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.
The hydration reaction of ethylene, C2H4+H2O → C2H5OH, catalyzed by oxoacids (H3PO4, H2SO4, and HClO4) and metal cations (B3+, Al3+, Sc3+, Ga3+, La3+, Be2+, Mg2+, Ca2+, Zn2+, and Sr2+) are studied systematically by density functional theory with a BLYP functional. The reaction profiles of the main reaction and some side reactions, such as ester formation, dimerization of ethylene, and dehydrogenation of ethanol, have been determined with a variety of catalysts. In each case, the intermediate states, the transition states, and their energetics are calculated. Metal cations react more efficiently for the main reaction than oxoacids, but they also make the dehydrogenation reaction active. While the dimerization reaction is strongly affected by the acidity of the catalyst, both the acidity and basicity of the catalyst are important for the dehydrogenation reaction. Efficient formation of ethanol from ethylene over a catalyst is suggested. © 2000 John Wiley & Sons, Inc. J Comput Chem 21: 1292–1304, 2000  相似文献   

2.
The reaction rate of the coulometric variant of the Karl-Fischer titration reaction (in which electrolytically generated triiodide is used as oxidant instead of iodine) has been measured in methanol. The reaction is first order in water, sulfur dioxide and triiodide, respectively. For pH<5 the reaction rate constant decreases logarithmically with decreasing pH. Addition of pyridine solely influences the pH (by fixing it to a value of about 6) and has no direct influence on the reaction rate. A linear relation exists between the reaction rate constant and the reciprocal value of the iodide concentration, from which we can calculate the individual reaction rates for the oxidation by iodine and triiodide, respectively. While the reaction rate constant for triiodide is relatively small (k3≈350 l2 mol?2s?1), the reaction rate constant for iodine is much larger (k3≈1.5×107 l2 mol?2 s?1.  相似文献   

3.
The dicarbollyliron (Cb2Fe-) redox reaction is studied at an amalgamated platinum electrode coated with a monolayer of behenic acid on top of which hemin is adsorbed. The redox reaction of Cb2Fe- involves two stages. First, an electrochemical reaction of adsorbed hemin proceeds, which involves the electron transfer through the dielectric monolayer; it is followed by the hemin’s chemical redox reaction with dissolved Cb2Fe-. It is shown that the adsorbed hemin transformation, no matter how small, is sufficient for the stimulation of the Cb2Fe- redox reaction.  相似文献   

4.
5.
The detailed kinetics of Cu(II) catalyzed reduction of toluidine blue (TB+) by phenyl hydrazine (Pz) in aqueous solution is studied. Toluidine white (TBH) and the diazonium ions are the main products of the reaction. The diazonium ion further decomposes to phenol (PhOH) and nitrogen. At low concentrations of acid, H+ ion autocatalyzes the uncatalyzed reaction and hampers the Cu(II) catalyzed reaction. At high concentrations, H+ hinders both the uncatalyzed and Cu(II) catalyzed reactions. Cu(II) catalyzed had stoichiometry similar to the uncatalyzed reaction, Pz+2 TB++H2O=PhOH+2 TBH+2 H++N2. Cu(II) catalyzed reaction occurs possibly through ternary complex formation between the unprotonated toluidine blue and phenyl hydrazine and catalyst. The rate coefficient for the Cu(II) catalyzed reaction is 2.1×104 M−2 s−1. A detailed 13‐step mechanistic scheme for the Cu(II) catalyzed reaction is proposed, which is supported by simulations. © 1999 John Wiley & Sons, Inc., Int J Chem Kinet 31: 271–276, 1999  相似文献   

6.
The dark reaction of NOx and H2O vapor in 1 atm of air was studied for the purpose of elucidating the recently discussed unknown radical source in smog chambers. Nitrous acid and nitric oxide were found to be formed by the reaction of NO2 and H2O in an evacuable and bakable smog chamber. No nitric acid was observed in the gas phase. The reaction is not stoichiometric and is thought to be a heterogeneous wall reaction. The reaction rate is first order with respect to NO2 and H2O, and the concentrations of HONO and NO initially increase linearly with time. The same reaction proceeds with a different rate constant in a quartz cell, and the reaction of NO2 and H218O gave H18ONO exclusively. Taking into consideration the heterogeneous reaction of NO2 and H2O, the upper limit of the rate constant of the third-order reaction NO + NO2 + H2O → 2HONO was deduced to be (3.0 ± 1.4) × 10?10 ppm?2-min?1, which is one order of magnitude smaller than the previously reported value. Nitrous acid formed by the heterogeneous dark reaction of NO2 and H2O should contribute significantly to both an initially present HONO and a continuous supply of OH radicals by photolysis in smog chamber experiments.  相似文献   

