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.
The CH3 + ion, formed in ionized methane, undergoes consecutive eliminative condensation reactions with methane to form the carbonium ions C2H5 +, i-C3H7 + and t-C4H9 +. AtT<500°K, \(N_{CH_4 } \) ?1016 cm?3 these ions react with NH3 in competitive condensation-H+ transfer reactions, e.g. $$\begin{gathered} C_2 H_5 ^ + + NH_3 \xrightarrow{M} C_2 H_5 NH_3 ^ + \hfill \\ - - - \to NH_4 ^ + + C_2 H_4 \hfill \\ \end{gathered} $$ At particle densities of \(N_{CH_4 } \) <1016 cm?3 proton transfer is the only significant reaction channel. At \(N_{CH_4 } \) >1017 cm?3 condensation constitutes 5–20% of the overall reactions. The product of the condensation reaction further associates with CO2 to form C2H5NH3 +·CO2; the atomic composition of this cluster ion is identical with the protonated amino acid alanine. The carbonium ions i-C3H7 + and t-C4H9 + condense also with HCN to yield protonated isocyanides. HCNH+ also appears to condense with HCN atT>570°K, and form cluster ions with HCN at lower temperatures. The rate constants of the condensation reactions vary with temperature and pressure in a complex manner. Under conditions similar to those on Titan at an altitude of 100 km (T=100–150°K, \(N_{CH_4 } \) ≈1018 cm?3), with a methane atmosphere containing 1% H2 and traces of NH3 and H2O, ion-molecule condensation reactions followed by H+ transfer are expected to lead to the atmospheric synthesis of C2H6, C3H8, CH3OH, C2H5OH and the terminal ions NH4 +, CH3NH3 + and C2H5NH3 +. At higher temperatures (250°K<T<400°K), the synthesis of i-C4H10, i-C3H7OH and t-C4H9OH and of the ions i-C3H7NH3 + and t-C4H9NH3 + is also expected. Electron recombination of the terminal ions may yield amines, imines and nitriles. Cycles of protonation and dissociative recombination of the alkanes and alcohols produced in condensation reactions will also produce unsaturated hydrocarbons, ketones and aldehydes in the ionized atmosphere.  相似文献   

2.
Photolysis of NH3 at 185 nm in the presence of a two-fold excess of CH4 results in the loss of about 0.25 mole of CH4 per mole of NH3 decomposed (ΔCH4/ΔNH3). The loss arises from the abstraction of hydrogen atoms from CH4 by photolytically generated hot hydrogen atoms, the presence of which is established by the constancy of ΔCH4/ΔNH3 between 298 and 156 K and by the quenching of the abstraction reaction when either H2 or SF6 is added. From the latter result, it can be concluded that NH3 photolysis in the H2-abundant atmosphere of Jupiter is not responsible for the presence of the carbon compounds observed there such as ethane, acetylene, and hydrogen cyanide, but may have had a role in the early atmosphere of Titan. Photolysis of PH3 with a 206 nm light source gives P2H4, which in turn is converted to a red-brown solid (P4?). The course of the photolysis is not changed appreciably when the temperature is lowered to 157 K except that the concentration of P2H4 increases. The presence of H2 has no effect on the P2H4 yield. Photolysis of 9∶1 NH3∶PH3 gives a rate of decomposition of PH3 that is comparable with that observed by the direct photolysis of PH3. Comparable amounts of P2H4 and the red-brown solid are also observed. The mechanisms of these photochemical reactions together with their implications to the atmospheric chemistry of Jupiter are discussed. The structures of the compounds responsible for the wide array of colorse.g., brown, red and white, observed in the atmosphere of Jupiter have been the subject of extensive speculation. One theory suggests that these colors are due to organic materials formed by the action of either solar ultraviolet light or electric discharges on mixtures of CH4, NH3 and NH4HS in the Jovian atmosphere (Ponnamperuma, 1976; Khareet al., 1978). An alternative hypothesis is that the colors are due to inorganic compounds resulting from the photolysis of NH4HS and PH3 (Lewis and Prinn, 1970; Prinn and Lewis, 1975). In this paper we will summarize our experiments which were designed to test some of these hypotheses.  相似文献   

