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 31 毫秒
1.
Soils are the largest sink of atmospheric hydrogen contributing about 75% to the total budget. Atmospheric H2 is assumed to be oxidized in soil by abiontic soil hydrogenases. Extraction of a forest soil with a slightly alkaline (pH 8.5) buffer containing polyethylene glycol (PEG), followed by filtration yielded a bacteria-free extract that oxidized H2 at ambient concentrations (0.2–2.0 ppmv). Hydrogenase activity was assayed by gas chromatographic analysis of H2 consumption and by conversion of 3H2 to tritiated water. Only less than 2% of the original activity was recovered in the extract. Kinetic analysis nevertheless resulted in a biphasic kinetics exhibiting two Km and Vmax values that were similar to those detected in the original soil. In addition, activities of both original soil and soil extract showed similar optima at pH 4–6 and at 30 °C, indicating that representative fractions of soil hydrogenases were recovered in the extract. Precipitation with PEG or ultrafiltration allowed further purification of the activity, albeit only about 20% of that in the crude extract could be recovered in the precipitate or the fraction >100 kDa.  相似文献   

2.
Low atmospheric H2 concentrations (0.55 ppmv) are oxidized in soils by a high-affinity activity with typical characteristics of an abiontic soil enzyme. This activity was measured in a meadow cambisol and a forest cambisol. In both soils, the maximum activity was reached at a soil moisture of about 20% water-holding capacity, and was localized in the top Ah horizon. The soils were fractionated by dry sieving and wet filtration into nine different particle-size fractions, ranging from 3 to 2000 m in size. H2 oxidation was measured by three different assays and was compared to the ATP content and microscopic counts of bacteria in the same fractions. In the meadow soil, the specific activities of H2 oxidation increased with the particle size (maximum at 200–500 m), whereas ATP and bacterial counts showed no trend. In the forest soil, the specific activities of H2 oxidation increased with the particle size up to 50–100 m, and then decreased again. ATP and bacterial counts, however, showed the opposite trend, i.e., decreased with an increasing particle size. Thus the H2-oxidizing activity was not correlated with characteristic microbial biomass parameters. Although significant percentage (29–64%) of randomly isolated heterotrophic bacteria was able to oxidize H2, this activity was too small to account for the H2 oxidation in the soil. In both soils, most of the activity present was found in particles of 100–500 m in size. The recovery shifted to smaller size fractions when larger soil aggregates were broken up by wet instead of dry sieving. Attempts to extract the H2-oxidizing activity from the soil particles were unsuccessful.  相似文献   

3.
Summary NO and N2O release rates were measured in an acidic forest soil (pH 4.0) and a slightly alkaline agricultural soil (pH 7.8) after the pH was adjusted to values ranging from pH 4.0 to 7.8. The total release of NO and N2O during 20 h of incubation was determined together with the net changes in the concentrations of NH 4 + , NO 2 and NO 3 in the soil. The release of NO and N2O increased after fertilization with NH 4 + and/or NO 3 ; it strongly decreased with increasing pH in the acidic forest soil; and it increased when the pH of the alkaline agricultural soil was decreased to pH 6.5. However, there was no simple correlation between NO and N2O release or between these compounds and activities such as the NO 2 accumulation, NO 3 reduction, or NH 4 + oxidation. We suggest that soil pH exerts complex controls, e.g., on microbial populations or enzyme activities involved in nitrification and denitrification.  相似文献   

4.
Summary The temperature dependence of the NO production rate and the NO consumption rate constant was measured in an Egyptian soil, a soil from the Bavarian Forest, and a soil from the Donau valley, together with the temperature dependence of the potential rates of ammonium oxidation, nitrite oxidation, and denitrification, and the temperature dependence of the growth of NH inf4 sup+ -oxidizing, NO inf2 sup- -oxidizing, and NO inf3 sup- -reducing bacteria in most probable number assays. In the acidic Bavarian Forest soil, NO production was only stimulated by the addition of NO inf3 sup- but not NH inf4 sup+ . However, NO production showed no temperature optimum, indicating that it was due to chemical processes. Most probable numbers and potential activities of nitrifiers were very low. NO consumption, in contrast, showed a temperature optimum at 25°C, demonstrating that consumption and production of NO were regulated individually by the soil temperature. In the neutral, subtropical Egyptian soil, NO production was stimulated only by the addition of NH inf4 sup+ but not NO inf3 sup- . All activities and most probable numbers showed a temperature optimum at 25° or 30°C and exhibited apparent activation energies between 61 and 202 kJ mol-1. However, a few nitrifiers and denitrifiers were also able to grow at 8° or 50°C. Similar temperature characteristics were observed in the Donau valley soil, although it originated from a temperate region. In this soil NO production was stimulated by the addition of NH inf4 sup+ or of NO inf3 sup- . Both NO production and consumption were stimulated by drying and rewetting.  相似文献   

