献   

4.
The metabolism of atmospheric methane in a forest soil was studied by radiotracer techniques. Maximum (sup14)CH(inf4) oxidation (163.5 pmol of C cm(sup-3) h(sup-1)) and (sup14)C assimilation (50.3 pmol of C cm(sup-3) h(sup-1)) occurred at the A(inf2) horizon located 15 to 18 cm below the soil surface. At this depth, 31 to 43% of the atmospheric methane oxidized was assimilated into microbial biomass; the remaining methane was recovered as (sup14)CO(inf2). Methane-derived carbon was incorporated into all major cell macromolecules by the soil microorganisms (50% as proteins, 19% as nucleic acids and polysaccharides, and 5% as lipids). The percentage of methane assimilated (carbon conversion efficiency) remained constant at temperatures between 5 and 20(deg)C, followed by a decrease at 30(deg)C. The carbon conversion efficiency did not increase at methane concentrations between 1.7 and 1,000 ppm. In contrast, the overall methane oxidation activity increased at elevated methane concentrations, with an apparent K(infm) of 21 ppm (31 nM CH(inf4)) and a V(infmax) of 188 pmol of CH(inf4) cm(sup-3) h(sup-1). Methane oxidizers from soil depths with maximum methanotrophic activity respired approximately 1 to 3% of the assimilated methane-derived carbon per day. This apparent endogenous respiration did not change significantly in the absence of methane. Similarly, the potential for oxidation of atmospheric methane was relatively insensitive to methane starvation. Soil samples from depths above and below the zone with maximum atmospheric methane oxidation activity showed a dramatic increase in the turnover of the methane assimilated (>20 times increase). Physical disturbance such as sieving or mixing of soil samples decreased methane oxidation and assimilation by 50 to 58% but did not alter the carbon conversion efficiency. Ammonia addition (0.1 or 1.0 (mu)mol g [fresh weight](sup-1)) decreased both methane oxidation and carbon conversion efficiency. This resulted in a dramatic decrease in methane assimilation (85 to 99%). In addition, ammonia-treated soil showed up to 10 times greater turnover of the assimilated methane-derived carbon (relative to untreated soil). The results suggest a potential for microbial growth on atmospheric methane. However, growth was regulated strongly by soil parameters other than the methane concentration. The pattern observed for metabolism of atmospheric methane in soils was not consistent with the physiology of known methanotrophic bacteria.  相似文献   

5.
In laboratory incubation experiments, application of a commercial formulation of the herbicide butachlor (N-butoxymethyl-2-chloro-2',6'-diethyl acetanilide) to three tropical rice soils, widely differing in their physicochemical characteristics, under flooded condition inhibited methane (CH4) production. The inhibitory effect was concentration dependent and most remarkable in the alluvial soil. Thus, following application of butachlor at 5, 10, 50 and 100 microg g(-1) soil, respectively, cumulative CH4 production in the alluvial soil was inhibited by 15%, 31%, 91% and 98% over unamended control. Since CH4 production was less pronounced in the sandy loam and acid sulfate soil, the impact of amendment with butchalor, albeit inhibitory, was less extensive than the alluvial soil. Inhibition of CH4 production in butachlor-amended alluvial soil was related to the prevention in the drop in redox potential as well as low methanogenic bacterial population especially at high concentrations of butachlor. CH4 oxidation was also inhibited in butachlor-amended alluvial soil with the inhibitory effect being more prevalent under flooded condition. Inhibition in CH4 oxidation was related to a reduction in the population of soluble methane monooxygenase producing methanotrophs. Results demonstrate that butachlor, a commonly used herbicide in rice cultivation, even at very low concentrations can affect CH4 production and its oxidation, thereby influencing the biogeochemical cycle of CH4 in flooded rice soils.  相似文献   

