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 27 毫秒
1.
The chain mechanism of oxidation of hydrazine in aqueous solutions saturated with oxygen at pH > 7.5 was established. The yields of radiochemical decomposition of hydrazine increase with a decrease in the absorbed dose rate, an increase in the concentration of hydrazine, and the pH of the solution and attains hundreds of molecules per 100 eV of absorbed energy. It is hypothesized that the chain process includes a stage of formation of the N2H3 radical, its reaction with O2, and the formation of O2 or N2H3O2. The chain propagation reaction is due to the reaction of molecules of N2H4 with N2H3O2 or O2 .Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 11, pp. 2450–2453, November, 1989.  相似文献   

2.
A mathematical model is proposed for the process of conversion of a highly concentrated gas into an aerosol over the surface of a solid reagent. On the basis of this model, the process of aerosol formation by the interaction between ammonia and a crystal hydrate of ferric nitrate in a flow-type reactor is studied. Recommendations for choosing the operating conditions required for ensuring proportional concentrations of the aerosol and reacting gas at high latter concentrations (1.2 × 10–6kmol m–3and higher) are given. The results obtained are compared with experimental data.  相似文献   

3.
The progress in the field of methane conversion into higher hydrocarbons including aromatics and oxygenated compounds in the recent five years will be reviewed shortly, together with a new type of the methane conversion reaction with carbon monoxide at lower temperatures (600–700 K) by supported group VIII metal catalysts. Benzene was formed selectively among hydrocarbons in the CH4–CO reaction over silica-supported Rh, Ru, Pd and Os catalysts under atmospheric pressure. Both CH4 and CO were required for benzene formation, and only ethane and ethylene were formed besides benzene. The amount of C3–C5 hydrocarbons was negligible, which suggests that a completely different mechanism from the CO–H2 reaction may be operating over these catalysts despite of the similarity in the reaction conditions with the CO–H2 reaction. The mechanism of benzene formation was studied deeply by means of kinetical investigation as well as infrared spectroscopy and isotopic tracer method in connection with that of CO hydrogenation.  相似文献   

4.
Experimental data on the influence of oxygen on the rate of formation of N2, N2O, and CO2 were obtained for a wide range of conditions, the goal being to substantiate and develop a previously proposed mechanism and reaction kinetics for the reaction of CO with NO and O2. The effect of the reversibility of the NO adsorption step on the kinetics of the process was also analyzed. The conditions necessary for acceleration of the reaction of CO + NO by oxygen were obtained. Good correspondence was also obtained between the calculated and experimental kinetic dependences.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 25, No. 3, pp. 289–294, May–June, 1989.  相似文献   

5.
Characteristics of the -induced chain reaction between sulfur dioxide and molecular oxygen in perchloric and sulfuric acid media in the presence of Ce(III) ions have been studied. The concentration effects of dissolved oxygen (0.2·10–3–9.4·10–3 mol/dm3, sulfur dioxide (0.3·10–1–2.0·10–1 mol/dm3 and Ce(III) (0.2·10–3–4.8·10–3 mol/dm3) and dose rate (0.26·1019–1.0·1019 eV/dm3·s) on the radiation — chemical yield of oxygen consumption G(–O2) and accumulation of sulfate G(HSO 4 ), have been investigated. The reaction proceeds with G(–O2) reaching 102–103 molecule/100eV in a catalytic regime. The reaction rate in perchloric acid medium is 3–4 times lower than that in the sulfuric acid medium and depends on the SO2, O2 and Ce(III) concentrations, the reaction order varying from 1.0 to 0 and/or in the reverse direction. The mechanism of the process involves chain propagation with 3 stages and 3 intermediates: SO3H, HSO5 and Ce(IV). The catalytic effect is caused by the interaction of HSO4 with Ce(IV) ions followed by their reduction when interacting with SO2, yielding SO3H radicals. Chain termination may be due to one or two of the three intermediates or due to all three particles, the kinetics depending on this. Kinetic equations describing the experimental data have been obtained.  相似文献   

