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 662 毫秒
1.
A new potential energy surface (PES) for the atmospheric formation of sulfuric acid from OH+SO2 is investigated using density functional theory and high-level ab initio molecular orbital theory. A pathway focused on the new PES assumes the reaction to take place between the radical complex SO3·HO2 and H2O. The unusual stability of SO3·HO2 is the principal basis of the new pathway, which has the same final outcome as the current reaction mechanism in the literature but it avoids the production and complete release of SO3. The entire reaction pathway is composed of three consecutive elementary steps:(1) HOSO2+O2→SO3·HO2, (2) SO3·HO2+H2O→SO3·H2O·HO2, (3) SO3·H2O·HO2→H2SO4+HO2. All three steps have small energy barriers, under 10 kcal/mol, and are exothermic, and the new pathway is therefore favorable both kinetically and thermodynamically. As a key step of the reactions, step (3), HO2 serves as a bridge molecule for low-barrier hydrogen transfer in the hydrolysis of SO3. Two significant atmospheric implications are expected from the present study. First, SO3 is not released from the oxidation of SO2 by OH radical in the atmosphere. Second, the conversion of SO2 into sulfuric acid is weakly dependent on the humidity of air.  相似文献   

2.
The gas‐phase reaction of organic acids with SO3 has been recognized as essential in promoting aerosol‐particle formation. However, at the air–water interface, this reaction is much less understood. We performed systematic Born–Oppenheimer molecular dynamics (BOMD) simulations to study the reaction of various organic acids with SO3 on a water droplet. The results show that with the involvement of interfacial water molecules, organic acids can react with SO3 and form the ion pair of sulfuric‐carboxylic anhydride and hydronium. This mechanism is in contrast to the gas‐phase reaction mechanisms in which the organic acid either serves as a catalyst for the reaction between SO3 and H2O or reacts with SO3 directly. The distinct reaction at the water surface has important atmospheric implications, for example, promoting water condensation, uptaking atmospheric condesation species, and incorporating “SO42?” into organic species in aerosol particles. Therefore, this reaction, typically occurring within a few picoseconds, provides another pathway towards aerosol formation.  相似文献   

3.
4.
The present paper describes the radiolytic oxidation of 10-4M K4[Fe(CN)6] aqueous aerated solution in the presence of different concentrations of SO2. In the absence of SO2, ferrocyanide is oxidized to ferricyanide with a G value of 7.2. In this solution two O2- react to form H2O2. Ferrocyanide is oxidized by OH and H2O2 both. At low concentration of SO2, O2- reacts with SO2 forming first SO4._, which leads to chain oxidation of SO2. The G [Fe(CN)63-) decrease from 7.2 to 4.5. At higher concentrations of SO2, H2O2 also reacts with SO2 and the G(Fe(CN)63-] further reduces to 2.7. In the presence of chloride ions, SO4._ converts them to chloride atoms which react with H2O2 and ferrocyanide and the G[Fe(CN)63-] is again increased to 4.3. The reaction of OH with SO2 was observed only at high concentrations of SO2 since the reaction of OH with ferrocyanide is very fast. The importance of water and ammonia in the conversion of sulphur dioxide to sulphuric acid probably lies in their reaction with SO42- to form H2SO4 or (NH4)2SO4.  相似文献   

5.
Preparation and Thermal Properties of Copper(I) Sulfate Cu2SO4 Copper(I) sulfate Cu2SO4 can be prepared in high purity by reaction of Cu2O with dimethyl sulfate (CH3)2SO4 at 160°C in an argon atmosphere. Using an extremely fine grained Cu2O, as obtained by reduction of cupric acetate with hydrazine, and a reaction time of 10 minutes a Cu2SO4 is obtained that contains less than 1% Cu2O. Longer reaction times lead to partial decomposition of the Cu2SO4 to Cu(met.) and CuSO4. In a closed system Cu2SO4 melts at about 400°C, however, the melt rapidly decomposes to Cu and CuSO4, solidifying simultaneously. When heated in a thermoanalyzer in flowing argon or in a vacuum, Cu and CuSO4 react under liberation of SO2. Increasing the temperature leads to CuO in three steps, which converts to Cu2O when heated to 1000°C. The question of formation of Cu2SO4, occasionally mentioned in the literature, being responsible for the liquid phases observed in the system Cu? S? O at temperatures below 500°C, is discussed.  相似文献   