7.
With Li+ or TBA+ as counter ion the electrodimerization of phenylpent-2-en-3-one occurs by a 0.2 F initiation. This reaction is regio- and stereospecific with Li+, the radical autocatalytic reaction gave 100% yield of a new dimer. If TBA+ is used the primary step is an ion-substrate reaction giving two isomers in equal proportion. All reaction products were structurally determined by 2D-NMR (SECSY or COSY and NOESY sequence).  相似文献   

8.
This study combined electrochemical synthesis with traditional ferrite method to remove Co2+ from simulated liquid radioactive waste (LRW). The experiment investigated the effects of various reaction conditions including current density, reaction time for electrosynthesis, reaction temperature, initial pH value and boric acid concentration as well as the type of power supply by measuring the concentration of Co2+ in the effluent, explored the reaction mechanism by measuring particles using XRD. The results showed that it was feasible to remove Co2+ from simulated LRW by electrochemical synthesis of ferrite. The best removal efficiency of 99.99% (the concentration of Co2+ in the effluent was 0.447 μg/L) was achieved under the optimal reaction conditions, the sediment was mainly composed of the mixture of CoFe2O4 and Fe3O4.  相似文献   

9.
Chemiluminescence (CL) accompanying the reaction of U4+ with O2 in 0.0004–0.1M HClO4 was studied. It was found that the electron-excited uranyl ion (UO2 2+)* is the CL emitter. The fact that the reaction rate and the CL yield increase as the solution acidity decreases was explained by different reactivities of the U aq 4+ aquation and the products of its stepwise hydrolysis, UOH3+ and U(OH)2 2+, toward O2. Based on the results of analysis of the chain-radical mechanism of the reaction between U4+ and O2, it was concluded that transfer of an electron from the UO2 + ion to the oxidizing agent (a ·OH radical) is the most plausible elementary step of the reaction of (UO2 2+)* formation. It was found that the reaction rate, as well as the CL yield, increase substantially in the presence of uranyl ion. Catalytic action of UO2 2+ was explained by the formation of a UO2 2+·UO2 + complex, which reduces the rate of the UO2 + disproportionation reaction (UO2 + is an intermediate of the reaction and is involved in chain propagation), and by regeneration of the active center, UO2 +, in the reaction of UO2 2+ with U4+. Published inIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 1522–1528, September, 2000.  相似文献   

10.
The reaction of OH with acetylene was studied in a discharge flow system at room temperature. OH was generated by the reaction of atomic hydrogen with NO2 and was monitored throughout the reaction using ESR spectroscopy. Mass-spectrometric analysis of the reaction products yielded the following results: (1) less than 3 molecules of OH were consumed, and less than 2 molecules of H2O were formed for every molecule of acetylene that reacted; (2) CO was identified as the major carbon-containing product; (3) NO, formed in the generation of OH, reacted with a reaction intermediate to give among other products N2O. These observations placed severe limitations on the choice of a reaction mechanism. A mechanism containing the reaction OH + C2H2 → HC2O + H2 better accounted for the experimental results than one involving the abstraction reaction OH + C2H2 → C2H + H2O. The rate constant for the initial reaction was measured as 1.9 ± 0.6 × 10?13 cm3 molecule?1 sec?1.  相似文献   

11.
通过加入NaBH4作为诱导剂, 可在室温下引发肼与Co2+在水-乙醇体系中的还原反应, 制得高纯度纳米金属钴粉. 机理研究表明, 该反应分二段进行: 第一段主要发生Co2+被N2H4还原的反应(2Co2++N2H4+4OH=2Co¯+N2­+4H2O), 第二段主要为金属Co催化的肼分解反应(N2H4=N2­+2H2­)和歧化反应(3N2H4=N2­+4NH3­). Co2+被N2H4还原是典型的自催化过程, 因此, 加入少量NaBH4即可在288 K下启动反应. 通过测量气体产物的生成速率, 获得了Co2+还原的反应动力学方程, 发现Co2+, N2H4和产物Co的反应级数分别为1, 0和1, 反应活化能约为89 kJ/mol. 调节Co2+的浓度, 纳米金属钴的表面积可从11增加到25 m2/g.  相似文献   