6.
Although CH 4 production is sensitive to temperature, it is not clear how temperature controls CH 4 production directly versus the production of organic substrates that methanogens convert into CH 4 . Therefore, this study was done to better understand how CH 4 production in rice paddy soil responded to temperature when the process was not limited by the availability of substrates. In a laboratory-incubation study using three Indian rice soils under flooded conditions, the effect of temperature on CH 4 production was examined. CH 4 production in acid sulphate, laterite, and alluvial soil samples under flooded conditions distinctly increased with increase in temperature from 15°C to 35°C. Laterite and acid sulphate soils produced distinctly less CH 4 than alluvial soils. CO 2 production increased with increase in temperature in all the soils. The readily mineralizable carbon C and Fe 2+ contents in soils were least at 15°C and highest at 35°C, irrespective of soil type. Likewise, a significant correlation existed between microbial population (methanogens and sulphate reducers) and CH 4 production. Comparing the temperature coefficients ( Q 10 ) for methane production within each soil type at low (15°C-25°C) and medium (25°C-35°C) temperature intervals revealed that these values were not uniform for both alluvial and laterite soils. But acid sulphate soil had Q 10 values that were near 2 at both temperature intervals. When these soil samples were amended with substrates (acetate, H 2 -CO 2 , and rice straw), there were stimulatory effects on methane production rates and consequently on the Q 10 values. The pattern of temperature coefficients was characteristic of the soil type and the nature of substrates used for amendment.  相似文献   

7.
Temperature change affects methane consumption in soil. However, there is no information on possible temperature control of methanotrophic bacterial populations. Therefore, we studied CH(4) consumption and populations of methanotrophs in an upland forest soil and a rice field soil incubated at different temperatures between 5 and 45 degrees C for up to 40 days. Potential methane consumption was measured at 4% CH(4). The temporal progress of CH(4) consumption indicated growth of methanotrophs. Both soils showed maximum CH(4) consumption at 25-35 degrees C, but no activity at >40 degrees C. In forest soil CH(4) was also consumed at 5 degrees C, but in rice soil only at 15 degrees C. Methanotroph populations were assessed by terminal restriction fragment length polymorphism (T-RFLP) targeting particulate methane monooxygenase (pmoA) genes. Eight T-RFs with relative abundance >1% were retrieved from both forest and rice soil. The individual T-RFs were tentatively assigned to different methanotrophic populations (e.g. Methylococcus/Methylocaldum, Methylomicrobium, Methylobacter, Methylocystis/Methylosinus) according to published sequence data. Two T-RFs were assigned to ammonium monooxygenase (amoA) gene sequences. Statistical tests showed that temperature affected the relative abundance of most T-RFs. Furthermore, the relative abundance of individual T-RFs differed between the two soils, and also exhibited different temperature dependence. We conclude that temperature can be an important factor regulating the community composition of methanotrophs in soil.  相似文献   

8.
Experiments were done to test the hypothesis that atmospheric CH(4) oxidizers in a well-drained alpine tundra soil are supported by CH(4) production from anaerobic microsites in the soil. Soil was subjected to 22 days of anaerobic conditions with elevated H(2) and CO(2) in order to stimulate methanogenesis. This treatment stimulated subsequent atmospheric CH(4) consumption, probably by increasing soil methanogenesis. After removal from anaerobic conditions, soils emitted CH(4) for up to 6 h, then oxidized atmospheric CH(4) at 111 (+/- 5.7) pmol (g dry weight)(-1) h(-1), which was more than 3 times the rate of control soils. Further supporting our hypothesis, additions of lumazine, a highly specific inhibitor of methanogenesis, prevented the stimulation of atmospheric CH(4) oxidation by the anaerobic treatment. The method used to create anaerobic conditions with elevated H(2) and CO(2) also elevated headspace CH(4) concentrations. However, elevated CH(4) concentrations under aerobic conditions did not stimulate CH(4) oxidation as much as preexposure to H(2) and CO(2) under anaerobic conditions. Anaerobic conditions created by N(2) flushing did not stimulate atmospheric CH4 oxidation, probably because N2 flushing inhibited methanogenesis by removing necessary precursors for methane production. We conclude that anaerobic conditions with elevated H(2) and CO(2) stimulate atmospheric CH(4) oxidation in this dry alpine tundra soil by increasing endogenous CH(4) production. This effect was prevented by inhibiting methanogenesis, indicating the importance of endogenous CH(4) production in a CH(4-) consuming soil.  相似文献   

