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 62 毫秒
1.
The thermal decomposition of the atmospheric constituent ethyl formate was studied by coupling flash pyrolysis with imaging photoelectron photoion coincidence (iPEPICO) spectroscopy using synchrotron vacuum ultraviolet (VUV) radiation at the Swiss Light Source (SLS). iPEPICO allows photoion mass-selected threshold photoelectron spectra (ms-TPES) to be obtained for pyrolysis products. By threshold photoionization and ion imaging, parent ions of neutral pyrolysis products and dissociative photoionization products could be distinguished, and multiple spectral carriers could be identified in several ms-TPES. The TPES and mass-selected TPES for ethyl formate are reported for the first time and appear to correspond to ionization of the lowest energy conformer having a cis (eclipsed) configuration of the O = C (H)– O – C (H2)–CH3 and trans (staggered) configuration of the O= C (H)– O – C (H2)– C H3 dihedral angles. We observed the following ethyl formate pyrolysis products: CH3CH2OH, CH3CHO, C2H6, C2H4, HC(O)OH, CH2O, CO2, and CO, with HC(O)OH and C2H4 pyrolyzing further, forming CO + H2O and C2H2 + H2. The reaction paths and energetics leading to these products, together with the products of two homolytic bond cleavage reactions, CH3CH2O + CHO and CH3CH2 + HC(O)O, were studied computationally at the M06-2X-GD3/aug-cc-pVTZ and SVECV-f12 levels of theory, complemented by further theoretical methods for comparison. The calculated reaction pathways were used to derive Arrhenius parameters for each reaction. The reaction rate constants and branching ratios are discussed in terms of the residence time and newly suggest carbon monoxide as a competitive primary fragmentation product at high temperatures.  相似文献   

2.
The production of phosphoryl species (PO, PO2, HOPO) is believed to be of great importance for efficient flame‐retardant action in the gas phase. We present a detailed investigation of the thermal decomposition of dimethyl methylphosphonate (DMMP) probed by vacuum ultraviolet (VUV) synchrotron radiation and imaging photoelectron photoion coincidence (iPEPICO) spectroscopy. This technique provides a snapshot of the thermolysis process and direct evidence of how the reactive phosphoryl species are generated during heat exposure. One of the key findings of this work is that only PO is formed in high concentration upon DMMP decomposition, whereas PO2 is absent. It can be concluded that the formation of PO2 needs an oxidative environment, which is typically the case in a real flame. Based on the identification of products such as methanol, formaldehyde, and PO, as well as the intermediates O?P?CH3, H2C?P?OH, and H2C?P(?O)H, supported by quantum chemical calculations, we were able to describe the predominant pathways that lead to active phosphoryl species during the thermal decomposition of DMMP.  相似文献   

3.
The adsorption and decomposition of trimethylgallium (Ga(CH3)3, TMG) on Pd(111) and the effect of pre-covered H and O were studied by temperature programmed desorption spectroscopy and X-ray photoelectron spectroscopy. TMG adsorbs dissociatively at 140 K and the surface is covered by a mixture of Ga(CH3)x (x=1, 2 or 3) and CHx(a) (x=1, 2 or 3) species. During the heating process, the decomposition of Ga(CH3)3 on clean Pd(111) follows a progressive Ga-C bond cleavage process with CH4 and H2 as the desorption products. The desorption of Ga-containing molecules (probably GaCH3) is also identi ed in the temperature range of 275-325 K. At higher annealing temperature, carbon deposits and metallic Ga are left on the surface and start to di use into the bulk of the substrate. The presence of precovered H(a) and O(a) has a signi cant effect on the adsorption and decomposition behavior of TMG. When the surface is pre-covered by saturated H2, CH4, and H2 desorptions are mainly observed at 315 K, which is ascribed to the dissociation of GaCH3 intermediate. In the case of O-precovered surface, the dissociation mostly occurs at 258 K, of which a Pd-O-Ga(CH3)2 structure is assumed to be the precusor. The presented results may provide some insights into the mechanism of surface reaction during the lm deposition by using trimethylgallium as precursor.  相似文献   

4.
Results of investigations of the adsorption and decomposition of methanol on the surface of transition metals such as Fe, Ni, Cu, Pd, Ag, Mo, W and Pt byuv and x-ray photoelectron spectroscopy, electron energy loss spectroscopy, Auger electron spectroscopy and thermal desorption spectroscopy have been reviewed. The first step in the decomposition of CH3OH on these metal surfaces is the formation of the methoxy species, OCH3 radical. In the case of Fe, Mo and W, complete decomposition of CH3OH occurs leaving CO(β), H2 and CH4 on the surface. Dissociation proceeds upto CO(α) and H2 on the surface of Ni, Pd and Pt whereas on Ag and Cu, selective oxidation of CH3OH to H2CO is preferred. The difference in the reactivity of metals towards CH3OH is rationalised from the heats of adsorption of O2, CO and H2 on these metals. Contribution No. 253 from the Solid State and Structural Chemistry Unit.  相似文献   

