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.
Suspensions of two strains of Pseudomonas aeruginosa (ON12 and ON12-1) were used to reduce NO3- and NO2-, respectively, to N2O. The evolved N2O was quantified by gas chromatography with electron capture detection, and the 15N abundance was determined by mass spectrometry with a special inlet system and triple-collector detection. Sample gas containing unknown N2O pools as small as 0.5 ng of N was analyzed by use of a spike technique, in which a reference gas of N2O of natural 15N abundance was added to obtain enough total N for the mass spectrometer. In NO3- or NO2- pools, the 15N abundance could be determined in samples as small as approximately 3.5 ng of N. No cross-contamination took place between the NO3- and NO2- pools. The excellent separation of NO3- and NO2- pools, small sample size required, and low contamination risk during N2O analysis offer great advantages in isotope studies of inorganic N transformations by, e.g., nitrifying or denitrifying bacteria in the environment.  相似文献   

2.
A more sensitive analytical method for NO(3) was developed based on the conversion of NO(3) to N(2)O by a denitrifier that could not reduce N(2)O further. The improved detectability resulted from the high sensitivity of the Ni electron capture gas chromatographic detector for N(2)O and the purification of the nitrogen afforded by the transformation of the N to a gaseous product with a low atmospheric background. The selected denitrifier quantitatively converted NO(3) to N(2)O within 10 min. The optimum measurement range was from 0.5 to 50 ppb (50 mug/liter) of NO(3) N, and the detection limit was 0.2 ppb of N. The values measured by the denitrifier method compared well with those measured by the high-pressure liquid chromatographic UV method above 2 ppb of N, which is the detection limit of the latter method. It should be possible to analyze all types of samples for nitrate, except those with inhibiting substances, by this method. To illustrate the use of the denitrifier method, NO(3) concentrations of <2 ppb of NO(3) N were measured in distilled and deionized purified water samples and in anaerobic lake water samples, but were not detected at the surface of the sediment. The denitrifier method was also used to measure the atom% of N in NO(3). This method avoids the incomplete reduction and contamination of the NO(3) -N by the NH(4) and N(2) pools which can occur by the conventional method of NO(3) analysis. N(2)O-producing denitrifier strains were also used to measure the apparent K(m) values for NO(3) use by these organisms. Analysis of N(2)O production by use of a progress curve yielded K(m) values of 1.7 and 1.8 muM NO(3) for the two denitrifier strains studied.  相似文献   

3.
A highly sensitive denitrification bioassay was developed for detection of NO3- and NO2- in rhizosphere soil samples. Denitrifying Pseudomonas aeruginosa ON12 was grown anaerobically in citrate (30 mM) minimal medium with KClO3 (10 mM) and NaNO2 (3 mM), which gave cells capable of NO2- reduction to N2O but incapable of NO3- reduction to NO2-. Growth on citrate minimal medium further resulted in the absence of N2O reduction. When added to small soil samples in O2-free vials, such cells could be used to convert the indigenous NO2- pool to N2O, which was subsequently quantified by gas chromatography. Cells grown in KClO3-free citrate medium with 10 mM NaNO3 as the electron acceptor were capable of reducing both NO3- and NO2-, and these cells could subsequently be added to the sample to convert the indigenous NO3- pool to N2O. Concentrations of both NO3- and NO2- were thus determined as N2O, with a detection limit of approximately 10 pmol of N. The bioassay could be used to determine NO3- and NO2- pools in 10-mg soil samples taken along a microgradient in the rhizosphere of field-grown barley plants. At both low (10%, wt/wt) and high (18%, wt/wt) water content, relatively high levels of NO2- were found in the rhizosphere compared with bulk soil. Under dry conditions, NO3- was also more abundant in the rhizosphere than in the bulk soil, whereas such a difference was not observed at the high water content. The roles of plant metabolism and bacterial nitrification and denitrification processes for NO3- and NO2- availability in the rhizosphere are discussed.  相似文献   