5.
Slurries of anoxic paddy soil were either freshly prepared or were partially depleted in endogenous electron donors by prolonged incubation under anaerobic conditions. Endogenous NO 3 was reduced within 4 h, followed by reduction of Fe3+ and SO 4 2– , and later by production of CH4. Addition of NO 3 slightly inhibited the production of Fe2+ in the depleted but not in the fresh paddy soil. Inhibition was overcome by the addition of H2, acetate, or a mixture of fatty acids (and other compounds), indicating that these compounds served as electron donors for the bacteria reducing NO 3 and/or ferric iron. Addition on NO 3 also inhibited the reduction of SO 4 2– in the depleted paddy soil. This inhibition was only overcome by H2, but not by acetate or a mixture of compounds, indicating that H2 was the predominant electron donor for the bacteria involved in NO 3 and/or SO 4 2– reduction. SO 4 2– reduction was also inhibited by exogenous Fe3+, but only in the depleted paddy soil. This inhibition was overcome by either H2, acetate, or a mixture of compounds, suggesting that they served as electron donors for reduction of Fe3+ and/or SO 4 2+ . CH4 production was inhibited by NO 3 both in depleted and in fresh paddy soil. Fe3+ and SO 4 2– also inhibited methanogenesis, but the inhibition was stronger in the depleted than in the fresh paddy soil. Inhibition of CH4 production was paralleled by a decrease in the steady state concentration of H2 to a level which provided a free enthalpy of less than G=–17 kJ mol-1 CH4 compared to more than G=–32 kJ mol-1 CH4 in the control. The results indicate that in the presence of exogenous fe3+ or SO 4 2+ , methanogenic bacteria were outcompeted for H2 by bacteria reducing Fe3+ or SO 4 2+ .Deceased on 27 December 1992  相似文献   

6.
Two complementary experimental approaches were utilized to examine the extent to which free soil hydrogenases and H2-oxidizing bacteria contribute to the soil uptake of atmospheric H2. First, high affinity hydrogenase activity and H2-oxidizing bacteria were fractionated in non-axenic soil and axenic soil colonized with the high affinity H2-oxidizing bacterium Streptomyces sp. PCB7. Non-axenic soil was fractionated by buoyant density centrifugation. High affinity H2 oxidation activity measured in individual fractions was proportional to the copy number of hhyL gene, specifying the large subunit of putative high affinity [NiFe]-hydrogenases. 2.5% of the hydrogenase activity was recovered in bacteria-free soil extract. Similarly, sequential centrifugation and wet filtrations of strain PCB7-colonized soil dispersed in solubilization buffer caused a loss of the activity, at a ratio proportional to the number of living cells removed. No abiontic hydrogenase activity was detected in bacteria-free fractions. The second experimental approach was designed to verify whether or not the [NiFe]-hydrogenase of strain PCB7 retains high affinity H2 oxidation activity in soil, under the abiontic state. H2 oxidation rates of crude enzyme extract of strain PCB7 measured under aerobic and anaerobic conditions were indistinguishable, indicating that the high affinity hydrogenase of strain PCB7 is oxygen-tolerant. The hydrogenase activity of sterile soil spiked with as much as 0.14 mg(protein) g(soil-dw)−1 was equivalent to the H2-oxidation activity of only 106-107 CFU of strain PCB7 g(soil-dw)−1. Taken together, our results indicate that high affinity hydrogenase activity is proportional to the abundance of H2-oxidizing bacteria in soil and, that abiontic hydrogenases contribute only a few percent of the total high affinity H2 oxidation activity detected in soil.  相似文献   