5.
The thermal decomposition of formaldehyde was investigated behind shock waves at temperatures between 1675 and 2080 K. Quantitative concentration time profiles of formaldehyde and formyl radicals were measured by means of sensitive 174 nm VUV absorption (CH2O) and 614 nm FM spectroscopy (HCO), respectively. The rate constant of the radical forming channel (1a), CH2O + M → HCO + H + M, of the unimolecular decomposition of formaldehyde in argon was measured at temperatures from 1675 to 2080 K at an average total pressure of 1.2 bar, k1a = 5.0 × 1015 exp(‐308 kJ mol?1/RT) cm3 mol?1 s?1. The pressure dependence, the rate of the competing molecular channel (1b), CH2O + M → H2 + CO + M, and the branching fraction β = k1a/(kA1a + k1b) was characterized by a two‐channel RRKM/master equation analysis. With channel (1b) being the main channel at low pressures, the branching fraction was found to switch from channel (1b) to channel (1a) at moderate pressures of 1–50 bar. Taking advantage of the results of two preceding publications, a decomposition mechanism with six reactions is recommended, which was validated by measured formyl radical profiles and numerous literature experimental observations. The mechanism is capable of a reliable prediction of almost all formaldehyde pyrolysis literature data, including CH2O, CO, and H atom measurements at temperatures of 1200–3200 K, with mixtures of 7 ppm to 5% formaldehyde, and pressures up to 15 bar. Some evidence was found for a self‐reaction of two CH2O molecules. At high initial CH2O mole fractions the reverse of reaction (6), CH2OH + HCO ? CH2O + CH2O becomes noticeable. The rate of the forward reaction was roughly measured to be k6 = 1.5 × 1013 cm3 mol?1 s?1. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 157–169 2004  相似文献   

6.
This study revisits the stability of the possible conformations and the decomposition reactions of ethyl formate in the S0 state using the (U)MP2, MP4SDTQ, CCSD(T), and (U)B3LYP methods with various basis sets. The transition states of the decomposition channels to HCOOH + C2H4, CO + CH3CH2OH, CH2O + CH3CHO, HCOH + CH3CHO, C2H6 + CO2, and H2 + CH2CHOCHO are determined. The microcanonical rate constants derived from the RRKM theory are calculated for each of the decomposition reactions. The high‐pressure limit rate constants are calculated for the decomposition channels to HCOOH + C2H4, CO + CH3CH2OH, and CH2O + CH3CHO.  相似文献   

7.
Complete dehydrogenation of methane is studied on model Pt catalysts by means of state‐of‐the‐art DFT methods and by a combination of supersonic molecular beams with high‐resolution photoelectron spectroscopy. The DFT results predict that intermediate species like CH3 and CH2 are specially stabilized at sites located at particles edges and corners by an amount of 50–80 kJ mol?1. This stabilization is caused by an enhanced activity of low‐coordinated sites accompanied by their special flexibility to accommodate adsorbates. The kinetics of the complete dehydrogenation of methane is substantially modified according to the reaction energy profiles when switching from Pt(111) extended surfaces to Pt nanoparticles. The CH3 and CH2 formation steps are endothermic on Pt(111) but markedly exothermic on Pt79. An important decrease of the reaction barriers is observed in the latter case with values of approximately 60 kJ mol?1 for first C? H bond scission and 40 kJ mol?1 for methyl decomposition. DFT predictions are experimentally confirmed by methane decomposition on Pt nanoparticles supported on an ordered CeO2 film on Cu(111). It is shown that CH3 generated on the Pt nanoparticles undergoes spontaneous dehydrogenation at 100 K. This is in sharp contrast to previous results on Pt single‐crystal surfaces in which CH3 was stable up to much higher temperatures. This result underlines the critical role of particle edge sites in methane activation and dehydrogenation.  相似文献   