4.
A highly sensitive denitrification bioassay was developed for detection of NO3- and NO2- in rhizosphere soil samples. Denitrifying Pseudomonas aeruginosa ON12 was grown anaerobically in citrate (30 mM) minimal medium with KClO3 (10 mM) and NaNO2 (3 mM), which gave cells capable of NO2- reduction to N2O but incapable of NO3- reduction to NO2-. Growth on citrate minimal medium further resulted in the absence of N2O reduction. When added to small soil samples in O2-free vials, such cells could be used to convert the indigenous NO2- pool to N2O, which was subsequently quantified by gas chromatography. Cells grown in KClO3-free citrate medium with 10 mM NaNO3 as the electron acceptor were capable of reducing both NO3- and NO2-, and these cells could subsequently be added to the sample to convert the indigenous NO3- pool to N2O. Concentrations of both NO3- and NO2- were thus determined as N2O, with a detection limit of approximately 10 pmol of N. The bioassay could be used to determine NO3- and NO2- pools in 10-mg soil samples taken along a microgradient in the rhizosphere of field-grown barley plants. At both low (10%, wt/wt) and high (18%, wt/wt) water content, relatively high levels of NO2- were found in the rhizosphere compared with bulk soil. Under dry conditions, NO3- was also more abundant in the rhizosphere than in the bulk soil, whereas such a difference was not observed at the high water content. The roles of plant metabolism and bacterial nitrification and denitrification processes for NO3- and NO2- availability in the rhizosphere are discussed.  相似文献   

5.
As genomic data for bacteria are unraveled at an increasing speed, there is a need for more efficient and refined techniques to characterize metabolic traits. The regulatory apparatus for denitrification, for instance, has been explored extensively for type strains, but we lack refined observations of how these and wild type denitrifiers respond metabolically to changing environmental conditions. There is a need for new "phenomic" approaches, and the present paper describes one; an automated incubation system for the study of gas kinetics in 15 parallel bacterial cultures. An autosampler with a peristaltic pump takes samples from the headspace, and replaces the sampled gas with He by reversing the pump. The sample flows through the injector of a micro GC (for determination of N(2), O(2), CH(4), CO(2), N(2)O) to the inlet of a chemoluminescence NO analyzer. The linear range for NO is 0.5-10(4) ppmv (CV=2%, detection limit 0.2 ppmv). The gas leakage of N(2) into the system is low and reproducible, allowing the quantification of N(2) production (in flasks with He+O(2) atmosphere) with a detection limit of 150-200 nmol N(2) for a single time increment. The gas loss by each sampling is taken into account, securing mass balance for all gases, thus allowing accurate estimation of electron flows to the various terminal acceptors (O(2), NO(2)(-), NO, N(2)O) throughout the culture's depletion of O(2) and NO(x). We present some experimental results with Agrobacterium tumefaciens, Paracoccus denitrificans and denitrifying communities, demonstrating the system's potential for unraveling contrasting patterns of denitrification gene expression as a function of concentrations of O(2) and NO in the medium.  相似文献   

6.
Dissimilatory reduction of NO(2) to N(2)O and NH(4) by a soil Citrobacter sp. was studied in an attempt to elucidate the physiological and ecological significance of N(2)O production by this mechanism. In batch cultures with defined media, NO(2) reduction to NH(4) was favored by high glucose and low NO(3) concentrations. Nitrous oxide production was greatest at high glucose and intermediate NO(3) concentrations. With succinate as the energy source, little or no NO(2) was reduced to NH(4) but N(2)O was produced. Resting cell suspensions reduced NO(2) simultaneously to N(2)O and free extracellular NH(4). Chloramphenicol prevented the induction of N(2)O-producing activity. The K(m) for NO(2) reduction to N(2)O was estimated to be 0.9 mM NO(2), yet the apparent K(m) for overall NO(2) reduction was considerably lower, no greater than 0.04 mM NO(2). Activities for N(2)O and NH(4) production increased markedly after depletion of NO(3) from the media. Amendment with NO(3) inhibited N(2)O and NH(4) production by molybdate-grown cells but not by tungstate-grown cells. Sulfite inhibited production of NH(4) but not of N(2)O. In a related experiment, three Escherichia coli mutants lacking NADH-dependent nitrite reductase produced N(2)O at rates equal to the wild type. These observations suggest that N(2)O is produced enzymatically but not by the same enzyme system responsible for dissimilatory reduction of NO(2) to NH(4).  相似文献   