6.
The investigations of reaction between Ag2SO4 and Ag2S in air atmosphere have been carried. Results of DTA and X-ray phase powder diffraction of a reaction mixture have confirmed that in the Ag?O?S system exists a new phase. A formula of the phase is Ag2SO2.  相似文献   

7.
The reaction of CuSO4 with Cu2SO2 to give Cu2SO4 was studied. The influence of the degree of reaction, the initial mixture composition and the temperature upon the reaction rate and the product composition was discussed. It was found that the reaction starts above 710 K and pure Cu2SO4 can be obtained under strictly defined conditions.  相似文献   

8.
We report a series of calculations to elucidate one possible mechanism of SO2 chemisorption in amino acid-based ionic liquids. Such systems have been successfully exploited as CO2 absorbents and, since SO2 is also a by-product of fossil fuels’ combustion, their ability in capturing SO2 has been assessed by recent experiments. This work is exclusively focused on evaluating the efficiency of the chemical trapping of SO2 by analyzing its reaction with the amino group of the amino acid. We have found that, overall, SO2 is less reactive than CO2, and that the specific amino acid side chain (either acid or basic) does not play a relevant role. We noticed that bimolecular absorption processes are quite unlikely to take place, a notable difference with CO2. The barriers along the reaction paths are found to be non-negligible, around 7–11 kcal/mol, and the thermodynamic of the reaction appears, from our models, unfavorable.  相似文献   

9.
Abstract— The gas phase photochemical reactions of SO2 induced by 3130 Å radiation have been studied in the presence of added alkanes or added CO. The quantum yields obtained in the reactions with the low molecular weight alkanes employed are lower than those obtained by previous workers. The quantum yields were found to be pressure dependent increasing slowly with increasing pressure. A stoichiometric ratio of one SO2 removed per molecule of hydrocarbon consumed was observed only under experimental conditions of [SO2] < [RH]. For reaction mixtures where [SO2] < [RH] the ratio of [SO2]/[RH] reacted always exceeded unity. The quantum yields decreased slightly with increasing temperature. In all the alkane reaction systems studied, the deposition of viscous, nonvolatile reaction products was observed. In the experiments with added CO, the quantum yields were computed with respect to the rate of CO2 formation. At 25°C and equal pressures of SO2 and CO, φco2 was observed to be 0.005 and it decreased slightly with increasing temperature. The results obtained are interpreted in terms of the sulfoxidation of the alkanes and the oxidation of CO proceeding by way of a 3SO2 reaction intermediate.  相似文献   

10.
Long B  Long ZW  Wang YB  Tan XF  Han YH  Long CY  Qin SJ  Zhang WJ 《Chemphyschem》2012,13(1):323-329
The formic acid catalyzed gas‐phase reaction between H2O and SO3 and its reverse reaction are respectively investigated by means of quantum chemical calculations at the CCSD(T)//B3LYP/cc‐pv(T+d)z and CCSD(T)//MP2/aug‐cc‐pv(T+d)z levels of theory. Remarkably, the activation energy relative to the reactants for the reaction of H2O with SO3 is lowered through formic acid catalysis from 15.97 kcal mol?1 to ?15.12 and ?14.83 kcal mol?1 for the formed H2O ??? SO3 complex plus HCOOH and the formed H2O ??? HCOOH complex plus SO3, respectively, at the CCSD(T)//MP2/aug‐cc‐pv(T+d)z level. For the reverse reaction, the energy barrier for decomposition of sulfuric acid is reduced to ?3.07 kcal mol?1 from 35.82 kcal mol?1 with the aid of formic acid. The results show that formic acid plays a strong catalytic role in facilitating the formation and decomposition of sulfuric acid. The rate constant of the SO3+H2O reaction with formic acid is 105 times greater than that of the corresponding reaction with water dimer. The calculated rate constant for the HCOOH+H2SO4 reaction is about 10?13 cm3 molecule?1 s?1 in the temperature range 200–280 K. The results of the present investigation show that formic acid plays a crucial role in the cycle between SO3 and H2SO4 in atmospheric chemistry.  相似文献   