9.
采用野外原位实验静态箱-气相色谱法,研究了兴安岭多年冻土不同程度退化地区生长季湿地土壤温室气体CH4、CO2和N2O的排放通量特征,同时分析了环境因子对土壤温室气体排放的影响。结果表明:1) 3种类型冻土区(季节性冻土区、岛状多年冻土区、连续多年冻土区,分别用D1、D2、D3表示)土壤在生长季时期表现为CO2和N2O的源; D1和D3为CH4的源,D2为CH4的汇。D1、D2、D3土壤在生长季中平均CH4排放通量分别为(0.127±0.021)、(-0.020±0. 006)、(0. 082±0. 019) mg·m-2·h-1; CO2排放通量分别为(371.50±66.73)、(318.43±55.67)、(213.19±37.05) mg·m-2·h-1; N2O排放通量分别为...  相似文献   

10.
Microorganisms that oxidize atmospheric methane in soils were characterized by radioactive labelling with (14)CH(4) followed by analysis of radiolabelled phospholipid ester-linked fatty acids ((14)C-PLFAs). The radioactive fingerprinting technique was used to compare active methanotrophs in soil samples from Greenland, Denmark, the United States, and Brazil. The (14)C-PLFA fingerprints indicated that closely related methanotrophic bacteria were responsible for the oxidation of atmospheric methane in the soils. Significant amounts of labelled PLFAs produced by the unknown soil methanotrophs coeluted with a group of fatty acids that included i17:0, a17:0, and 17:1omega8c (up to 9.0% of the total (14)C-PLFAs). These PLFAs are not known to be significant constituents of methanotrophic bacteria. The major PLFAs of the soil methanotrophs (73.5 to 89.0% of the total PLFAs) coeluted with 18:1 and 18:0 fatty acids (e.g., 18:1omega9, 18:1omega7, and 18:0). The (14)C-PLFAs fingerprints of the soil methanotrophs that oxidized atmospheric methane did not change after long-term methane enrichment at 170 ppm CH(4). The (14)C-PLFA fingerprints of the soil methanotrophs were different from the PLFA profiles of type I and type II methanotrophic bacteria described previously. Some similarity at the PLFA level was observed between the unknown soil methanotrophs and the PLFA phenotype of the type II methanotrophs. Methanotrophs in Arctic, temperate, and tropical regions assimilated between 20 and 54% of the atmospheric methane that was metabolized. The lowest relative assimilation (percent) was observed for methanotrophs in agricultural soil, whereas the highest assimilation was observed for methanotrophs in rain forest soil. The results suggest that methanotrophs with relatively high carbon conversion efficiencies and very similar PLFA compositions dominate atmospheric methane metabolism in different soils. The characteristics of the methane metabolism and the (14)C-PLFA fingerprints excluded any significant role of autotrophic ammonia oxidizers in the metabolism of atmospheric methane.  相似文献   