8.
Photocatalysis of CH3OH on the ZnO(0001) surface has been investigated by using temperature-programmed desorption (TPD) method with a 266 nm laser light. TPD results show that part of the CH3OH adsorbed on ZnO(0001) surface are in molecular form, while others are dissociated. The thermal reaction products of H2, CH3·, H2O, CO, CH2O, CO2 and CH3OH have been detected. Experiments with the UV laser light indicate that the irradiation can promote the dissociation of CH3OH/CH3O· to form CH2O, which can be future converted to HCOO- during heating or illumination. The reaction between CH3OHZnand OHad can form the H2O molecule at the Zn site. Both temperature and illumination promote the desorption of CH3· from CH3O·. The research provides a new insight into the photocatalytic reaction mechanism of CH3OH on ZnO(0001).  相似文献   

9.
The kinetics and mechanism of Cl-atom initiated reactions of CH3C(O)CHO were studied using the FTIR detection method in the photolysis (λ < 300 nm) of Cl2? CH3C(O)CHO mixtures in 700 torr of N2? O2 diluent at 298 ± 2 K. The observed product distribution over the O2 pressure range from 0–700 torr, combined with relative rate measurements, provided evidence that: (1) the primary step is Cl + CH3C(O)CHO → HCl + CH3C(O)CO with a rate constant of (4.8 ± 1.1) × 10?11 cm3 molecule?1 s?1; and (2) the predominant fate of the primary radical CH3C(O)CO under atmospheric conditions is unimolecular dissociation to CH3C(O) radicals and CO, rather than O2-addition to yield the corresponding carbonylperoxy radical CH3C(O)C(O)OO.  相似文献   

10.
This study investigates the decomposition reactions of ethyl formate in the S1 and T1 states. The dissociation channels leading to HCOOH + C2H4, CH3CH2O + HCO, CH3CH2OCO + H, and CH3CH2 + HCO2 were studied. The major reactions of ethyl formate in the S1 and T1 states are isomerization to the biradical CH2CH2OC(OH)H and dissociation to CH3CH2O + HCO. All the stationary and intersection points were optimized at the CAS(10,8) level of theory with the 6‐31G(d,p) and 6‐311G(2df,2pd) basis sets. Single‐point CASPT3 energy was calculated for each of the stationary and intersection points. Microcanonical rate constants were also calculated for each of the reactions by using the RRKM theory.  相似文献   

11.
The kinetics of the thermal decomposition of acetylacetone has been studied in a shock tube in the temperature range of 1120–1660 K. Detailed analyses of CO and H2O formation data indicate that H2O is formed by a four-center molecular channel, whereas CO is formed by the rapid dissociation of CH3CO produced by the C? C bond dissociation of acetylacetone. The Arrhenius equations for H2O and CH3CO formation channels are ??2 = 1014.24±0.21 exp(?60,800 ± 1,220/RT)sec?1 and ??3 = 1017.05±0.28 exp(?74,600 ± 1,680/RT) sec?1, respectively. The results of the study suggest that the six-center molecular channel for the production of acetone and ketene is not important under the condition used in this investigation.  相似文献   

12.
《Chemphyschem》2003,4(5):466-473
The influence of potassium, in the submonolayer regime, on the adsorption and coadsorption of CO2 and H on a stepped copper surface, Cu(115), has been studied by photoelectron spectroscopy, temperature‐programmed desorption, and work‐function measurements. Based on the fast recording of C 1s and O 1s core‐level spectra, the uptake of CO2 on K/Cu(115) surfaces at 120 K has been followed in real time, and the different reaction products have been identified. The K 2p3/2 peak exhibits a chemical shift of ?0.4 eV with CO2 saturation, the C 1s peaks of the CO3 and the CO species show shifts of ?0.8 and ?0.5 eV, respectively, and the C 1s peak of the physisorbed CO2 exhibits no shift. The effects of gradually heating the CO2/K/Cu(115) surface include the desorption of physisorbed CO2 at 143 K; the desorption of CO at 193 K; the ordering of the CO3 species, and subsequently the dissociation of the carbonate with desorption at 520–700 K. Formate, HCOO?, was synthesized by the coadsorption of H and CO2 on the K/Cu(115) surface at 125 K. Formate formed exclusively for potassium coverages of less than 0.4 monolayer, whereas both formate and carbonate were formed at higher coverages. The desorption of formate‐derived CO2 took place in the temperature range 410–425 K and carbonate‐derived CO2 desorbed at 645–660 K, depending on the potassium coverage.  相似文献   