6.
Oxidative conversion of a mixture of natural gas and oxygen in a barrier-discharge plasma-chemical reaction was investigated experimentally. The process was conducted at atmospheric pressure and room temperature. The discharge was initiated by high-voltage pulses of 50–100 s duration at a repetition frequency of up to 2 kHz. The principal feature of the process was that in the plasma-chemical reactor conditions were created which stimulated the condensation of the products of incomplete oxidation of methane that resulted in the formation of aerosol even from nonsaturated vapor. The removal of intermediate reagents from the gaseous phases into the aerosol prevented them from further oxidation. Depending on the experimental conditions, the mass percentage of the components of the condensate formed varied within the following limits: formic acid from 20 to 40%, methanol from 8 to 15%, methylformate from 4 to 8%, and water from 40 to 60%. The conversion process has been realized on a laboratory setup of average power up to 1 kW. In the single-pass mode, a 57% degree of conversion of the mixture has been achieved. The energy value of the condensate is 15–20 kWh/kg.  相似文献   

7.
Along with hydrodechlorination, the formation of C1 and higher hydrocarbons takes place in a flow system in the presence of catalysts containing 0.5–5.0% Pd supported on a Sibunit carbon carrier at 150–230°C. In the entire range of conditions examined, the reaction products are primarily methane, C2–C4 hydrocarbon fractions, and C5 traces. The catalysts are stable in operation, and a high conversion of CCl4 was retained for a long time interval. The nonselective formation of linear and branched hydrocarbons is indicative of a radical mechanism of the process.  相似文献   

8.
Summary In laboratory-scale experiments sea salt particles are exposed to SO2 at a temperature of 22°C and relative humidities of 40, 60 and 80%; the SO2 gas concentration is fixed to 0.2, 0.5 and 1.0 ppm (v), respectively. In further test series NO2 is added to the gas phase. As kinetic data the capacity values of the sea salt particles (mg formed sulfate/g dry aerosol) are determined as function of time and from this the reaction rates (mg formed sulfate/g dry aerosol and minute) are calculated in dependence of the yield. The relative humidity (r.h.) has proved to be a decisive reaction parameter. For example, the rate (at a reaction time of one hour) increases at a SO2 concentration of 0.5 ppm (v) from 0.01 to approx. 0.1 mg SO 4 2– /g·min, if the r.h. will increase from 40 to 80%. However, the gas concentration has only an importance at high humidities (where the reaction takes place in droplets) for the sulfate formation in sea salt aerosols. If the SO2 concentration is reduced from 1.0 to 0.2 ppm (v) at a r.h. of 80%, the rate will be decreased from 0.2 to about 0.07 mg SO 4 2– /g·min; however, at a r.h. of 60% from 0.075 to 0.04 mg SO 4 2– /g·min. As an increased sulfate formation but no nitrate formation can be detected when NO2 is added to the gas phase, it can be assumed that SO2 is oxidized in the electrolyte layer around the sea salt particles whereas NO2 is reduced. If NO2 (SO2:NO2=1:1) is added to the gas phase, the rate — for example at a r.h. of 40% — will be increased from 0.01 to 0.24 mg SO 4 2– /g·min.  相似文献   

9.
Summary The kinetics and mechanism of the system: [FeL(OH)]2–n + 5 CN [Fe(CN)5(OH)]3– + Ln–, where L=DTPA or HEDTA, have been investigated at pH= 10.5±0.2, I=0.25 M and t=25±0.1 C.As in the reaction of [FeEDTA(OH)]2–, the formation of [Fe(CN)5(OH)]3– through the formation of mixed ligand complex intermediates of the type [FeL(OH)(CN)x]2–n–x, is proposed. The reactions were found to consist of three observable stages. The first involves the formation of [Fe(CN)5(OH)]3–, the second is the conversion of [Fe(CN)5(OH)]3– into [Fe(CN)6]3– and the third is the reduction of [Fe(CN)6]3– to [Fe(CN)6]4– by oxidation of Ln– The first reaction exhibits a variable order dependence on the concentration of cyanide, ranging from one at high cyanide concentration to three at low concentration. The transition between [FeL(OH)]2–n and [Fe(CN)5(OH)]3– is kinetically controlled by the presence of four cyanide ions around the central iron atom in the rate determining step. The second reaction shows first order dependence on the concentration of [Fe(CN)5(OH)]3– as well as on cyanide, while the third reaction follows overall second order kinetics; first order each in [Fe(CN)6]3– and Ln–, released in the reaction. The reaction rate is highly dependent on hydroxide ion concentration.The reverse reaction between [Fe(CN)5(OH)]3– and Ln– showed an inverse first order dependence on cyanide concentration along with first order dependence each on [Fe(CN)5– (OH)]3– and Ln–. A five step mechanism is proposed for the first stage of the above two systems.  相似文献   