11.
Aerobic grasslands may consume significant amounts of atmospheric methane (CH4). We aimed (i) to assess the spatial and temporal variability of net CH4 fluxes from grasslands on aerobic sandy soils, and (ii) to explain the variability in net CH4 fluxes by differences in soil moisture content and temperature. Net CH4 fluxes were measured with vented closed flux chambers at two sites with low N input on sandy soils in the Netherlands: (i) Wolfheze, a heather grassland, and (ii) Bovenbuurtse Weilanden, a grassland which is mown twice a year. Spatial variability of net CH4 fluxes was analysed using geostatistics. In incubation experiments, the effects of soil moisture content and temperature on CH4 uptake capacity were assessed. Temporal variability of net CH4 fluxes at Wolfheze was related to differences in soil temperature (r2 of 0.57) and soil moisture content (r2 of 0.73). Atmospheric CH4 uptake was highest at high soil temperatures and intermediate soil moisture contents. Spatial variability of net CH4 fluxes was high, both at Wolfheze and at Bovenbuurtse Weilanden. Incubation experiments showed that, at soil moisture contents lower than 5% (w/w), CH4 uptake was completely inhibited, probably due to physiological water stress of methanotrophs. At soil moisture contents higher than 50% (w/w), CH4 uptake was greatly reduced, probably due to the slow down of diffusive CH4 and O2 transport in the soil, which may have resulted in reduced CH4 oxidation and possibly some CH4 production. Optimum soil moisture contents for CH4 uptake were in the range of 20 – 35% (w/w), as prevailing in the field. The sensitivity of CH4 uptake to soil moisture content may result in short-term variability of net atmospheric CH4 uptake in response to precipitation and evapotranspiration, as well as in long-term variability due to changing precipitation patterns as a result of climate change.  相似文献   

12.
The global methane cycle includes both terrestrial and atmospheric processes and may contribute to feedback regulation of the climate. Most oxic soils are a net sink for methane, and these soils consume approximately 20 to 60 Tg of methane per year. The soil sink for atmospheric methane is microbially mediated and sensitive to disturbance. A decrease in the capacity of this sink may have contributed to the approximately 1%. year(-1) increase in the atmospheric methane level in this century. The organisms responsible for methane uptake by soils (the atmospheric methane sink) are not known, and factors that influence the activity of these organisms are poorly understood. In this study the soil methane-oxidizing population was characterized by both labelling soil microbiota with (14)CH(4) and analyzing a total soil monooxygenase gene library. Comparative analyses of [(14)C]phospholipid ester-linked fatty acid profiles performed with representative methane-oxidizing bacteria revealed that the soil sink for atmospheric methane consists of an unknown group of methanotrophic bacteria that exhibit some similarity to type II methanotrophs. An analysis of monooxygenase gene libraries from the same soil samples indicated that an unknown group of bacteria belonging to the alpha subclass of the class Proteobacteria was present; these organisms were only distantly related to extant methane-oxidizing strains. Studies on factors that affect the activity, population dynamics, and contribution to global methane flux of "atmospheric methane oxidizers" should be greatly facilitated by use of biomarkers identified in this study.  相似文献   

13.
Production and consumption processes in soils contribute to the global cycles of many trace gases (CH4, CO, OCS, H2, N2O, and NO) that are relevant for atmospheric chemistry and climate. Soil microbial processes contribute substantially to the budgets of atmospheric trace gases. The flux of trace gases between soil and atmosphere is usually the result of simultaneously operating production and consumption processes in soil: The relevant processes are not yet proven with absolute certainty, but the following are likely for trace gas consumption: H2 oxidation by abiontic soil enzymes; CO cooxidation by the ammonium monooxygenase of nitrifying bacteria; CH4 oxidation by unknown methanotrophic bacteria that utilize CH4 for growth; OCS hydrolysis by bacteria containing carbonic anhydrase; N2O reduction to N2 by denitrifying bacteria; NO consumption by either reduction to N2O in denitrifiers or oxidation to nitrate in heterotrophic bacteria. Wetland soils, in contrast to upland soils are generally anoxic and thus support the production of trace gases (H2, CO, CH4, N2O, and NO) by anaerobic bacteria such as fermenters, methanogens, acetogens, sulfate reducers, and denitrifiers. Methane is the dominant gaseous product of anaerobic degradation of organic matter and is released into the atmosphere, whereas the other trace gases are only intermediates, which are mostly cycled within the anoxic habitat. A significant percentage of the produced methane is oxidized by methanotrophic bacteria at anoxic-oxic interfaces such as the soil surface and the root surface of aquatic plants that serve as conduits for O2 transport into and CH4 transport out of the wetland soils. The dominant production processes in upland soils are different from those in wetland soils and include H2 production by biological N2 fixation, CO production by chemical decomposition of soil organic matter, and NO and N2O production by nitrification and denitrification. The processes responsible for CH4 production in upland soils are completely unclear, as are the OCS production processes in general. A problem for future research is the attribution of trace gas metabolic processes not only to functional groups of microorganisms but also to particular taxa. Thus, it is completely unclear how important microbial diversity is for the control of trace gas flux at the ecosystem level. However, different microbial communities may be part of the reason for differences in trace gas metabolism, e.g., effects of nitrogen fertilizers on CH4 uptake by soil; decrease of CH4 production with decreasing temperature; or different rates and modes of NO and N2O production in different soils and under different conditions.  相似文献   