13.
Atomic clusters are being actively studied for activation of methane, the most stable alkane molecule. While many cluster cations are very reactive with methane, the cluster anions are usually not very reactive, particularly for noble metal free anions. This study reports that the reactivity of molybdenum carbide cluster anions with methane can be much enhanced by adsorption of CO. The Mo2C2? is inert with CH4 while the CO addition product Mo2C3O? brings about dehydrogenation of CH4 under thermal collision conditions. The cluster structures and reactions are characterized by mass spectrometry, photoelectron spectroscopy, and quantum chemistry calculations, which demonstrate that the Mo2C3O? isomer with dissociated CO is reactive but the one with non‐dissociated CO is unreactive. The enhancement of cluster reactivity promoted by CO adsorption in this study is compared with those of reported systems of a few carbonyl complexes.  相似文献   

14.
The thermal decompositions of methyl azidoformate (N3COOMe), ethyl azidoformate (N3COOEt) and 2-azido-N,N-dimethylacetamide (N3CH2CONMe2) have been studied by matrix isolation infrared spectroscopy and real-time ultraviolet photoelectron spectroscopy. N2 appears as an initial pyrolysis product in all systems, and the principal interest lies in the fate of the accompanying organic fragment. For methyl azidoformate, four accompanying products were observed: HNCO, H2CO, CH2NH and CO2, and these are believed to arise as a result of two competing decomposition routes of a four-membered cyclic intermediate. Ethyl azidoformate pyrolysis yields four corresponding products: HNCO, MeCHO, MeCHNH and CO2, together with the five-membered-ring compound 2-oxazolidone. In contrast, the initial pyrolysis of 2-azido-N,N-dimethyl acetamide, yields the novel imine intermediate Me2NCOCH=NH, which subsequently decomposes into dimethyl formamide (HCONMe2), CO, Me2NH and HCN. This intermediate was detected by matrix isolation IR spectroscopy, and its identity confirmed both by a molecular orbital calculation of its IR spectrum, and by the temperature dependence and distribution of products in the PES and IR studies. Mechanisms are proposed for the formation and decomposition of all the products observed in these three systems, based on the experimental evidence and the results of supporting molecular orbital calculations.  相似文献   

15.
The methods of temperature-programmed reaction/desorption (TPR/TPD) are used to study azomethane (CH3N=NCH3) decomposition and the reactions of the products of its pyrolysis (CH 3 * radicals and N2) on the polycrystalline molybdenum surface. A TPR spectrum of adsorbed azomethane decomposition shows mainly N2, H2, and unreacted azomethane. Upon preliminary adsorption of azomethane pyrolysis products on a catalyst sample, a TPR spectrum shows N2, H2, and CH4 in comparable amounts. The difference in the composition of desorption products found for these two types of experiments shows that, in the decomposition of adsorbed azomethane, surface methyl moieties are not formed. The rate constants were calculated for the dissociation of adsorbed CH3, CH2, and CH, recombination of hydrogen atoms with each other and with CH3 and CH2, and the recombinative desorption of nitrogen atoms. Deceased.  相似文献   

16.
The major products of the thermal decomposition of methyl formate in the gas phase are CH(3)OH, CH(2)O, and CO. Experimental studies have proposed that the mechanism to describe these observations involves two key steps: (1) unimolecular decomposition of methyl formate to yield CH(3)OH + CO, followed by (2) thermal decomposition of methanol to yield CH(2)O + H(2). The present study shows that there exists an alternative mechanism that is energetically more favorable. The new mechanism involves two competing parallel unimolecular decomposition pathways to yield the observed major products.  相似文献   

17.
Iodinated hydrocarbons are often used as precursors for hydrocarbon radicals in shock-tube experiments. The radicals are produced by C─I bond fission reaction, and their formation can be followed through time-resolved monitoring of the complementary I-atom concentrations, for example, by I-atom resonance absorption spectroscopy (I-ARAS). This very sensitive technique requires, however, an independent calibration. As a very clean source of I atoms, CH3I is particularly well suited as calibration system for I-ARAS presumed the yield of I atoms and the rate coefficient of I-atom formation from CH3I are known with sufficient accuracy. But if the formation of I atoms from CH3I by I-ARAS is to be characterized, an independent calibration system is required. In this study, we propose a cross-calibration approach for I-ARAS based on the simultaneous time-resolved monitoring of I and H atoms by ARAS in C2H5I pyrolysis experiments. For this reaction system, it can be shown that at sufficiently short reaction times very similar amounts of I and H atoms are formed (difference <1%). As calibration of H-ARAS, with mixtures of N2O and H2, is a well-established technique, we calibrated I-atom absorption–time profiles with respect to simultaneously recorded H-atom concentration–time profiles. Using this approach, we investigated the thermal decomposition of CH3I in the temperature range 950–2050 K behind reflected shock waves at two different nominal pressures (p ∼ 0.4 and 1.6 bar, bath gas: Ar). From the obtained absolute I-atom concentration–time profiles at temperatures T < 1250 K, we inferred a second-order rate coefficient k(T) = (1.7 ± 0.7) × 1015 exp(–20020 K/T) cm3 mol–1 s–1 for the reaction CH3I + Ar → CH3 + I + Ar. A small mechanism to describe the pyrolysis of CH3I under shock-tube conditions is presented and discussed.  相似文献   