7.
Uptake of 15N labelled NH4NO3 by two Sphagnum mosses on a raised bog in north east Scotland was measured at different times of the year. In a field experiment, fortnightly additions of NH4NO3 at natural abundance, equivalent to 3 g N m-2 yr-1, were made over 14 months to cores of Sphagnum capillifolium occupying hummocks and S. recurvum colonizing hollows. Pre-harvested cores were treated with 15NH415NO3 two weeks before harvesting and 15N abundance determined for the total N in the moss, inorganic and dissolved organic N (DON) in the moss water and extractable inorganic, organic and microbial N in the underlying peat. The proportion of added 15N taken up by the mosses two weeks after each addition averaged 72% and ranged between 11 and 100%, tending to be least during October when the rising water table reached the surface, particularly for S. recurvum. A small proportion of the 15N was detected in the moss water as NH4+ (0.01%) and as DON (0.03%) and on occasions a large proportion remained unaccounted for. In waters from S. capillifolium, DON was proportional to the amount of inorganic N added, but this was not the case for S. recurvum. Little or no 15N was detected in the underlying peat partly because of the large size and variability of the NH4+, DON and microbial N pools.  相似文献   

8.
To evaluate the denitrification abilities of many Bradyrhizobium field isolates, we developed a new (15)N-labeled N(2) detection methodology, which is free from interference from atmospheric N(2) contamination. (30)N(2) ((15)N(15)N) and (29)N(2) ((15)N(14)N) were detected as an apparent peak by a gas chromatograph equipped with a thermal conductivity detector with N(2) gas having natural abundance of (15)N (0.366 atom%) as a carrier gas. The detection limit was 0.04% (30)N(2), and the linearity extended at least to 40% (30)N(2). When Bradyrhizobium japonicum USDA110 was grown in cultures anaerobically with (15)NO(3)(-), denitrification product ((30)N(2)) was detected stoichiometrically. A total of 65 isolates of soybean bradyrhizobia from two field sites in Japan were assayed by this method. The denitrification abilities were partly correlated with filed sites, Bradyrhizobium species, and the hup genotype.  相似文献   

9.
The pathway of anaerobic reduction of nitrite to nitrogen gas (N2) by cell suspensions of the denitrifier, Pseudomonas aeruginosa, was studied using the techniques of gas chromatography and mass spectrometry. While release of nitrous oxide (N2O) is not normally detected during the reduction of nitrite to N2 by this organism, 15N from [15N]nitrite nevertheless can be trapped quantitatively as 15N2O in a pool of added N2O. In such experiments the abundance of 15N in N2O always exceeds that in product N2, consistent with the absence of a major reductive route from nitrite to N2 which by-passes N2O. During the reduction of a mixture of [15N]nitrite and nitric oxide (NO), 15NO produced at most only in trace amounts. The final products are chiefly 15N2 and 14N2 with only a small fraction of the scrambled product, 14N15N. Much of the 14N15N can be accounted for as an artifact caused by traces of molecular oxygen, which promote the conversion of NO to nitrite by autooxidation and thereby degrade slightly the isotopic purity of [15N]nitrite. Nitrous oxide shows all the properties of a free obligatory intermediate during the denitrification of nitrite to N2 by P. aeruginosa, whereas NO does not. The inability to trap 15NO in a pool of NO indicates that NO is not a free obligatory intermediate in the reduction of nitrite. The small mole fractions of 14N15N produced from a mixture of [15N]nitrite and NO require that the main reductive pathways for these nitrogen oxides cannot share any freely diffusible mono-nitrogen intermediate in common. The simplest interpretation is that nitrite and NO are denitrified by separate pathways, at least prior to the formation of the first bi-nitrogen compound.  相似文献   

10.
The oxygen control of denitrification and its emission of NO/N2O/N2 was investigated by incubation of Nycodenz-extracted soil bacteria in an incubation robot which monitors O2, NO, N2O and N2 concentrations (in He+O2 atmosphere). Two consecutive incubations were undertaken to determine (1) the regulation of denitrification by O2 and NO2(-) during respiratory O2 depletion and (2) the effects of re-exposure to O2 of cultures with fully expressed denitrification proteome. Early denitrification was only detected (as NO and N2O) at 相似文献   