3.
Yields based on carbon are usually reported in prebiotic experiments, while energy yields (moles cal–1) are more useful in estimating the yields of products that would have been obtained from the primitive atmosphere of the earth. Energy yields for the synthesis of HCN and H2CO from a spark discharge were determined for various mixtures of CH4, CO, CO2, H2, H2O, N2 and NH3. The maximum yields of HCN and H2CO from CH4, CO, and CO2 as carbon sources are about 4×10–8 moles cal–1.  相似文献   

4.
We studied the influence of inorganic nitrogen sources (NO3 ? or NH4 +) and potassium deficiency on expression and activity of plasma membrane (PM) H+-ATPase in sorghum roots. After 15 d of cultivation at 0.2 mM K+, the plants were transferred to solutions lacking K+ for 2 d. Then, K+ depletion assays were performed in the presence or absence of vanadate. Further, PMs from K+-starved roots were extracted and used for the kinetic characterization of ATP hydrolytic activity and the immunodetection of PM H+-ATPase. Two major genes coding PM H+-ATPase (SBA1 and SBA2) were analyzed by real-time PCR. PM H+-ATPase exhibited a higher Vmax and Km in NH4 +-fed roots compared with NO3 ? -fed roots. The optimum pH of the enzyme was slightly lower in NO3 ? -fed roots than in NH4 +-fed roots. The vanadate sensitivity was similar. The expressions of SBA1 and SBA2 increased in roots grown under NH4 +. Concomitantly, an increased content of the enzyme in PM was observed. The initial rate of K+ uptake did not differ between plants grown with NO3 ? or NH4 +, but it was significantly reduced by vanadate in NH4 +-grown plants.  相似文献   

5.
Laboratory studies provide a fundamental understanding of photochemical processes in planetary atmospheres. Photochemical reactions taking place on giant planets like Jupiter and possibly comets and the interstellar medium are the subject of this research. Reaction pathways are proposed for the coupled photochemistry of NH3 (ammonia) and C2H2 (acetylene) within the context Jupiter’s atmosphere. We then extend the discussion to the Great Red Spot, Extra-Solar Giant Planets, Comets and Interstellar Organic Synthesis. Reaction rates in the form of quantum yields were measured for the decomposition of reactants and the formation of products and stable intermediates: HCN (hydrogen cyanide), CH3CN (acetonitrile), CH3CH = N-N = CHCH3 (acetaldazine), CH3CH = N-NH2 (acetaldehyde hydrazone), C2H5NH2 (ethylamine), CH3NH2 (methylamine) and C2H4 (ethene) in the photolysis of NH3/C2H2 mixtures. Some of these compounds, formed in our investigation of pathways for HCN synthesis, were not encountered previously in observational, theoretical or laboratory photochemical studies. The quantum yields obtained allowed for the formulation of a reaction mechanism that attempts to explain the observed results under varying experimental conditions. In general, the results of this work are consistent with the initial observations of Ferris and Ishikawa (1988). However, their proposed reaction pathway which centers on the photolysis of CH3CH = N-N = CHCH3 does not explain all of the results obtained in this study. The formation of CH3CH = N-N = CHCH3 by a radical combination reaction of CH3CH = N? was shown in this work to be inconsistent with other experiments where the CH3CH = N? radical is thought to form but where no CH3CH = N-N = CHCH3 was detected. The importance of the role of H atom abstraction reactions was demonstrated and an alternative pathway for CH3CH = N-N = CHCH3 formation involving nucleophilic reaction between N2H4 and CH3CH = NH is advanced.  相似文献   