7.
The soils of the boreal zone, characterized by acidic, low-organic-matter sands in uplands and organic deposits in lowlands, represent unique environments for heavy metals. The mobility and plant uptake of Pb can be substantially different than in other soils. A survey of natural levels of Pb in northern Ontario revealed concentrations of 26 mg kg–1 dry soil and 1.3 mg kg–1 dry blueberry leaf, with an apparent plant/soil concentration ratio (CR) of 0.051. In outdoor lysimeters with an acidic sand profile (pH 4.9) and under a boreal climate, 67% of a pulse of Pb, applied as Pb(NO3)2, was essentially immobile over 4 yr. The 33% that leached may have been mobilized by soluble organic ligands or the N03 companion ion. The solid/liquid partition coefficient (Kd) for this soil, using either applied 210Pb or stable Pb, was very low: 20 L kg–1 The CR for 210Pb in the same soil was correspondingly high: 0.10 for blueberry and 0.059 over all crops studied. In two organic soils, the Kd values were 9 × 103 L kg–1 (Sphagnum, pH 4.8) and 3 × 104 L kg–1 (sedge, pH 5.5) with corresponding CR values of 8 × 10–4 and 0.0085 for blueberries (0.0027 overall in the latter soil). The CR was most closely related to soil cation exchange capacity, although organic matter content and pH were undoubtedly important related factors. In combination, the acidic sand and organic soils of boreal settings represent extremes for the mobility of Pb.  相似文献   

8.
Summary NO production rates, NO uptake rate constants, NO compensation points, and different soil variables were determined for various soil types and different soil horizons, and checked for mutual correlations. NO production was detected in all, and NO comsuption in most soils tested. Only soils in a very early state of soil genesis showed no NO consumption activity. NO consumption was positively correlated with soil water and NH 4 + contents. NO production rates were not correlated with any soil variable. Both NO production and NO consumption tended to decrease from the upper organic to the deeper mineral horizons in different climax soils. The seasonal variation of NO production and NO consumption in a calcic cambisol and a luvisol showed highest rates in summer. The rates of NO production and NO consumption were correlated with a few of the soil variables, but showed no uniform, theoretically comprehensible pattern. However, NO production in samples of the calcic cambisol was stimulated by fertilization with NH 4 + , but not with NO 3 and was inhibited by nitrapyrin, indicating that NO was produced by nitrification. NO production made up about 3% of the nitrification rates. In the luvisol, in contrast, NO production was not affected by the addition of NH 4 + or NO 3 . Nitrification was also undetectable in this acidic soil, except for a few patches where NO production was also detected.  相似文献   

9.
We evaluated the effect of elemental S (S0) under three moisture (40, 60, 120% water-filled pore space; WFPS) and three temperature regimes (12, 24, 36°C) on changes in pH and available P (0.5 N NaHCO3-extractable P) concentrations in acidic (pH 4.9), neutral (pH 7.1) and alkaline (pH 10.2) soils. Repacked soil cores were incubated for 0, 14, 28 and 42 days. Application of S0 did not alter the trends of pH in acidic and neutral soils at all moisture regimes but promoted a decrease in the pH of alkaline soil under aerobic conditions (40%, 60% WFPS). Moisture and temperature had profound effects on the available P concentrations in all three soils, accumulation of available P being greatest under flooded conditions (120% WFPS) at 36°C. Application of S0 in acidic, neutral and alkaline soils resulted in the net accumulation of 16.5, 14.5 and 13 g P g–1 soil after 42 days at 60% WFPS, but had no effect under flooded conditions. The greatest available P accumulations in the respective soils were 19, 19.5 and 20 g P g–1 soil (equivalent to 38, 41, 45 kg P ha–1) with the combined effects of 36°C, 60% WFPS and applied S0. The results of our study revealed that oxidation of S0 lowered the pH of alkaline soil (r=–0.88, P<0.01), which in turn enhanced available P concentrations. Also, considering the significant relationship between the release of sulphate and accumulation of P, even in acidic soil (r=0.92, P<0.01) and neutral soil (r=0.85, P<0.01) where the decrease in pH was smaller, it is possible that the stimulatory effect of sulphate on the availability of P was due to its concurrent desorption from the colloidal surface, release from fixation sites and/or mineralization of organic P. Thus, in the humid tropics and irrigated subtropics where high moisture and temperature regimes are prevalent, the application of S0 could be beneficial not only in alleviating S deficiency in soils but also for enhancing the availability of P in arable soils, irrespective of their initial pH.  相似文献   