14.
Laboratory-scale soil microcosms containing different soils were permeated with CH(inf4) for up to 6 months to investigate their capacity to develop a methanotrophic community. Methane emissions were monitored continuously until steady states were established. The porous, coarse sand soil developed the greatest methanotrophic capacity (10.4 mol of CH(inf4) (middot) m(sup-2) (middot) day(sup-1)), the greatest yet reported in the literature. Vertical profiles of O(inf2), CH(inf4), and methanotrophic potential in the soils were determined at steady state. Methane oxidation potentials were greatest where the vertical profiles of O(inf2) and CH(inf4) overlapped. A significant increase in the organic matter content of the soil, presumably derived from methanotroph biomass, occurred where CH(inf4) oxidation was greatest. Methane oxidation kinetics showed that a soil community with a low methanotrophic capacity (V(infmax) of 258 nmol (middot) g of soil(sup-1) (middot) h(sup-1)) but relatively high affinity (k(infapp) of 1.6 (mu)M) remained in N(inf2)-purged control microcosms, even after 6 months without CH(inf4). We attribute this to a facultative, possibly mixotrophic, methanotrophic microbial community. When purged with CH(inf4), a different methanotrophic community developed which had a lower affinity (k(infapp) of 31.7 (mu)M) for CH(inf4) but a greater capacity (V(infmax) of 998 nmol (middot) g of soil(sup-1) (middot) h(sup-1)) for CH(inf4) oxidation, reflecting the enrichment of an active high-capacity methanotrophic community. Compared with the unamended control soil, amendment of the coarse sand with sewage sludge enhanced CH(inf4) oxidation capacity by 26%; K(inf2)HPO(inf4) amendment had no significant effect, while amendment with NH(inf4)NO(inf3) reduced the CH(inf4) oxidation capacity by 64%. In vitro experiments suggested that NH(inf4)NO(inf3) additions (10 and 71 (mu)mol (middot) g of soil(sup-1)) inhibited CH(inf4) oxidation by a nonspecific ionic effect rather than by specific inhibition by NH(inf4)(sup+).  相似文献   

15.
Methanotrophic bacteria play a crucial role in regulating the emission of CH4 from rice fields into the atmosphere. We investigated the CH4 oxidation activity together with the diversity of methanotrophic bacteria in ten rice field soils from different geographic locations. Upon incubation of aerated soil slurries under 7% CH4, rates of CH4 oxidation increased after a lag phase of 1-4 days and reached values of 3-10 micromol d(-1) g-dw(-1) soil. The methanotrophic community was assayed by retrieval of the pmoA gene which encodes the a subunit of the particulate methane monooxygenase. After extraction of DNA from actively CH4-oxidizing soil samples and PCR-amplification of the pmoA, the community was analyzed by Denaturant Gradient Gel Electrophoresis (DGGE) and Terminal Restriction Fragment Length Polymorphism (T-RFLP). DGGE bands were excised, the pmoA re-amplified, sequenced and the encoded amino acid sequence comparatively analyzed by phylogenetic treeing. The analyses allowed the detection of pmoA sequences related to the following methanotrophic genera: the type-I methanotrophs Methylobacter, Methylomicrobium, Methylococcus and Methylocaldum, and the type-II methanotrophs Methylocystis and Methylosinus. T-RFLP analysis detected a similar diversity, but type-II pmoA more frequently than DGGE. All soils but one contained type-II in addition to type-I methanotrophs. Type-I Methylomonas was not detected at all. Different combinations of methanotrophic genera were detected in the different soils. However, there was no obvious geographic pattern of the distribution of methanotrophs.  相似文献   