11.
The photodetachment of NO from [M(II)(CN)5NO]2- with M = Fe, Ru, and Os, upon laser excitation at various wavelengths (355, 420, and 480 nm) was followed by various techniques. The three complexes showed a wavelength-dependent quantum yield of NO production Phi(NO), as measured with an NO-sensitive electrode, the highest values corresponding to the larger photon energies. For the same excitation wavelength the decrease of Phi(NO) at 20 degrees C in the order Fe > Ru > Os, is explained by the increasing M-N bond strength and inertness of the heavier metals. Transient absorption data at 420 nm indicate the formation of the [M(III)(CN)5H2O]2- species in less than ca. 1 micros for M = Fe and Ru. The enthalpy content of [Fe(III)(CN)5H2O]2- with respect to the parent [Fe(II)(CN)5NO]2- state is (190 +/- 20) kJ mol(-1), as measured by laser-induced optoacoustic spectroscopy (LIOAS) upon excitation at 480 nm. The production of [Fe(III)(CN)5H2O]2- is concomitant with an expansion of (8 +/- 3) ml mol(-1) consistent with an expansion of the water bound through hydrogen bonds to the CN ligands plus the difference between NO release into the bulk and water entrance into the first coordination sphere. The activated process, as indicated by the relatively strong temperature dependence of the Phi(NO) values and by the temperature dependence of the appearance of the [Fe(III)(CN)5H2O]2- species, as determined by LIOAS, is attributed to NO detachment in less than ca. 100 ns from the isonitrosyl (ON) ligand (MS1 state).  相似文献   

12.
Pure cultures of the marine ammonium-oxidizing bacterium Nitrosomonas sp. were grown in the laboratory at oxygen partial pressures between 0.005 and 0.2 atm (0.18 to 7 mg/liter). Low oxygen conditions induced a marked decrease in the rate for production of NO(2), from 3.6 x 10 to 0.5 x 10 mmol of NO(2) per cell per day. In contrast, evolution of N(2)O increased from 1 x 10 to 4.3 x 10 mmol of N per cell per day. The yield of N(2)O relative to NO(2) increased from 0.3% to nearly 10% (moles of N in N(2)O per mole of NO(2)) as the oxygen level was reduced, although bacterial growth rates changed by less than 30%. Nitrifying bacteria from the genera Nitrosomonas, Nitrosolobus, Nitrosospira, and Nitrosococcus exhibited similar yields of N(2)O at atmospheric oxygen levels. Nitrite-oxidizing bacteria (Nitrobacter sp.) and the dinoflagellate Exuviaella sp. did not produce detectable quantities of N(2)O during growth. The results support the view that nitrification is an important source of N(2)O in the environment.  相似文献   

13.
长效碳酸氢铵对土壤硝化-反硝化过程和NO与N2O排放的影响   总被引:3,自引:0,他引:3  
Compared with ammonium bicarbonate(AB), the effect of modified ammonium bicarbonate (MAB) on nitrification and denitrification processes and NO and N2O emissions in a clay soil (C soil) and a loam soil (L soil) was studied in laboratory (25 degrees C and 50% WFPS). The inhibition effect of DCD from MAB on nitrification was relatively small in C soil, but considerably great in L soil. Compared with AB, MAB extended 7 days and 33 days for retaining NH4+. During 15 days, the NO emission from C soil and L soil respectively accounted for 0.60% and 1.06% of applied N under AB application (100 micrograms N.g-1), which were as 30 and 12 times as the N2O emission from corresponding soils. After applying MAB, the emission of NO from C soil and L soil decreased by 67% and 95%, and the emission of N2O decreased by 64% and 95%, respectively. After 39 days of aerobic incubation, then anaerobically flooded incubation with nitrate addition (200 micrograms KNO3-N.g-1) for 7 days, the total loss of denitrification in MAB in L soil was 50% less, and N2O emission was 113% more than in AB in same soil.  相似文献   