6.
7.
Four methods of monitoring the anaerobic digestion process were studied at pilot scale. The methods employed were Micro Gas Chromatography (μ-GC) and Membrane Inlet Mass Spectrometry (MIMS) for measurements in the gas phase, Near Infrared Spectroscopy (NIRS) and pH in the liquid phase. Micro Gas Chromatography accurately measured H2, CH4, H2S, N2 and O2 in the headspace whereas the MIMS accurately measured CH4, CO2, H2S, reduced organic sulfur compounds and p-cresol, also in the headspace. In the liquid phase, NIRS was found to be suitable for estimating the concentrations of acetate, propionate and total volatile fatty acids (VFA) but the error of prediction was too large for accurate quantification. Both the μ-GC and NIRS were low maintenance methods whereas the MIMS required frequent cleaning and background measurements.  相似文献   

8.
The prospects of a control for a novel gallium nitride pseudo-halide vapor phase epitaxy (PHVPE) with HCN were thoroughly analyzed for hydrocarbons–NH3–Ga gas phase on the basis of quantum chemical investigation with DFT (B3LYP, B3LYP with D3 empirical correction on dispersion interaction) and ab-initio (CASSCF, coupled clusters, and multireference configuration interaction including MRCI+Q) methods. The computational screening of reactions for different hydrocarbons (CH4, C2H6, C3H8, C2H4, and C2H2) as readily available carbon precursors for HCN formation, potential chemical transport agents, and for controlled carbon doping of deposited GaN was carried out with the B3LYP method in conjunction with basis sets up to aug-cc-pVTZ. The gas phase intermediates for the reactions in the Ga-hydrocarbon systems were predicted at different theory levels. The located π-complexes Ga…C2H2 and Ga…C2H4 were studied to determine a probable catalytic activity in reactions with NH3. A limited influence of the carbon-containing atmosphere was exhibited for the carbon doping of GaN crystal in the conventional GaN chemical vapor deposition (CVD) process with hydrocarbons injected in the gas phase. Our results provide a basis for experimental studies of GaN crystal growth with C2H4 and C2H2 as auxiliary carbon reagents for the Ga-NH3 and Ga-C-NH3 CVD systems and prerequisites for reactor design to enhance and control the PHVPE process through the HCN synthesis.  相似文献   

9.
The effects of mineral nutrient were examined on in vitro growth of Gerbera hybrida (G. jamesonii?×?G. viridifolia), specifically Gerbera hybrida cv. Pasadena. Four types of experiments were conducted to quantify the effects of mineral nutrients on four in vitro growth responses (quality, shoot number, leaf number, and shoot height) of gerbera and included groups of mineral nutrients (macros/mesos, micros, and Fe), individual salts (CuSO4·5H2O, MnSO4·4H2O, ZnSO4·7H2O, and Fe/EDTA), and the specific ions NO3 ?, NH4 +, and K+. Experiments included mixture-amount designs that are essential for separating the effects of proportion and concentration. Highly significant effects were observed in all experiments, but the mineral nutrients with the largest effects varied among the four growth responses. For example, leaf number was strongly affected by the macronutrient group in one experiment and by NH4 + and K+, which were in the macronutrient group, in the NO3 ?/NH4 +/K+ ion-specific experiment, whereas quality was strongly affected by the micronutrients ZnSO4 and Fe/EDTA. Because mineral nutrient effects varied significantly with the response measured, defining an appropriate formulation requires a clear definition of “optimal” growth.  相似文献   

10.
The reaction mechanisms and rates for the H abstraction reactions between CH3SS and CN radicals in the gas phase were investigated with density functional theory (DFT) methods. The geometries, harmonic vibrational frequencies, and energies of all stationary points were obtained at B3PW91/6-311G(d,p) level of theory. Relationships between the reactants, intermediates, transition states and products were confirmed, with the frequency and the intrinsic reaction coordinate (IRC) analysis at the same theoretical level. High accurate energy information was provided by the G3(MP2) method combined with the standard statistical thermodynamics. Gibbs free energies at 298.15 K for all of the reaction steps were reported, and were used to describe the profile diagrams of the potential energy surface. The rate constants were evaluated with both the classical transition state theory and the canonical variational transition state theory, in which the small-curvature tunneling correction was included. A total number of 9 intermediates (IMs) and 17 transition states (TSs) were obtained. It is shown that IM1 is the most stable intermediate by the largest energy release, and the channel of CH3SS?+?CN?→?IM3?→?TS10?→?P1(CH2SS?+?HCN) is the dominant reaction with the lowest energy barrier of 144.7 kJ mol?1. The fitted Arrhenius expressions of the calculated CVT/SCT rate constants for the rate-determining step of the favorable channel is k =7.73?×?106? T 1.40exp(?14,423.8/T) s?1 in the temperature range of 200–2000 K. The apparent activation energy E a(app.) for the main channel is ?102.5 kJ mol?1, which is comparable with the G3(MP2) energy barrier of ?91.8 kJ mol?1 of TS10 (relative to the reactants).  相似文献   