16.
Corrinoids, such as aquocobalamin, methylcobalamin, and (cyanoaquo)cobinamide, catalyze the reductive dehalogenation of CCl4 with titanium(III) citrate as the electron donor [Krone et al. (1989) Biochemistry 28, 4908-4914]. We report here that this reaction is also effectively mediated by the nickel-containing porphinoid, coenzyme F430, found in methanogenic bacteria. Chloroform, methylene chloride, methyl chloride, and methane were detected as intermediates and products. Ethane was formed in trace amounts, and several as yet unidentified nonvolatile compounds were also generated. The rate of dehalogenation decreased in the series of CCl4, CHCl3, and CH2Cl2. With coenzyme F430 as the catalyst, the reduction of CH3Cl to CH4 proceeded more than 50 times faster than with aquocobalamin. Cell suspensions of Methanosarcina barkeri were found to catalyze the reductive dehalogenation of CCl4 with CO as the electron donor (E'0 = -0.524 V). Methylene chloride was the main end product. The kinetics of CHCl3 and CH2Cl2 formation from CCl4 were similar to those with coenzyme F430 or aquocobalamin as catalysts and titanium(III) citrate as the reductant.  相似文献   

17.
The potential activity of methane production was determined in the vertical profiles of the peat deposits of three bogs in Tver oblast, which were representative of the boreal zone. In the minerotrophic fen, the rates of methane production measured throughout the profile did not change significantly with depth and comprised 3-6 ng CH4-C g(-1) h(-1). In ombrotrophic peat bogs, the rate did not exceed 5 ng CH4-C g(-1) h(-1) in the upper layer of the profile (up to 1.5 m) and increased to 15-30 ng CH4-C g(-1) h(-1) in the deep layers of the peat deposits. The distribution of fermentative microorganisms and methanogens in the profiles of peat deposits was uniform in all the studied bogs. In bog water samples, the presence of butyrate (up to 14.1 mg l(-1)) and acetate (up to 2.4 mg l(-1)) was revealed throughout the whole profile; in the upper 0.5-m layer of the ombrotrophic bogs, formate (up to 8.9 mg l(-1)) and propionate (up to 0.3 mg l(-1)) were detected as well. The arrangement of local maxima of the fatty acid content and methanogenic activity in the peat deposits, as well as the decrease in the acetate concentrations during summer, support the hypothesis that the initial substrates for methanogenesis come from the upper peat layers. It was established that the addition of sulfate and nitrate inhibits methane production in peat samples: the changes in the concentrations, recorded in situ, may also influence the methane content in peat layers.  相似文献   

18.
Anoxic soils, such as flooded rice fields, are major sources of the greenhouse gas CH(4) while oxic upland soils are major sinks of atmospheric CH(4). Nevertheless, CH(4) is also consumed in rice fields where up to 90% of the produced CH(4) is oxidized in a narrow oxic zone around the rice roots and in the soil surface layer before it escapes into the atmosphere. After 1 day drainage of rice field soil, CH(4) oxidation was detected in the top 2-mm soil layers, but after 8 days drainage the zone of CH(4) oxidation extended to 8 mm depth. Simultaneously, the potential for CH(4) production decreased, but some production was still detectable after 8 days drainage throughout the soil profile. The vertical distribution of the methanotrophic community was also monitored after 1 and 8 days drainage using denaturing gradient gel electrophoresis after PCR amplification with primer sets targeting two regions on the 16S rRNA gene that are relatively specific for methylotrophic alpha- and gamma-Proteobacteria, and targeting two functional genes encoding subunits of key enzymes in all methanotrophs, i.e. the genes for the particulate methane monooxygenase (pmoA) and the methanol dehydrogenase (mxaF). Drainage stimulated the methanotrophic community. Eight days after drainage, new methanotrophic populations appeared and a distinct methanotrophic community developed. The population structure of type I and II methanotrophs was differently affected by drainage. Type II methanotrophs (alpha-Proteobacteria) were present throughout the soil core directly after drainage (1 day), and the community composition remained largely unchanged with depth. Only two new type II populations appeared after 8 days of drainage. Drainage had a more pronounced impact on the type I methanotrophic community (gamma-Proteobacteria). Type I populations were not or only weakly detected 1 day after drainage. However, after 8 days of drainage, a large diversity of type I methanotrophs were detected, altough they were not evenly distributed throughout the soil core but dominated at different depths. A distinct type I community structure had developed within each soil section between 0 and 20 mm soil depth, indicating the widening of suitable habitats for methanotrophs in the rice field soil within 1 week of drainage.  相似文献   