14.
15.
A sensitive NO2- biosensor that is based on bacterial reduction of NO2- to N2O and subsequent detection of the N2O by a built-in electrochemical N2O sensor was developed. Four different denitrifying organisms lacking NO3- reductase activity were assessed for use in the biosensor. The relevant physiological aspects examined included denitrifying characteristics, growth rate, NO2- tolerance, and temperature and salinity effects on the growth rate. Two organisms were successfully used in the biosensor. The preferred organism was Stenotrophomonas nitritireducens, which is an organism with a denitrifying pathway deficient in both NO3- and N2O reductases. Alternatively Alcaligenes faecalis could be used when acetylene was added to inhibit its N2O reductase. The macroscale biosensors constructed exhibited a linear NO2- response at concentrations up to 1 to 2 mM. The detection limit was around 1 microM NO2-, and the 90% response time was 0.5 to 3 min. The sensor signal was specific for NO2-, and interference was observed only with NH2OH, NO, N2O, and H2S. The sensor signal was affected by changes in temperature and salinity, and calibration had to be performed in a system with a temperature and an ionic strength comparable to those of the medium analyzed. A broad range of water bodies could be analyzed with the biosensor, including freshwater systems, marine systems, and oxic-anoxic wastewaters. The NO2- biosensor was successfully used for long-term online monitoring in wastewater. Microscale versions of the NO2- biosensor were constructed and used to measure NO2- profiles in marine sediment.  相似文献   

16.
Nitrite reductase (cytochrome c,d1) was purified from Pseudomonas aeruginosa. In the presence of the reducing system, ascorbate-N,N,N',N'-tetramethylphenyl-enediamine, which alone had no ability to reduce nitrite or NO at pH 7.5, the enzyme catalyzed the reduction of nitrite to NO and N2O as major and minor products, respectively, as determined by gas chromatography-mass spectrometry. The rate of reduction of NO to N2O was considerably lower than the rate of reduction of nitrite to N2O and might be zero. The N2O produced in a system containing [15N]nitrite and natural NO was more highly enriched in 15N than was the NO pool and, in this regard, closely resembled the enrichment of the nitrite pool. The amount of 14N in the NO pool changed little, if any, as the result of enzymatic processes. For the enzyme, free NO seems not to be an intermediate between nitrite and N2O, just as was found by this laboratory for certain intact denitrifying bacteria. The results are consistent with reduction of nitrite to enzyme-bound NO, which can partition between release and further reduction.  相似文献   

17.
Ions of structure X[N(O)NO]-, examples of which have seen increasing use as probes for studying the biology of nitric oxide (NO) over the past decade, have a varied chemical history spanning nearly two centuries. Nevertheless, they have not been widely appreciated for their physicochemical similarities. Here we begin a series of systematic inquiries into the fundamental chemistry of such compounds aimed at identifying both the characteristics that justify considering them as a group and the factors that contribute to observed differences in their physicochemical properties. In the present paper, X-ray structures in which X is SO3- (1), O- (2), Ph (3), and Et2N (5), as well as that of the gem-disubstituted carbon derivative CH2[N(O)NO]2-(2) (4), are compared. All their O-N-N-O systems are essentially planar, with cis oxygens and an N-N linkage exhibiting considerable double-bond character. The ultraviolet spectrum of the isolated chromophore consists of a relatively intense ( approximately 6-10 mM(-1) x cm(-1) per [N(O)NO]- group) absorption at 248-250 nm (for 2 and 5) that is red shifted by through-space Stark interactions (e.g., by approximately 10 nm in 1 and 4) as well as by conjugative interaction with X (lambda(max) = 284 nm for 3). Infrared and Raman spectra for the widely used pharmacological probe 5 were determined, with analysis of vibrational modes being aided by comparison with the spectra of the [15N(O)15NO]- isotopomer and density functional theory calculations at the B3LYP/6-311++G** level. To address confusion that has arisen in the literature resulting from rather widespread use of differing trivial designations for this class of compounds, a unifying nomenclature system is recommended in which compounds containing the [N(O)NO]- moiety are named as diazeniumdiolates. It is hoped that these and other efforts to understand and predict the physicochemical similarities and differences among different members of the diazeniumdiolate class will aid in reaping their full potential in the area of rational drug design.  相似文献   