10.
Agricultural factors affecting methane oxidation in arable soil   总被引:9,自引:0,他引:9  
CH4 oxidation activity in a sandy soil (Ardoyen) and agricultural practices affecting this oxidation were studied under laboratory conditions. CH4 oxidation in the soil proved to be a biological process. The instantaneous rate of CH4 consumption was in the order of 800 mol CH4 kg–1 day–1 (13 mg CH4 kg–1 day–1) provided the soil was treated with ca. 4.0 mmol CH4 kg–1 soil. Upon repeated supplies of a higher dose of CH4, the oxidation was accelerated to a rate of at least 198 mg CH4 kg–1 day–1. Addition of the plant-growth promoting rhizopseudomonad strains Pseudomonas aeruginosa 7NSK2 and Pseudomonas fluorescens ANP15 significantly decreased the CH4 oxidation by 20 to 30% during a 5-day incubation. However, with further incubation this suppression was no longer detectable. Growing maize plants prevented the suppression of CH4 oxidation. The numbers of methanotrophic bacteria and fungi increased significantly after the addition of CH4, but there were no significant shifts in the population of total bacteria and fluorescent pseudomonads. Drying and rewetting of soil for at least 1 day significantly reduced the activity of the indigenous methanotrophs. Upon rewetting, their activity was regained after a lag phase of about 3 days. The herbicide dichlorophenoxy acetic acid (2,4-D) had a strong negative effect on CH4 oxidation. The application of 5 ppm increased the time for CH4 removal; at concentrations above 25 ppm 2,4-D CH4–oxidizing activity was completely hampered. After 3 days of delay, only the treatments with below 25 ppm 2,4-D showed recovery of CH4–oxidizing activity. This finding suggests that it can be important to include a CH4–removal bioassay in ecotoxicology studies of the side effects of pesticides. Changes in the native soil pH also affected the CH4–oxidizing capacity. Permanent inhibition occurred when the soil pH was altered by 2 pH units, and partial inhibition by 1 pH unit change. A rather narrow pH range (5.9–7.7) appeared to allow CH4 oxidation. Soils pre-incubated with NH 4 + had a lower CH4–removal capacity. Moreover, the nitrification inhibitor 2-chloro-6-trichloromethyl pyridine (nitrapyrin) strongly inhibited CH4 oxidation. Probably methanotrophs rather than nitrifying microorganisms are mainly responsible for CH4 removal in the soil studied. It appears that the causal methanotrophs are remarkably sensitive to soil environmental disturbances.  相似文献   

11.
High rates of cattle slurry application induce NO inf3 sup- leaching from grassland soils. Therefore, field and lysimeter trials were conducted at Gumpenstein (Austria) to determine the residual effect of various rates of cattle slurry on microbial biomass, N mineralization, activities of soil enzymes, root densities, and N leaching in a grassland soil profile (Orthic Luvisol, sandy silt, pH 6.6). The cattle slurry applications corresponded to rates of 0, 96, 240, and 480 kg N ha-1. N leaching was estimated in the lysimeter trial from 1981 to 1991. At a depth of 0.50 m, N leaching was elevated in the plot with the highest slurry application. In October 1991, deeper soil layers (0–10, 10–20, 20–30, 30–40, and 40–50 cm) from control and slurry-amended plots (480 kg N ha-1) were investigated. Soil biological properties decreased with soil depth. N mineralization, nitrification, and enzymes involved in N cycling (protease, deaminase, and urease) were enhanced significantly (P<0.05) at all soil depths of the slurry-amended grassland. High rates of cattle slurry application reduced the weight of root dry matter and changed the root distribution in the different soil layers. In the slurry-amended plots the roots were mainly located in the topsoil (0–10 cm). As a result of this study, low root densities and high N mineralization rates are held to be the main reasons for NO inf3 sup- leaching after heavy slurry applications on grassland.  相似文献   