11.
The extensive bands observed from the helium afterglow reaction of SO2 in the 250–540 nm region are assigned to the new SO+(A2Π-X2Πr) system produced from the He+/SO2 dissociative charge-transfer reaction at thermal energy. They had been erroneously interpreted as the SO+2 (C?-X?) system produced from He(23S)/SO2 Penning ionization. The spectroscopic constants for the SO+A2Π) and SO+(X2Πr) states were determined.  相似文献   

12.
The rate of adsorption of SO2 on a prototype carbonaceous surface was measured at low pressure in a flow reactor. The measured rate indicates a maximum atmospheric loss of SO2 by heterogeneous reaction of 1%/h for a particle density of 100 μg/m3. The capacity of carbon particles to adsorb SO2 is limited at ~1 mg SO2 g?1 C. NO2 has no effect on the rate of SO2 adsorption or the saturation behavior.  相似文献   

13.
The effects of SO2, V2O5 loading and reaction temperature on the activity of activated carbon supported vanadium oxide catalyst have been studied for the reduction of NO with NH3 at low temperatures (150—250°C). It is found that SO2 significantly promotes the catalyst activity. Both V2O5 loading and reaction temperature are vital to the promoting effect of SO2. The catalysts with V2O5 loadings of 1—5 weight percent have a positive effect on the promotion of SO2, while the catalysts with V2O5 loadings of above 7 weight percent have not such an effect or show a negative effect. At lower temperatures (<180°C) SO2 poisons the catalyst but at higher temperatures promotes it. The reason of the SO2 promotion was also discussed; it may results from the formation of SO4 2? on the catalyst surface, which increases the surface acidity and hence the catalytic activity.  相似文献   

14.
Flow reactor experiments were performed to study moist CO oxidation in the presence of trace quantities of NO (0–400 ppm) and SO2 (0–1300 ppm) at pressures and temperatures ranging from 0.5–10.0 atm and 950–1040 K, respectively. Reaction profile measurements of CO, CO2, O2, NO, NO2, SO2, and temperature were used to further develop and validate a detailed chemical kinetic reaction mechanism in a manner consistent with previous studies of the CO/H2/O2/NOX and CO/H2O/N2O systems. In particular, the experimental data indicate that the spin‐forbidden dissociation‐recombination reaction between SO2 and O‐atoms is in the fall‐off regime at pressures above 1 atm. The inclusion of a pressure‐dependent rate constant for this reaction, using a high‐pressure limit determined from modeling the consumption of SO2 in a N2O/SO2/N2 mixture at 10.0 atm and 1000 K, brings model predictions into much better agreement with experimentally measured CO profiles over the entire pressure range. Kinetic coupling of NOX and SOX chemistry via the radical pool significantly reduces the ability of SO2 to inhibit oxidative processes. Measurements of SO2 indicate fractional conversions of SO2 to SO3 on the order of a few percent, in good agreement with previous measurements at atmospheric pressure. Modeling results suggest that, at low pressures, SO3 formation occurs primarily through SO2 + O(+M) = SO3(+M), but at higher pressures where the fractional conversion of NO to NO2 increases, SO3 formation via SO2 + NO2 = SO3 + NO becomes important. For the conditions explored in this study, the primary consumption pathways for SO3 appear to be SO3 + HO2 = HOSO2 + O2 and SO3 + H = SO2 + OH. Further study of these reactions would increase the confidence with which model predictions of SO3 can be viewed. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 317–339, 2000  相似文献   

15.
The SO(A → X) fluorescence resulting from the Ar(3P2,0)+SO2 reaction has been studied in a flowing afterglow system. Various reaction pathways for the formation of the SO(A) fragment have been considered. A linear surprisal analysis of the vibrational distributions indicates that the resonance excitation with subsequent predissociation of the intermediate bound state is responsible for the formation of SO(A). A significant fraction of the available energy has been found to be partitioned into translation. Because the predissociation of SO2 in the exit channel is dominating the reaction pathway in the Ar*+SO2 → Ar+O+SO(A) system, all the dynamical features can then be explained by the simple classical treatments on the dissociation of SO2.  相似文献   