12.
Aminofluorsilanes are obtained by the reaction of fluorosilanes with the lithium salts of the corresponding amines. The reaction of aminofluorosilanes with butyllithium in a (2 + 2)cyclo addition reaction leads to the formation of fourmembered silicon-nitrogen ring compounds. The mechanism of the reaction is discussed, the mass, 1H and 19F NMR spectra of the compounds are reported.  相似文献   

13.
The mechanism of ascorbic acid (DH2) oxidation with molecular oxygen catalysed by the polynuclear complex of Cu2+ with poly-4-vinylpyridine (PVP), partially quaternized by dimethylsulphate, has been studied. The half-conversion time of the reaction of DH2 with Cu(II) PVP under anaerobic conditions is independent of [Cu2+]. At pH 3.5, t0.5 (sec) = 0.8 + 5 × 10?4 [DH2]. The formation of an intermediate cupric-ascorbate complex is suggested (Kc ≈ 104 M?1). Free radicals of ascorbic acid are detected by the ESR-method combined with a flow technique. The small steady-state concentration of radicals indicates that their decay occurs inside the macromolecular complex. The rate constant of the PVP Cu(II) DH? ternary complex dissociation is ≈0.4 sec?1 (pH 3.5). The reaction of Cu(I) PVP with O2 is not accompanied by formation of O2? outside the macromolecule bulk. The rate constant of this reaction is 1.3 ± 0.15) × 102 M?1 sec?1 (pH 3.5). The cyclic mechanism of the catalytic reaction is suggested to include interchange of the redox state of copper-ions. About 23 of the total copper ion exists in the form Cu(I) PVP during the reaction at pH 3.5. The rate of DH2 oxidation under these conditions is limited by the rate of Cu(I) PVP reaction with O2. At pH 4.5 the overall reaction rate is limited by the rate of interaction of Cu(II) PVP with DH?.  相似文献   

14.
Ni-B超细非晶合金的化学制备、反应机理及性质研究   总被引:10,自引:0,他引:10  
本文研究了水溶液中NiCl2与KBH4的反应及其产物Ni-B超细非晶合金的性质。详细探讨了反应的电化学基础和反应过程机理。发现Ni^2+与BH^-4的反应独立的基本反应组成,其中不可避免地存在BH^-4的分解即与H2O的反应,Ni-B则产生于Ni^2+的还原  相似文献   

15.
The Berthelot reaction is a well-established colorimetric method for determination of ammonia. In this work, the effects of different bivalent ions (Ba2+, Cd2+, Co2+, Cu2+, Fe2+, Hg2+, Mg2+, Mn2+, Ni2+, and Pb2+) were studied as catalyst on the Berthelot reaction efficiency. CuCl2 was generally found as the best catalyst that provides a rapid and stable blue indophenol color. The Taguchi experimental design methodology has been applied to find optimum conditions. Four factors including temperature, pH, reaction time, and CuCl2 concentration at five levels were considered to achieve optimum conditions. Blue indophenol color stability for 40 min, and linearity response up to 20 mM of ammonium sulfate were achieved by further validation experiments. Limit of detection and quantification for this approach was 0.15 and 0.5 mM, respectively. Inhibitory activity of three traditional medicinal plants extract (Citrus aurantifolia, Laurus nobilis, and Zingiber officinale) was evaluated against jack bean urease activity by Berthelot reaction in the presence of CuCl2 as catalyst, and results were compared with traditional Berthelot reaction.  相似文献   

16.
Spectrophotometric methods were used to investigate the rate of the reaction of Br2 with HCOOH in aqueous, acidic media. The reaction products are Br? and CO2. The kinetics of this reaction are complicated by both the formation of Br3? as Br? is formed and the dissociation of HCOOH into HCOO? and H+. Previous work on this reaction was carried out at acidities lower than the highest used here and led to the conclusion that only HCOO? reacts with Br2. It is agreed that this is by far the principal reaction. However, at the highest acidity experiments, an added small component of reaction was found, and it is suggested that it results from the direct reaction of Br2 with HCOOH itself. On this assumption, values of the rate constants for both reactions are derived here. The rate constant for the reaction of HCOO? with Br2 agrees with values previously reported, within a factor of 2 on the low side. The reaction involving HCOOH is more than 2000 times slower than the reaction involving HCOO?, but it does contribute to the overall rate as [H+] approaches 1M. These derived rate constants are able to simulate quantitatively the authors' absorbance-versus-time data, demonstrating the validity of their data treatment methods, if not mechanistic assignments. Finally, activation parameters were determined for both rate constants. The values obtained are: ΔE?(HCOOH + Br2) = 13.3 ± 1.1 kcal/mol, ΔS? (HCOOH + Br2) = ?28 ± 3 cal/deg mol, ΔE? (HCOO? + Br2) = 13.1 ± 0.9 kcal/mol, and ΔS?(HCOO? + Br2) = ?12 ± 1 cal/deg mol. That the activation energies of the two reactions turn out to be essentially identical does not support the authors' suggestion that both HCOOH and HCOO? react with Br2.  相似文献   