11.
Biosurfactans are amphiphilic compounds synthesized by a wide group of microorganisms and tend to interact with surfaces of different polarities. In the present study we purified and characterized a biosurfactant produced by Dietzia cinnamea KA-1 when cultured by n-hexadecane as sole carbon source. The crude biosurfactant was extracted with ethyl acetate and purified by freezing at–20°C and then silica Gel column chromatography. The purified biosurfactant applied for more characterization using Elemental analysis (CHNS), Fourier Transform Infrared Spectroscopy (FTIR), Mass Spectroscopy (MS) and Nuclear Magnetic Resonance (1H and 13C-NMR) analysis. CHNS analysis showed the presence of C (74.92%) and H (11.63%) but not N or S. Functional groups of OH, CH2, CH3, C=O and aliphatic C?O revealed by FTIR analysis. The presence and position of these groups were confirmed by NMR analysis, and molecular mass of biosurfactant calculated using MS analysis. Finally, the product characterized as a methylated ester compound with molecular formula of C21H42O4. This is the first report of biosurfactant of species D. cinnamea identified as ester, furthermore the ester was found to be in the methylated form.  相似文献   

12.
The reaction of the racemic chiral methyl complex (η5-C5H5)Re(NO)(PPh3)(CH3) (1) with CF3SO3H and then NH2CH2C6H5 gives [(η5-C5H5)Re(NO)(PPh3)(NH2CH2C6H5)]+ ([4a-H]+; 73%), and deprotonation with t-BuOK affords the amido complex (η5-C5H5)Re(NO)(PPh3)(NHCH2C6H5) (76%). Reactions of 1 with Ph3C+ X and then primary or secondary amines give [(η5-C5H5)Re(NO)(PPh3)(CH2NHRR′)]+ X ([6-H]+ X; R/R′/X = a, H/NH2CH2C6H5/BF4; a′, H/NH2CH2C6H5/PF6; b, H/NH2CH2(CH2)2CH3/PF6; c, H/(S)-NH2CH(CH3)C6H5/BF4); d, CH2CH3/CH2CH3/PF6; e, CH2(CH2)2CH3/CH2(CH2)2CH3/PF6; f, CH2C6H5/CH2C6H5/PF6; g, -CH2(CH2)2CH2-/PF6; h, -CH2(CH2)3CH2-/PF6; i, CH3/CH2CH2OH/PF6 (62-99%). Deprotonations with t-BuOK afford the amines (η5-C5H5)Re(NO)(PPh3)(CH2NRR′) (6a-i; 99-40%), which are more stable and isolated in analytically pure form when R ≠ H. Enantiopure 1 is used to prepare (RReSC)-[6c-H]+, (RReSC)-6c, (S)-[6g-H]+, and (S)-6g. The crystal structures of [4a-H]+, a previously prepared NH2CH2Si(CH3)3 analog, [6a′,d,f,h-H]+, (RReSC)-6c, and 6f are determined and analyzed in detail, particularly with respect to cation/anion hydrogen bonding and conformation. In contrast to analogous rhenium containing phosphines, 6a-i show poor activities in reactions that are catalyzed by organic amines.  相似文献   

13.
The photochemical reaction of HCN at 184.9 nm is studied in the gas phase. (CN)2, H2, CH4, NH3, N2H4, C2H6, and CH3NH2 are identified as gas phase products, and a reaction mechanism is proposed. HCN polymers are also obtained as solid reaction products, and their structure is investigated by Infrared Spectorscopy, UV-Visible Spectroscopy, Mass Spectrometry, and Amino Acid Analysis. The process and nature of the formation of the polymers are discussed.  相似文献   