12.
Soils are the most important sink for atmospheric hydrogen, which is assumed to be oxidized by abiontic soil hydrogenases or by putative high-affinity hydrogenases of microbial origin. The activity of soil hydrogenases has been found to change with soil temperature as it changes during the day and the season. However, it is unclear whether and to which extent the soil hydrogenases are deactivated by increased temperature. Therefore, we incubated soils from different climates and different ecosystems (forest, agricultural, arid, hyper-arid, paddy, peat) at elevated temperature and measured the residual activity at 25 °C after different incubation times. We found that at least part of the soil hydrogenase is deactivated irreversibly already at relatively low temperatures (>30 °C) and short exposure times (>10 min) and that the deactivation was more pronounced in soil from cold versus hot climate. The deactivation kinetics could be fitted to a biexponential model, but were complex with respect to soil type and deactivation temperature. The results show that new hydrogenase activity has to be generated in the soil to compensate for activity loss by diurnal and seasonally increased temperature.  相似文献   

13.
Summary The rate of H2 release from broad beans (Vicia faba) infected with Rhizobium leguminosarum Hup- was much faster than from beans infected with the Hup+ strain. Acetylene reduction and H2 release were abolished by cutting the plants down, by incubation in darkness, or after the addition of ammonium, indicating that the H2 was released by N2-fixing bacterial symbionts. In laboratory cultures using non-sterile soil, the bean plants released H2 until an equilibrium between H2 production and H2 oxidation was reached. The H2 equilibrium concentration was higher in Hup--infected bean cultures (about 3 ppm H2 in the gas phase) than in Hup+-infected cultures (0.3 ppm H2) because of the higher H2 production. The H2 release from Hup--infected bean cultures in sterile soil did not reach equilibrium. An equilibrium occurred, if Knallgas bacteria were added. However, the equilibrium value was higher (13 ppm H2) than in non-sterile soil, which seemed to be more efficient at H2 oxidation. The Knallgas bacteria exhibited a relatively high K m for H2 (> 1300 ppmv H2); this activity was observed in unplanted non-sterile soil, and in nonsterile soil planted with Hup+-infected beans or planted with Hup--infected beans which had been cut down before being assayed. All these soils also showed a second, low-K m (<50 ppm) level of H2 oxidation activity, which was presumably due to abiontic soil enzymes. In contrast, only one level of activity, which had an intermediate K m (about 200 ppm H2), was observed when the soil was planted with Hup--infected beans. The origin of this activity, which was only observed in the presence of intact, H2-producing beans, is still unknown.  相似文献   

14.
Short-timed pH-buffering of disturbed and undisturbed forest soil samples The pH buffering of disturbed and undisturbed soils under spruce (podzol and podzolic cambisol derived from phyllite, eutric cambisol derived from basalt) was studied in the laboratory by adding H2SO4 in ecologically relevant concentrations (pH 5.6–2.0). For the cambisol with crumb structure no difference was found. 80–90% of the added protons were neutralized by release of Ca and Mg. Disturbed samples of the podzol buffer less than 70% of the applicated acid. For undisturbed samples the maximum buffering rate of 185 g H+/ha · h is reached with a proton load of about 500 g H+/ha · h (related to 4 cm soil depth). Buffering behaviour of the podzolic cambisol lies between the podzol and the cambisol. 70–90% of the proton input is buffered in the disturbed samples while the undisturbed one does not reach its maximum buffering rate, even with high proton load. In this soil Al-release is the most reactive buffer.  相似文献   

15.
This study investigated the long-term effect of lime application and tillage systems (no-till, ridge-till and chisel plow) on the activities of arylamidase and amidohydrolases involved in N cycling in soils at four long-term research sites in Iowa, USA. The activities of the following enzymes were assayed: arylamidase, -asparaginase, -glutaminase, amidase, urease, and -aspartase at their optimal pH values. The activities of the enzymes were significantly (P<0.001) and positively correlated with soil pH, with r values ranging from 0.42* to 0.99*** for arylamidase, 0.81*** to 0.97*** for -asparaginase, 0.62*** to 0.97*** for -glutaminase, 0.61*** to 0.98*** for amidase, 0.66** to 0.96*** for urease, and 0.80*** to 0.99*** for -aspartase. The Δactivity/ΔpH values were calculated to assess the sensitivity of the enzymes to changes in soil pH. The order of the sensitivity of enzymes was as follows: -