18.
The pyrolysis of anisole (C6H5OCH3) was studied behind reflected shock waves via highly sensitive absorption measurements of CO concentration using a rotational transition in the fundamental vibrational band near 4.7 µm. Time‐resolved CO mole fractions were monitored in shock‐heated C6H5OCH3/Ar mixtures between 1000 and 1270 K at 1.3–1.6 bar. The decomposition of C6H5OCH3 proceeds exclusively via homolytic dissociation, with reaction rate k 1, forming methyl (CH3) and phenoxy (C6H5O) radicals. The subsequent decomposition of C6H5O by ring rearrangement and bond dissociation yields CO. To determine the rate constant k 2 of C6H5O decomposition avoiding secondary reactions, allyl phenyl ether (C6H5OC3H5) was used as an alternative source for C6H5O. Its decomposition was studied between 970 and 1170 K at ∼1.4 bar. The potential‐energy surface of C6H5O dissociation has been reevaluated at the G4 level of theory. Rate constants determined from unimolecular rate theory are in good agreement with the present experiments. However, the obtained rates k 2 = 9.1 × 1013 exp(−220.3 kJ mol−1/RT )s−1 are significantly higher than those reported before (factor 6, 2, and 1.5 faster than those data reported by Lin and Lin, J. Phys. Chem . 1986, 90, 425–431; Frank et al., 1994; Carstensen and Dean, 2012, respectively). Good agreement was found between the measured CO concentration profiles and simulations based on the mechanism of Nowakowska et al. after substituting k 2 by the value obtained from experiments on C6H5OC3H5 in this work. The bimolecular reaction of C6H5O and CH3 toward cresol was identified as the most important reaction influencing the CO concentration at longer reaction time.  相似文献   

19.
In this paper previous results are compared for two different types of velocity mapping studies which probe vibrational energy disposal following the A-band photodissociation of methyl iodide, CH3I + hv → CH3 (v) + 1(2P3/2), 1*(2P1/2). Full three-dimensional state-specific speed and angular distributions of the nascent fragments have been recorded for the photoelectrons, iodine atoms, and methyl radicals, using state- and mass-selective (2+1) resonance-enhanced multi-photon ionization (REMPI)/time-of-flight spectrometry. Two sources of information on the vibrational energy disposal are available from velocity mapping: (a) the photoelectron images, which give information on the initial stages of vibrational excitation in electronically excited CH3I, and (b) methyl radical images, which indicate the final energy disposal channels. Even though the two signals are believed to probe very different time-scales of the dissociation process, good agreement between the two is found for the vibrational energy disposal trends. Several trends found in the data for methyl iodide photodissociation indicate that readjustment of the ab initio multi-dimensional potential energy surfaces calculated for this molecule appears to be needed.  相似文献   

20.
A study of the atmospheric photochemical reaction of CF3 radical with CO and O2 was performed by using a homemade ultraviolet photoelectron spectrometer-photoionization mass spectrometer (PES-PIMS). The electronic structures and mechanism of ionization and dissociation of CF3OC(O)OOC(O)-OCF3 were investigated. It was indicated that the two bands on the photoelectron spectrum of CF3OC(O)OOC(O)OCF3 are the result of ionization of an electron from a lone pair of oxygen and a fluorine lone pair of CF3 group. The outermost electrons reside in the oxygen lone pair. The experimental and theoretical first vertical ionization energy is 13.21 and 13.178 eV, respectively, with the PES and OVGF method. They are in good agreement. The photo ionization and dissociation processes were discussed with the help of theoretical calculations and PES-PIMS experiment. After ionization, the parent ions prefer the dissociation of the C—O bond and giving the fragments CF3OCO+ and CF3+. It demonstrated that the ultraviolet photoelectron and photoionization mass spectrometer could be applied widely in the study of atmospheric photochemical reaction. Supported by the Knowledge Innovation Program of the Chinese Academy of Sciences (Grant No. KZCX2-YW-205), Hundred Talents Fund, 973 Program of Ministry of Science and Technology of China (Grant No. 2006CB403701) and the National Natural Science Foundation of China (Grant Nos. 20577052, 20673123)  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