14.
The solvatothermal reactions of V2O5, the appropriate organoamine and HF in the temperature range 100-180 °C yielded a series of vanadium fluorides and oxyfluorides. The compounds [NH4][H3N(CH2)2NH3][VF6] (1) and [H3N(CH2)2NH3][VF5(H2O)] (2) contain mononuclear V(III) anions, while [H3N(CH2)2NH2(CH2)2NH3]2 [VF5(H2O)]2[VOF4(H2O)] (3) exhibits both V(IV) and V(III) mononuclear anions. Both compound 4, [H3NCH2(C6H4)CH2NH3][VOF4]·H2O (4·H2O) and compound 5, [HN(C2H4)3NH][V2O2F6 (H2O)2] (5) contain binuclear anions constructed from edge-sharing V(IV) octahedra. In contrast, [H3N(CH2)2NH2(CH2)2NH3]2[V4O4F14(H2O)2], (6) exhibits a tetranuclear unit of edge- and corner-sharing V(IV) octahedra. Compound 7, [H3N(CH2)2NH2][VF5], contains chains of corner-sharing {VIVF6} octahedra, while [H2N(C2H4)2NH2]3[V4F17O]·1.5H2O (8·1.5H2O) is two-dimensional with a layer of V(III) and V(IV) octahedra in an edge- and corner-sharing arrangement. In the case of [H3N(CH2)2NH3][V2O6] (9), there was no fluoride incorporation, and the anion is a one-dimensional chain of corner-sharing V(V) tetrahedra.  相似文献   

15.
The combined effects of arbuscular mycorrhizal fungi (AMF) and low temperature (LT) on cucumber plants were investigated with respect to biomass production, H2O2 accumulation, NADPH oxidase, ATPase activity and related gene expression. Mycorrhizal colonization ratio was gradually increased after AMF-inoculation. However, LT significantly decreased mycorrhizal colonization ability and mycorrhizal dependency. Regardless of temperature, the total fresh and dry mass, and root activity of AMF-inoculated plants were significantly higher than that of the non-AMF control. The H2O2 accumulation in AMF-inoculated roots was decreased by 42.44 % compared with the control under LT. H2O2 predominantly accumulated on the cell walls of apoplast but was hardly detectable in the cytosol or organelles of roots. Again, NADPH oxidase activity involved in H2O2 production was significantly reduced by AMF inoculation under LT. AMF-inoculation remarkably increased the activities of P-type H+-ATPase, P-Ca2+-ATPase, V-type H+-ATPase, total ATPase activity, ATP concentration and plasma membrane protein content in the roots under LT. Additionally, ATP concentration and expression of plasma membrane ATPase genes were increased by AMF-inoculation. These results indicate that NADPH oxidase and ATPase might play an important role in AMF-mediated tolerance to chilling stress, thereby maintaining a lower H2O2 accumulation in the roots of cucumber.  相似文献   

16.
《Inorganica chimica acta》1988,148(2):233-240
The complexes CodptX3 and [Codpt(H2O)X2]ClO4 (X = Cl, Br; dpt = dipropylenetriamine = NH(CH2CH2CH2NH2)2) have been prepared and characterized. Rate constants (s−1) for aqueous solution at 25 °C and μ = 0.5 M (NaClO4), for the acid-independent sequential ractions.
have been measured spectrophotometrically. For X = Cl: k1 ⋍ 2 × 10−2, k2 = 1.7 × 10−4 and k3 = 4.8 × 10−6, and for X = Br: k1 ⋍ 2 × 10−2, k2 = 5.25 × 10−4 and k3 = 2.5 × 10−5 The primary equation was found to be acid independent, while the secondary and tertiary aquations were acid-inhibited reactions. For the second step, the rate of the reaction was given by the rate equation
where Ct is the complex concentration in the aqua-and hydroxodihalo species, k2 is the rate constant for the acid-dependent pathway and Ka is the equilibrium constant between the hydroxo and aqua complex ions. The activation parameters were evaluated, for X = Cl: ΔH2 = 106.3 ± 0.4 kJ mol−1 and ΔS2 = 40.2 ± 1.7 J K−1 mol, and for X = Br: ΔH2 = 91.6 ± 0.4 kJ mol−1 and ΔS2 = 0.4 ± 1.7 J K−1 mol−1. The results are discussed and detailed comparisons of the reactivities of these complexes with other haloaminecobalt(III) species are presented.  相似文献   