-aspartase. The enzyme activities were greater in the samples of the 0–5 cm depth than those of the 0–15 cm samples under no-till treatment. Most of the enzyme activities were significantly (P<0.001) and positively correlated with microbial biomass C (Cmic) and N (Nmic). Lime application significantly affected the specific activities of the six enzymes studied. Results showed that soil management practices, including liming and type of tillage significantly affect soil biological and biochemical properties, which may lead to changes in nitrogen cycling, including N mineralization in soils.  相似文献   

16.
The direct application of Sokoto phosphate rock to restore phosphorus in the savanna soil of Nigeria has not been very successful. The dissolution of Sokoto phosphate rock was investigated in three electrolyte solutions – 0.01 m CaCl2, NaCl and KCl – at pH range 3.5–7.0 under laboratory conditions to provide solubility and kinetic data that are required to develop guidelines for direct application in the field. The phosphate rock dissolved in the salt solutions in the order KCl > NaCl > CaCl2. Particle size and ionic strength had no significant effect on the dissolution. The standard free energy of reaction ΔG°R in an acidic solution with no basic cations was ?38 kJ mol?1. If Ca2+ ions were in the acidic solution, then ΔG°R increased to 210 kJ mol?1, 170 kJ mol?1 for Na+ ions, and 107 kJ mol?1 for K+ ions in the solution. The theoretical solubility constant (Ks) calculated from the relation ΔG° = ?RT ln Ks gave Ks = 106.7 in an acidic solution without basic cations, but decreased to 10?36.8 with Ca2+ ions in solution, 10?29.8 with Na+ ions, and 10?18.8 with K+ ions in solution. At pH ≥ 5.5, the dissolution was more constrained by Ca2+ ions or basic cations in solution than by availability of protons. The kinetics of the dissolution reaction was best described by a power function: Ct = atb, where Ct is the amount of P released from the rock phosphate at time t, and a and b are fitting parameters. An Elovich and a parabolic diffusion expression equally gave satisfactory fits to the dissolution data, suggesting that the rate of dissolution was limited by a combination of film‐ and intra‐particle diffusion. To utilize this rock phosphate as an effective source of P, management practices that increase Ca sinks and the supply of protons to the soil are necessary. In the savanna, increasing the soil's organic matter greatly enhances cation exchange capacity and availability of protons. The practice should provide adequate sinks for Ca2+ and the acidic environment required for the release of P from rock phosphate.  相似文献   

17.
Influence of oxygen on production and consumption of nitric oxide in soil   总被引:1,自引:0,他引:1  
Summary NO and N2O release rates were measured in an acidic forest soil (pH 4.0) and a slightly alkaline agricultural soil (pH 7.8), which were incubated at different O2 concentrations (<0.01 – 20% O2) and at different NO concentrations (40 – 1000 ppbv NO). The system allowed the determination of simultaneously operating NO production rates and NO uptake rate constants, and the calculation of a NO compensation concentration. Both NO production and NO consumption decreased with increasing O2. NO consumption decreased to a smaller extent than NO production, so that the NO compensation concentrations also decreased. However, the NO compensation concentrations were not low enough for the soils to become a net sink for atmospheric NO. The release of N2O increased relative to NO release when the gases were allowed to accumulate instead of being flushed out. The forest soil contained only denitrifying, but not nitrifying bacteria, whereas the agricultural soil contained both. Nevertheless, NO release rates were less sensitive to O2 in the forest soil compared to the agricultural soil.  相似文献   

18.
Concentrations of CH4, a potent greenhouse gas, have been increasing in the atmosphere at the rate of 1% per year. The objective of these laboratory studies was to measure the effect of different forms of inorganic N and various N-transformation inhibitors on CH4 oxidation in soil. NH 4 + oxidation was also measured in the presence of the inhibitors to determine whether they had differential activity with respect to CH4 and NH 4 + oxidation. The addition of NH4Cl at 25 g N g-1 soil strongly inhibited (78–89%) CH4 oxidation in the surface layer (0–15 cm) of a fine sandy loam and a sandy clay loam (native shortgrass prairie soils). The nitrification inhibitor nitrapyrin (5 g g-1 soil) inhibited CH4 oxidation as effectively as did NH4Cl in the fine sandy loam (82–89%), but less effectively in the sandy clay loam (52–66%). Acetylene (5 mol mol-1 in soil headspace) had a strong (76–100%) inhibitory effect on CH4 consumption in both soils. The phosphoroamide (urease inhibitor) N-(n-butyl) thiophosphoric triamide (NBPT) showed strong inhibition of CH4 consumption at 25 g g-1 soil in the fine sandy loam (83%) in the sandy clay loam (60%), but NH 4 + oxidation inhibition was weak in both soils (13–17%). The discovery that the urease inhibitor NBPT inhibits CH4 oxidation was unexpected, and the mechanism involved is unknown.  相似文献   