18.
Despite numerous approaches to measuring nitric oxide ((.-)NO) formation from purified NO synthase (NOS), it is still not clear whether (.-)NO is a direct or indirect product of the NO synthase reaction. The direct detection of catalytically formed (.-)NO is complicated by side reactions with reactive oxide species like H(2)O(2) and superoxide. The aim of the present study was therefore to reinvestigate these reactions both electrochemically and by chemiluminescence detection with particular emphasis on the requirement for cofactors and their interference with (.-)NO detection. Flavins were found to generate large amounts of H(2)O(2) and were therefore excluded from subsequent incubations. Under conditions of both coupled and uncoupled catalysis, SOD was absolutely required to detect (.-)NO from NOS. H(2)O(2) formation took place also in the presence of SOD and gave a smaller yet significant interfering signal. Similar data were obtained when the proposed intermediate N(omega)-hydroxy-l-arginine was utilized as substrate. In conclusion, standard Clark-type ()NO electrodes are cross-sensitive to H(2)O(2) and therefore both SOD and catalase are absolutely required to specifically detect (.-)NO from NOS.  相似文献   

19.
McCalley CK  Sparks JP 《Oecologia》2008,156(4):871-881
Emissions of reactive N compounds produced during terrestrial N cycling can be an important N loss pathway from ecosystems. Most measurements of this process focus on NO and N(2)O efflux; however, in alkaline soils such as those in the Mojave Desert, NH(3) production can be an important component of N gas loss. We investigated patterns of NO and NH(3) emissions in the Mojave Desert and identified seasonal changes in temperature, precipitation and spatial heterogeneity in soil nutrients as primary controllers of soil efflux. Across all seasons, NH(3) dominated reactive N gas emissions with fluxes ranging from 0.9 to 10 ng N m(-2) s(-1) as compared to NO fluxes of 0.08-1.9 ng N m(-2) s(-1). Fluxes were higher in April and July than in October; however, a fall precipitation event yielded large increases in both NO and NH(3) efflux. To explore the mechanisms driving field observations, we combined NO and NH(3) soil flux measurements with laboratory manipulations of temperature, water and nutrient conditions. These experiments showed a large transient NH(3) pulse (~70-100 ng N m(-2) s(-1)) following water addition, presumably driven by an increase in soil NH(4) (+) concentrations. This was followed by an increase in NO production, with maximum NO flux rates of 34 ng N m(-2) s(-1). Our study suggests that immediately following water addition NH(3) volatilization proceeds at high rates due to the absence of microbial competition for NH(4) (+); during this period N gas loss is insensitive to changes in temperature and soil nutrients. Subsequently, NO emission increases and rates of both NO and NH(3) emission are sensitive to temperature and nutrient constraints on microbial activity. Addition of labile C reduces gaseous N losses, presumably by increasing microbial immobilization, whereas addition of NO(3) (-) stimulates NO and NH(3) efflux.  相似文献   

20.
To react with peptides, nitric oxide.NO has to be activated by oxidation, or by coupling with superoxide (O.-2) thereby producing peroxynitrite. In the course of.NO oxidation,.NO2 free radicals and N2O3 may be formed. Using gamma-irradiation methods, we characterized the products formed by these nitrogen oxides with angiotensin II. Angiotensin II is specifically nitrated at its tyrosinyl residue by.NO2 or peroxynitrite. Equimolecular amounts of each reagent in K+/Pi solutions at pH 7.4 led to 56% and 5% nitration yields, respectively. Nitrogen oxides produced by autoxidation of.NO, as well as.NO2 under.NO, reacted only with the arginine residue, giving a mixture of peptides containing citrulline, a N-(hydroxylamino-cyanamido-) instead of guanido group, and a conjugated diene derived from an arginine side-chain. However, nitrosation reactions by N2O3 occurred only when the initial concentration of.NO2 was 10 times that able to react with angiotensin II. Thus, in this case.NO appears to protect against.NO2 action.  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