16.
Transients of the open-circuit potential observed in the reaction of methanol with oxygen (Oads) preliminarily adsorbed on smooth polycrystalline platinum (pcPt) are measured in 0.05 M HClO4, 0.5 M HClO4, 0.05 M H2SO4, 0.05 M H2SO4 + 0.45 M Na2SO4, and 0.05 M H2SO4 + 0.45 M Cs2SO4. It is shown that the solution pH has a weak effect on the transient characteristics (when the reversible hydrogen electrode potential scale is used). This confirms the chemical nature of rate-controlling stages in the reaction mechanism. The changes in the reaction rate, observed upon going from one electrolyte to another, are preferentially associated with the involvement of solution ions in the formation of activated surface complexes that include CH3OH, Oads, and supporting-electrolyte components.  相似文献   

17.
主要通过XPS表征、热力学计算以及一系列设计的评价实验等方法,对硫化CoMo/Al2O3催化剂上H2同时催化还原SO2和NO反应的活性相、吸附活性位以及反应机理进行了研究.结果表明,金属硫化物相是SO,和NO转化的主要活性相,并与载体Al2O3共同承担H2S转化为单质硫的作用.此外,反应过程中产生的品格空位也对NO转化起着重要作用.催化剂表面的阴离子空位是SO2和NO共同的吸附活性位,SO2对NO的吸附有抑制作用,而催化剂表面的L碱佗也是SO2的吸附活性位,NO可促进SO2的氧化吸附.最后,本文从反应分子的吸附与活化、NO的转化及品格硫的流失、SO2还原到H2S、H2S的转化、晶格硫的补充等5个方面提出了反应机理.  相似文献   

18.
Reactivity of Monophosphine Platinum(0) Complexes with SO2 . The addition reaction of (PPh3)Pt(ViSi) (ViSi = {η2-H2C?CHSiMe2}2O) ( 1 ) with SO2 gives within 30 min the red SO2 complex (PPh3)Pt(η2-H2C?CHSiMe2- OSiMe2CH?CH2)(SO2) ( 2 ). A reaction time of 24 h with SO2 leads to the elimination of the ViSi ligand, and the unstable monomeric intermediate (PPh3)Pt(SO2) cyclo- trimerizes to the stable cluster [Pt(PPh3)(SO2)]3 ( 3 ). 3 is also obtained within 30 min by the reaction of (PPh3)Pt(C2H4)2 ( 4 ) with SO2. The crystal structure of 3 has been determined; space group P21/n, Z = 4, a = 1 606.1(3), b = 1 019.3(1), c = 3 624.6(5) pm, β = 93.67°, R/Rw = 0.102/0.121.  相似文献   

19.
This paper has reported an anionic SO3H-functionalized ionic liquid N-methylimidazolium sulfomethylsulfonate ([Hmim][HO3SCH2SO3]) for the synthesis of coumarins by Pechmann reaction. The [Hmim][HO3SCH2SO3] is easier to prepare by one-step neutralization reaction of N-methylimidazole with methanedisulfonic acid and show high catalytic performance for Pechmann reaction. Besides, the catalyst can simply be separated from the reaction mixture and recycled ten times without noticeable loss of activity.  相似文献   

20.
Hydrolysis of α-cellulose by H2SO4 is a heterogeneous reaction. As such the reaction is influenced by physical factors. The hydrolysis reaction is therefore controlled not only by the reaction conditions (acid concentration and temperature) but also by the physical state of the cellulose. As evidence of this, the reaction rates measured at the high-temperature region (above 200°C) exhibited a sudden change in apparent activation energy at a certain temperature, deviating from Arrhenius law. Furthermore, α-cellulose, once it was dissolved into concentrated H2SO4 and reprecipitated, showed a reaction rate two orders of magnitude higher than that of untreated cellulose, about the same magnitude as cornstarch. The α-cellulose when treated with a varying level of H2SO4 underwent an abrupt change in physical structure (fibrous form to gelatinous form) at about 65% H2SO4. The sudden shift of physical structure and reaction pattern in response to acid concentration and temperature indicates that the main factor causing the change in cellulose structure is disruption of hydrogen bonding. Finding effective means of disrupting hydrogen bonding before or during the hydrolysis reaction may lead to a novel biomass saccharification process.  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