10.
Over the past few decades, numerous model systems have been discovered that create carbon–carbon bonds from CO. These reactions are of potential relevance to the Fischer–Tropsch process, a technology that converts syngas (H2/CO) into mixtures of hydrocarbons. In this paper, a homogeneous model system that constructs carbon chains from CO is reported. The system exploits the cooperative effect of a transition metal complex and main group reductant. An entire reaction sequence from C1 → C2 → C3 → C4 has been synthetically verified. The scope of reactivity is broad and includes a variety of transition metals (M = Cr, Mo, W, Mn, Re, Co), including those found in industrial heterogeneous Fischer–Tropsch catalysts. Variation of the transition metal fragment impacts the relative rate of the steps of chain growth, allowing isolation and structural characterisation of a rare C2 intermediate. The selectivity of carbon chain growth is also impacted by this variable; two distinct isomers of the C3 carbon chain were observed to form in different ratios with different transition metal reagents. Based on a combination of experiments (isotope labelling studies, study of intermediates) and calculations (DFT, NBO, ETS-NOCV) we propose a complete mechanism for chain growth that involves defined reactivity at both transition metal and main group centres.

A homogeneous model system that constructs carbon chains from CO is reported. The system exploits the cooperative effect of a transition metal complex and main group reductant. An entire reaction sequence from C1 → C2 → C3 → C4 has been synthetically verified.  相似文献   

11.
Oxidation of secondary alcohols (2-propanol, 2-butanol, and cyclohexanol) by dioxygen, catalyzed by CuI ando-phenanthroline complexes, in the presence of alkali, was studied. The conditions under which oxidative dehydrogenation of secondary alcohols result in fast formation of ketones as the only primary oxidation products were found. Bis-phenanthrolinates [Cu(phen)2]+ are the active forms of the catalyst. The catalytic turnover number for complexes between copper(i) ando-phenanthroline is 1 to 2 s–1 at room temperature.Kinetic regularities of the reaction are similar to those of the oxidation of alcohols in the presence of oxidases. The mechanism of the process is proposed, suggesting that the oxidation of secondary alcohols occursvia a concerted two-electron mechanism involving a stage of formation of the ternary complex [O2...Cu(phen)2 +...OCHR1R2]. It is significant for the oxidation mechanism that a hydrogen atom is transferred from the anionic form of a substrate to oxygen, which is confirmed by the value of the kinetic isotope effectk H/k D = 2.1.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 1952–1958, October, 1995.The work was financially supported by the Russian Foundation for Basic Research (Project No. 94-03-08733a) and the International Science Foundation (Grant MN4 000).  相似文献   