19.
Ring-U-[14C]methoxychlor [1,1-bis(p-methoxyphenyl)-2,2,2-trichloroethane] was incubated in soil under aerobic and anaerobic conditions. Primary degradation of methoxychlor occurred under anaerobic conditions, but not under aerobic conditions, after 3 months of incubation. Analysis of soil extracts, using gas chromatography, demonstrated that only 10% of the compound remained at initial concentrations of 10 and 100 ppm (wt/wt) of methoxychlor. Evidence is presented that a dechlorination reaction was responsible for primary degradation of methoxychlor. Analysis of soils treated with 100 ppm of methoxychlor in the presence of 2% HgCl2 showed that 100% of the compound remained after 3 months, indicating that degradation in the unpoisoned flasks was biologically mediated. Methanogenic organisms, however, are probably not involved, as strong inhibition of methane production was observed in all soils treated with methoxychlor. During the 3-month incubation period, little or no evaluation of 14CO2 or 14CH4 occurred under either aerobic or anaerobic conditions. Cometabolic processes may be responsible for the extensive molecular changes which occurred with methoxychlor because the rate of its disappearance from soil was observed to level off after exhaustion of soil organic matter. After this incubation period, soils previously incubated under anaerobic conditions were converted to aerobic conditions. The rates of 14CO2 evolution from soils exposed to anaerobic and aerobic sequences of environments ranged from 10- to 70-fold greater than that observed for soils exposed solely to an aerobic environment.  相似文献   

20.
Emissions of N2O from cover soils of both abandoned (> 30 years) and active landfills greatly exceed the maximum fluxes previously reported for tropical soils, suggesting high microbial activities for N2O production. Low soil matrix potentials (<-0.7 MPa) indicate that nitrification was the most likely mechanism of N2O formation during most of the time of sampling. Soil moisture had a strong influence on N2O emissions. The production of N2O was stimulated by as much as 20 times during laboratory incubations, when moisture was increased from -2.0 MPa to -0.6 MPa. Additional evidence from incubation experiments and delta13C analyses of fatty acids (18:1) diagnostic of methanotrophs suggests that N2O is formed in these soils by nitrification via methanotrophic bacteria. In a NH3(g)-amended landfill soil, the rate of N2O production was significantly increased when incubated with 100 ppmv methane compared with 1.8 ppmv (atmospheric) methane. Preincubation of a landfill soil with 1% CH4 for 2 weeks resulted in higher rates of N2O production when subsequently amended with NH3(g) relative to a control soil preincubated without CH4. At one location, at the soil depth (9-16 cm) of maximum methane consumption and N2O production, we observe elevated concentrations of organic carbon and nitrogen and distinct minima in delta15N (+1.0%) and delta13C (-33.8%) values for organic nitrogen and organic carbon respectively. A delta13C value of -39.3% was measured for 18:1 carbon fatty acids in this soil, diagnostic of type II methanotrophs. The low delta15N value for organic nitrogen is consistent with N2 fixation by type II methanotrophs. These observations all point to a methanotrophic origin for the organic matter at this depth. The results of this study corroborate previous reports of methanotrophic nitrification and N2O formation in aqueous and soil environments and suggest a predominance of type II rather than type I or type X methanotrophs in this landfill soil.  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