17.
Degradation of methyl tert-butyl ether (MTBE) with Fe2+/H2O2 was studied by purge-and-trap gas chromatography-mass spectrometry. MTBE was degraded 99% within 120 min under optimum conditions. MTBE was firstly degraded rapidly based on a Fe2+/H2O2 reaction and then relatively slower based on a Fe3+/H2O2 reaction. The dissolved oxygen decreased rapidly in the Fe2+/H2O2 reaction stage, but showed a slow increase in the Fe3+/H2O2 reaction stage. tert-Butyl formate, tert-butyl alcohol, methyl acetate and acetone were identified as primary degradation products by mass spectrometry. A preliminary reaction mechanism involving two different pathways for the degradation of MTBE with Fe2+/H2O2 was proposed. This study suggests that degradation of MTBE can be achieved using the Fe2+/H2O2 process.  相似文献   

18.
Reactions of HCCCO and NCCO radicals with O2 have been studied by a combination of pulsed laser photolysis and photoionization mass spectrometry. HCCCO was produced by 193‐nm photolysis of methylpropiolate or 3‐butyn‐2‐one, and NCCO was formed by 193‐nm photolysis of acetylcyanide. The rate constants obtained at 298 ± 3 K were (6.5 ± 0.7) × 10?12 cm3 molecule?1 s?1 for the HCCCO + O2 reaction, and no pressure dependence was observed between 1.5 and 16 Torr of N2 as a bath gas. Because HCO and HCCO radicals were observed as reaction products, it was confirmed that the reaction proceeds by a two‐body reaction. On the other hand, the rate constants of NCCO with O2 depended on the total pressure and were (5.4–8.8) × 10?13 cm3 molecule?1 s?1 for total pressures 2.0–15.5 Torr of N2, confirming that the reaction proceeds by a three‐body process. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 440–448, 2001  相似文献   

19.
The potential energy surfaces for the La+SCO and La++ SCO reactions have been theoretically investigated by using the DFT (B3LYP/ECP/6-311+G(2d)) level of theory. Both ground and excited state potential energy surfaces (PES) are discussed. The present results show that the reaction mechanism is insertion mechanism both along the C-S and C-O bond activation branches, but the C-S bond activation is much more favorable in energy than the C-O bond activation. The reaction of La atom with SCO was shown to occur preferentially on the ground state (doublet) PES throughout the reaction process, and the experimentally observed species, have been explained according to the mechanisms revealed in this work. While for the reaction between La+ cation with SCO, it involves potential energy curve-crossing which dramatically affects reaction mechanism, and the crossing points (CPs) have been localized by the approach suggested by Yoshizawa et al. Due to the intersystem crossing existing in the reaction process of La+ with SCO, the products SLa+2CO) and OLa+2CS) may not form. This mechanism is different from that of La + SCO system. All our theoretical results not only support the existing conclusions inferred from early experiment, but also complement the pathway and mechanism for this reaction.  相似文献   

20.
The dynamics of the H(2S) + FO(2Π) → OH(2Π) + F(2P) reaction on the adiabatic potential energy surface of the 13A′ and 13A″ states is investigated. The initial state selected reaction probabilities for total angular momentum J = 0 have been calculated by using the quantum mechanical real wave packet method. The integral cross sections and initial state selected reaction rate constants have been obtained from the corresponding J = 0 reaction probabilities by means of the simple J‐Shifting technique. The initial state‐selected reaction probabilities and reaction cross section do not manifest any sharp oscillations and the initial state selected reaction rate constants are sensitive to the temperature. © 2010 Wiley Periodicals, Inc. J Comput Chem 2010  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