17.
Using density functional theory calculations, we investigated properties of a functionalized BC2N nanotube with NH3 and five other NH2-X molecules in which one of the hydrogen atoms of NH3 is substituted by X = ?CH3, ?CH2CH3, ?COOH, ?CH2COOH and ?CH2CN functional groups. It was found that NH3 can be preferentially adsorbed on top of the boron atom, with adsorption energy of ?12.0 kcal mol?1. The trend of adsorption-energy change can be correlated with the trend of relative electron-withdrawing or -donating capability of the functional groups. The adsorption energies are calculated to be in the range of ?1.8 to ?14.2 kcal mol?1, and their relative magnitude order is found as follows: H2N(CH2CH3) > H2N(CH3) > NH3 > H2N(CH2COOH) > H2N(CH2CN) > H2N(COOH). Overall, the functionalization of BC2N nanotube with the amino groups results in little change in its electronic properties. The preservation of electronic properties of BC2N coupled with the enhancement of solubility renders their chemical modification with either NH3 or amino functional groups to be a way for the purification of BC2N nanotubes.  相似文献   

18.

Key message

NH 4 + acts as a mild oxidative stressor, which triggers antioxidant cellular machinery and provide resistance to salinity.

Abstract

NH4 + nutrition in Carrizo citrange (Citrus sinensis L. Osbeck × Poncirus trifoliata L) plants acts as an inducer of resistance against salinity conditions. NH4 + treatment triggers mild chronic stress that primes plant defence responses by stress imprinting and confers protection against subsequent salt stress. In this work, we studied the influence of NH4 + nutrition on antioxidant enzymatic activities and metabolites involved in detoxification of reactive oxygen species (ROS) to clarify their involvement in NH4 +-mediated salt resistance. Our results showed that NH4 + nutrition induces in citrus plants high levels of H2O2, strongly inhibits superoxide dismutase (SOD) and glutathione reductase (GR) activities, and leads to higher content of oxidised glutathione (GSSG) than in control plants in the absence of salt, thus providing evidence to confirm mild stress induced by NH4 + nutrition. However, upon salinity, plants grown with NH4 + (N-NH4 + plants) showed a reduction of H2O2 levels in parallel to an increase of catalase (CAT), SOD, and GR activities compared with the control plants. Moreover, N-NH4 + plants were able to keep high levels of reduced glutathione (GSH) upon salinity and were able to induce glutathione-S-transferase (GST) and phospholipid hydroperoxide glutathione peroxidise (PHGPx) mRNA accumulation. Based on this evidence, we confirm that sublethal concentrations of NH4 + might act as a mild oxidative stressor, which triggers antioxidant cellular machinery that can provide resistance to subsequent salt stress.  相似文献   