12.
The formation of 2-aminoacetamide from ammonia and glycine and N-glycylglycine from two glycine molecules with and without Mg2+, Cu2+, and Zn2+ cations as catalysts have been studied as model reactions for peptide bond formation using the B3LYP functional with 6–311+G(d,p) and 6–31G(d) basis sets. The B3LYP method was also used to characterize the nine gas–phase complexes of neutral glycine, its amide (2-aminoacetamide), and N-glycylglycine with Lewis acids Mg2+, Cu2+, and Zn2+, respectively. Further, the gas-phase hydration of metal-coordinated complexes of glycine, 2-aminoacetamide, and N-glycylglycine was also investigated. Finally, the effect of water on the structure and reactivity of the metal coordinated complexes was determined. Enthalpies and Gibbs energies for the stationary points of each reaction have been calculated to determine the thermodynamics of the reactions investigated. A substantial decrease in reaction enthalpies and Gibbs energies was found for glycine–ammonia and glycine–glycine reactions coordinated by Mg2+, Cu2+, and Zn2+ ions compared to those of the uncoordinated 2-aminoacetamide bond formation. The formation of a dipeptide is a more exothermic process than the creation of simple 2-aminoacetamide from glycine. The energetic effect of the transition metal ions Cu2+ and Zn2+ is of similar strength and more pronounced than that of the Mg2+ cation. The basicity order of the amides investigated shows the order: NH2CH2CO2H < NH2CH2CONH2 < NH2CH2CONHCH2CO2H. Interaction enthalpies and Gibbs energies of metal ion–amide complexes increase as Mg2+2+2+. In both reactant (glycine) and reaction products (2-aminoacetamide, N-glycylglycine) dihydration caused considerable reduction (about 200–500 kJ-mol–1) of the strength of the bifurcated metal–amide bonds. Solvent effects also reduce the reaction enthalpy and Gibbs energy of reactions under study.  相似文献   

13.
The methods of optical, ESR, and IR spectroscopy were used to obtain data on the structure and mechanism for the formation of the products in the reaction of dioxasilirane groups (Si–O)2Si 2 (DOSG) stabilized on the silica surface. Depending on the regime of the reaction (temperature and methane pressure), the process is accompanied by the formation of various products: methoxy (–O–CH3) and ethoxy (–O–C2H5) groups. The process mechanism is elucidated: this is a free-radical reaction in which paramagnetic sites are generated in the reaction between DOSG and methane molecules. The formation of final products is due to the reactions >Si(O)(OCH3) + CH4 >Si(OH)(OCH3) + CH3 and >Si(O–CH2)(OH) + CH3 >Si(OH)(OC2H5). The ratio of the rate constants of methyl radical addition to (Si–O)2Si: and (Si–O)2Si 2 at room temperature was determined experimentally (4.6 ± 1.0).  相似文献   

14.
The yields of formation of radiolytic hydrogen (H2) and orthopositronium (o-Ps) in aqueous and alcoholic acetone solutions were experimentally determined. A decrease in the o-Ps yield with an increase in the acetone concentration is much weaker than the decline in the yield of solvated electrons (e s) under picosecond pulse radiolysis conditions. In contrast, the decrease in the o-Ps yield is minimal in higher alcohols where the inhibiting action of acetone e s is most pronounced. These findings seem to contradict the conventional concepts of Ps formation via the intratrack reaction of positron recombination with a track electron (e), which competes with the reaction of e scavenging by dissolved acetone molecules. This contradiction can be eliminated, assuming that the scavenging of e by acetone begins from the formation of the weakly bound transient state (CH3)2CO···e capable of donating e to a positron. This opens up an additional pathway for the formation of the Ps atom.  相似文献   

15.
The partial oxidation of methane to syngas is studied in the presence of Pt- and Ni-containing catalysts. The process kinetics does not provide unequivocal information on the order of formation of products (including carbon oxides) when either methane–oxygen or methane–oxygen–CO2 mixtures are used. Experiments with 13C-labeled carbon dioxide added show the difference in the behavior of the catalysts. In the presence of Pt/ZrO2, there is no noticeable transfer of the isotopic label to the CO molecules. On the nickel catalyst, 13CO is formed in substantial amounts, which can probably be explained by the redox reaction of 13CO2 with metallic nickel under oxygen-free conditions behind the zone of the main reaction of methane oxidation.  相似文献   