19.
A field study was conducted to investigate the long-term effect of surface application of sewage sludge composts vs chemical N fertilizer on total N, total C, soluble organic C, pH, EC, microbial biomass C and N, protease activity, deaminase activity, urease activity, gross and net rates of N mineralization and nitrification, CO2 evolution, and N2O production. Soil samples were taken from five depths (0–15, 15–20, 20–30, 30–40, and 40–50 cm) of a long-term experiment at the University of Tokyo, Japan. Three fields have been receiving sewage sludge composted with rice husk (RH), sawdust (SD), or mixed chemical fertilizer NPK (CF), applied at the rate of 240 kg N ha–1 each in split applications in summer and autumn since 1978. Significantly higher amounts of total N and C and soluble organic C were found in the compost than in the CF treatments up to the 40-cm soil depth, indicating improved soil quality in the former. In the CF treatment, soil pH values were significantly lower and electrical conductivity values were significantly higher than those of compost-treated soils of up to 50 cm depth. Soil microbial biomass C and N, CO2 evolution, protease, deaminase, and urease activities were significantly higher in the compost than in the CF treatments due to greater availability of organic substrates that stimulated microbial activity. Gross N mineralization rates determined by 15N dilution technique were eight and five times higher in the SD and RH treatments than in the CF treatment, respectively, probably due to high levels of microbial and enzyme activities. Net N mineralization rates were also significantly higher in the compost treatments and were negative in the CF treatment indicating immobilization. Net nitrification rates were higher in compost treatments and negative in the CF treatment. Nitrous oxide productions from compost treatments were higher than the CF treatment due to the greater availability of mineral N as a result of higher mineralization and nitrification rates and soluble organic C in the former. Most of the measured parameters were highest in the surface soil (0–15 cm) and were significantly higher in the SD treatment than in the RH treatment.  相似文献   

20.
Summary Soil enzyme activities (acid and alkaline phosphatase, arylsulfatase, -glucosidase, urease and amidase) were determined (0- to 20-cm depth) after 55 years of crop-residue and N-fertilization treatment in a winter wheat (Triticum aestivum L.)-fallow system on semiarid soils of the Pacific Northwest. All residues were incorporated and the treatments were: straw (N0), straw with fall burn (N0FB), straw with spring burn (N0SB), straw plus 45 kg N ha–1 (N45), straw plus 90 kg N ha–1 (N90), straw burned in spring plus 45 kg N ha–1 (N45SB), straw burned in spring plus 90 kg N ha–1 (N90SB), straw plus 2.24 T ha–1 pea-vine residue and straw plus 22.4 T ha–1 of straw-manure. Enzyme activities were significantly (P<0.001) affected by residue management. The highest activities were observed in the manure treated soil, ranging from 36% (acid phosphatase) to 190% increase in activity over the control (N0). The lowest activities occurred in the N0FB (acid phosphatase, arylsulfatase and -glucosidase) and N90 treated soils (alkaline phosphatase, amidase and urease). Straw-burning had a significant effect only on acid phosphatase activity, which decreased in spring burn treated soil when inorganic N was applied. Urease and amidase activity decreased with long-term addition of inorganic N whereas the pea vine and the manure additions increased urease and amidase activity. There was a highly significant effect from the residue treatments on soil pH. Arylsulfatase, urease, amidase and alkaline phosphatase activities were positively correlated and acid phosphatase activity was negatively correlated with soil pH. Enzyme activities were strongly correlated with soil organic C and total N content. Except for acid phosphatase, there was no significant relationship between enzyme activity and grain yield.Journal Paper No. 8072 of the Agricultural Experimental Station, Oregon State University, Corvallis, OR 97331, USA  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