19.
The acidic reduction of Hg using a continuous-flow analytical system was evaluated. With 25% SnCl2 as the reductant, characteristic concentrations (sensitivities) of 0.44 μg/L (open cell) and 0.29 μg/L (flow-through cell) were obtained using inorganic Hg2+ standards in 1.5% HCl. When CH3Hg+ standards were used, absorption signals were an order of magnitude lower, indicating that Sn(II) is incapable of producing Hg° from organic Hg in this acidic reduction system. Addition of CdCl2 to the SnCl2 reductant, as suggested by Magos (1) for the reduction of organomercurials under alkaline conditions, was without beneficial effect. Similarly, combining Sn, with another reducing agent (hydroxylamine hydrochloride), or a strong alkaline solution (40% NaOH), in the reaction coil of the flow-through system did not significantly enhance the Hg absorption signal for either inorganic or organic Hg. Recovery of Hg from spiked liver homogenates digested at 70–80°C using a HNO3/H2SO4/HCl procedure and stabilized with 0.5 mM K2Cr2O7 was >85%, using either inorganic Hg2+ or CH3Hg+, indicating that this digestion procedure successfully breaks the C-Hg bond to form readily reducible Hg species. Usingl-cysteine to stabilize standards of inorganic Hg2+ in HCl caused significant depressions of the Hg absorption signal atl-cysteine concentrations >0.001% (≈0.5 mM); 0.1%l-cysteine caused total suppression of the Hg signal. These results indicate that: (1) acidic reduction of Hg by Sn in this continuous-flow system requires breakdown of organomercurials prior to analysis; (2) tissue digestion using HNO3/H2SO4/HCl followed by the addition of K2Cr2O7 to stabilize Hg2+ achieves this breakdown and allows good recovery of total Hg; and (3) use ofl-cysteine to complex and prevent losses of Hg should be avoided in systems using acidic reduction of Hg. Concentrations of endogenous tissue sulfhydryls are generally lower than those associated with depressed absorbance signals during the acidic reduction of Hg.  相似文献   