16.
It is established that the reaction of recharging trioxalate complexes of ruthenium(III) occurs in the case of solutions with excess supporting oxalate salts of alkali metals K+, Na+, and Cs+ in reversible conditions, and limiting recharge currents are caused by diffusion. At the same time, values of diffusion coefficients for complex anion [Ru(C2O4)3]–3 decrease by almost two times upon going from potassium to sodium and cesium electrolytes. Substantial differences in the limiting currents in solutions containing excess amounts of the above salts are explained by the formation, at least in the case of cesium and sodium electrolytes, of ionic associates whose reduction rate at a fixed potential is lower than that of nonassociated anion [Ru(C2O4)3]–3. With solution dilution by supporting salts, transition is observed from reversible recharge conditions to absolutely irreversible conditions and a change in the above sequence of the effect of supporting cations on the recharge rate; at a fixed potential, the process decelerates in the series Cs+ > K+ > Na+. The reduction wave of the ruthenium(II) oxalate complexes in solutions with excess supporting electrolyte happens to depend on pH and, probably, is determined by simultaneous formation of adsorbed atoms of hydrogen (or ruthenium hydride) on atoms of ruthenium(0).  相似文献   

17.
Summary A kinetic study is reported on the substitutions: FoPdX+YFoPdY+X (H2Fo=l-(2-hydroxyphenyl)-3,5-diphenylformazan) with X=NH3 and py and Y = thiourea, tetramethylthiourea, triphenylphosphine and the thiocyanate ion, in seven nonaqueous pure solvents. Under pseudo first-order conditions a two-term rate-law is obeyed: k(obs)= k1+k2 [Y]. The results in terms of reactivity pattern and solvent effects on initial and transition states are very similar to the ones found for the analogous substitutions at platinum(II). The rate of solvolysis (k1) is determined by the donor number of the solvent and is not related to the transfer free energy of solvation of the substrate. Nonspecific solvation effects dominate. The entropy of activation for the direct nucleophilic displacement, FoPdNH3+Ph3P, is found to lie between –110 and –20 JK–1 mol–1, indicating the associative character of the substitutions. The named reaction exhibits an isokinetic relationship in the various solvents. In spite of that, an initial state — transition state — final state comparison shows the position of the transition state on the reaction coordinate to be solvent dependent. The importance of charge transfer from the donor solvent to the metal ion in determining the Gibbs free energy of the transition state is emphasized.  相似文献   

18.
The reaction of fluorobenzene with urotropine in trifluoroacetic acid (TFAA) at high pressures and temperatures affords predominantly fluorobenzaldehydes andN-(fluorophenylmethyl)trifluoroacetamides. The yields of these products depend considerably on the reaction conditions. The rates of their formation have the maximum values at the momemt of the phase transition (PT) of TFAA. A new efficient cyclic (dynamic) regime is proposed for the synthesis at high pressures. The regime involves periodically occurring PT of the solvent. The change in the relative rate of product formation with the degree of fluorobenzene conversion is wave-like.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 319–323, February, 1995.  相似文献   

19.
A process to prepare microparticles of narrow size distribution having a particle size in the range of approximately 1–8 μm was developed. The primary objective of this work was to study the formation and morphology of copolyester microparticles prepared using a sulfonated copolyester emulsion by an emulsion–aggregation process. Molecular weight of the copolyesters was measured by gel permeation chromatography. The glass transition temperature (Tg) of the copolyesters was found to be in the range of 40–70 °C. Aggregating agents used in this study were 1–5% (wt.%) solutions of divalent ions of zinc acetate and magnesium chloride salts. Emulsion–aggregation experiments were performed at various temperatures: 40, 50, 60, and 80 °C. Particle morphologies studied by field emission-scanning electron microscopy measurements provided an understanding of the conditions and mechanism leading to formation of microparticles by the emulsion–aggregation process. Molecular weight and Tg of the copolyester, the concentration of aggregating agent, and the temperature were determined to be the most important parameters influencing the preparation of microparticles. This process illustrates the preparation of microparticles of uniform size with morphology of controlled shape from a nanometer-sized emulsion by ionic crosslinking.  相似文献   

20.
Based on the Belousov-Zhabotinskii reaction in CSFR we observe that the mixing rate decrease in the reaction volume of the mixture to the value 0 leads to a qualitative change in the nature of the chaotic regime, similar to a second order phase transition. We show that the largest Lyapunov exponent for the system is directly proportional to 0 2.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 28, No. 1, pp. 52–56, January–February, 1992.  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