20.
NH4+ inhibition kinetics for CH4 oxidation were examined at near-atmospheric CH4 concentrations in three upland forest soils. Whether NH4+-independent salt effects could be neutralized by adding nonammoniacal salts to control samples in lieu of deionized water was also investigated. Because the levels of exchangeable endogenous NH4+ were very low in the three soils, desorption of endogenous NH4+ was not a significant factor in this study. The Km(app) values for water-treated controls were 9.8, 22, and 57 nM for temperate pine, temperate hardwood, and birch taiga soils, respectively. At CH4 concentrations of ≤15 μl liter−1, oxidation followed first-order kinetics in the fine-textured taiga soil, whereas the coarse-textured temperate soils exhibited Michaelis-Menten kinetics. Compared to water controls, the Km(app) values in the temperate soils increased in the presence of NH4+ salts, whereas the Vmax(app) values decreased substantially, indicating that there was a mixture of competitive and noncompetitive inhibition mechanisms for whole NH4+ salts. Compared to the corresponding K+ salt controls, the Km(app) values for NH4+ salts increased substantially, whereas the Vmax(app) values remained virtually unchanged, indicating that NH4+ acted by competitive inhibition. Nonammoniacal salts caused inhibition to increase with increasing CH4 concentrations in all three soils. In the birch taiga soil, this trend occurred with both NH4+ and K+ salts, and the slope of the increase was not affected by the addition of NH4+. Hence, the increase in inhibition resulted from an NH4+-independent mechanism. These results show that NH4+ inhibition of atmospheric CH4 oxidation resulted from enzymatic substrate competition and that additional inhibition that was not competitive resulted from a general salt effect that was independent of NH4+.Atmospheric CH4 contributes substantially to the greenhouse effect, and the concentration of atmospheric CH4 has increased dramatically in the past century because of human activity associated with agriculture, land use changes, and industry (34, 35). Bacterial oxidation of atmospheric CH4 in well-drained soils is an important regulator of atmospheric CH4 concentration, yet the organisms responsible remain unidentified and the physiology of the process is poorly understood (9, 35, 36). Although soil CH4 consumption is inhibited by a wide variety of anthropogenic disturbances, such as agriculture, N deposition, and forestry (12, 17, 22, 23, 32, 43, 44), predictable inhibition patterns have failed to emerge, which has made it difficult to predict the effects of disturbance on soil CH4 flux in various ecosystems. The most commonly reported disturbance effect is that of NH4+ fertilizers, which can suppress soil CH4 consumption by up to 70% (1, 8, 10, 17, 22, 32, 33, 37, 38, 43). In the field, inhibition may occur immediately following fertilization, may be delayed for months to years, or may never occur despite years of chronic fertilization (9, 17). This variety of responses may stem at least in part from the distribution of physiologically diverse methane oxidizer populations across sites (17, 18, 20).Of the various NH4+ inhibition patterns, immediate inhibition is the best documented. As in field studies, however, physiological laboratory studies have produced variable results, suggesting that there may be multiple inhibition mechanisms (15, 17, 2628, 36, 39). Physicochemical similarities between CH4 and NH3 may permit these two compounds to compete for enzyme active sites so that fortuitous NH3 oxidation competitively inhibits CH4 oxidation (38). Although this mechanism has been demonstrated to occur in pure cultures of methanotrophic bacteria (6) and in a CH4-producing agricultural soil (15), it has not been demonstrated to occur in well-drained, nonagricultural mineral soils, which comprise the dominant terrestrial sink for atmospheric CH4 (14, 38, 45), nor has it been demonstrated to occur at near-atmospheric CH4 concentrations. In many cases, the kinetics of immediate NH4+ inhibition in soil cannot be reconciled easily with substrate competition (15, 16, 2628, 39). An alternative mechanism has been proposed, whereby the toxicity of NO2 or NH2OH produced by fortuitous NH4+ oxidation suppresses methanotrophic activity (26, 27, 39). Hence, multiple inhibition mechanisms may be involved, and these mechanisms may vary with the physiology of different CH4 oxidizer populations (17).Two physiologically distinct communities of CH4 oxidizers apparently exist in soil. One group, generally associated with atmospheric CH4 consumption, exhibits an extremely high affinity for CH4. Representatives of this group have yet to be cultivated or otherwise identified (9). The second group exhibits a much lower affinity for CH4 and is generally associated with common methanotrophs, such as those that have been studied in pure culture for many years (2, 9). In upland mineral soils, only high-affinity activity is usually detectable without artificial enrichment with high CH4 concentrations in the laboratory. However, the only prior study in which kinetic constants for NH4+ inhibition of soil CH4 oxidation were reported was conducted in a periodically moist, organic matter-rich agricultural soil with demonstrable methanogenesis (15, 16). Such a soil potentially harbors a rich community of CH4 oxidizers representing a continuum from low-affinity organisms to high-affinity organisms. Although this important investigation demonstrated that NH4+ inhibits CH4 oxidation via enzymatic substrate competition in an agricultural humisol, it is unclear to what extent its results apply to well-drained mineral soils lacking endogenous CH4 sources. Physiological studies of soil CH4 oxidation typically derive kinetic constants from oxidation rates at CH4 concentrations ranging from atmospheric levels (∼1.7 μl liter−1) to ≫Km for high-affinity CH4 oxidizers. Even in soil in which only high-affinity organisms are active, the CH4-oxidizing enzyme(s) could respond differently to NH4+ at high CH4 concentrations than at near-atmospheric concentrations (15, 39). Thus, to study NH4+ inhibition of high-affinity CH4 oxidizers per se, it would be preferable to examine inhibition kinetics at near-atmospheric CH4 concentrations in a soil with no apparent endogenous CH4 source.A common shortcoming of NH4+ inhibition studies, regardless of the organisms involved, has been a lack of attention to nonammoniacal salt effects despite numerous reports of substantial inhibition by such salts (1, 10, 15, 17, 24). King and Schnell (28) examined the effects of several Cl and SO42− salts and concluded that nonammoniacal salts indirectly inhibit CH4 oxidation by desorbing endogenous NH4+ from cation exchange sites in the soil, which then directly inhibit CH4 oxidation. Many N-limited soils, however, have extremely low concentrations of exchangeable NH4+, yet are substantially inhibited by nonammoniacal salts (17), suggesting that these salts have NH4+-independent effects on atmospheric CH4 oxidizers. Additional mechanisms may alter inhibition kinetics, thus hindering the diagnosis of NH4+-specific inhibition.With the limitations described above in mind, we used a simple steady-state kinetics approach to assess the mechanism of NH4+ inhibition of CH4 oxidation at near-atmospheric concentrations (1.8 to 15 μl liter−1) in three well-drained, N-limited forest soils that lack known endogenous CH4 sources. In addition, we examined the effects of nonammoniacal salts in parallel samples to judge the utility of these salts as experimental controls for neutralizing NH4+-independent salt effects.  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