首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 140 毫秒
1.
利用ReaxFF分子动力学模拟方法对正庚烷在高温条件下的热解行为进行了研究.细致分析了温度对正庚烷高温裂解过程以及产物分布的影响.结果显示温度对正庚烷的热解过程的影响是分阶段的.高温能加速正庚烷的分解,但是当温度达到一定阶段之后这种影响逐渐变小.正庚烷的热解可以分为三个阶段.主要产物C2H4、C3和C4的质量百分数随转化率的变化规律与实验值符合很好.利用一阶动力学模型得到的正庚烷热解的表观活化能和指前因子分别为53.96 kcal/mol和55.34×1013 s-1,与实验值相符.  相似文献   

2.
本文介绍了真空紫外光电离质谱结合理论计算研究环戊酮单分子的光电离解离过程. 在9.0∽15.5 eV能量范围内,测量了环戊酮离子及其碎片离子的光电离效率曲线. 通过光电离效率曲线,将环戊酮分子的电离能确定为9.23±0.03 eV,并确认碎片离子为:C5H7O+,C4H5O+,C4H8+,C3H3O+,C4H6+,C2H4O+,C3H6+,C3H5+,C3H4+,C3H3+,C2H5+, C2H4+. 利用量子化学计算方法,在ωB97X-D/6-31+G(d,p)理论水平基础上,提出了C5H8O+的解离机制. 通过对环戊酮解离路径的分析,发现开环和氢迁移过程为环戊酮离子解离的主要路径.  相似文献   

3.
利用CBS-QB3理论计算方法研究了异戊二烯的可能解离通道.获得了主要碎片离子C5H7+,C5H5+,C4H5+,C3H6+,C3H5+,C3H4+,C3H3+的C2H3+的结构以及这些解离通道的解离能,并给出了相应的过渡态和中间体的结构和位垒.得到的异戊二烯电离势及主要碎片离子的出现势均与实验值符合的较好.最后,通过理论和实验结果的对比讨论了各通道的解离机理.  相似文献   

4.
利用具有同步辐射源的反射式飞行时间质谱仪,研究甲基环己烷的真空紫外光电离和光解离. 观测到母体离子C7H14+和碎片离子C7H13+,C6H11+,C6H10+,C5H10+,C5H9+,C4H8+,C4H7+和C3H5+的光电离效率曲线. 测定甲基环己烷的电离能为9.80±0.03 eV,通过光电离效率曲线确定其碎片离子的出现势. 在B3LYP/6-31G(d)水平上对过渡态、中间体和产物离子的优化结构进行表征,并使用G3B3方法计算其能量. 提出主要碎片离子的形成通道. 分子内氢迁移和碳开环是甲基环己烷裂解途径中最重要的过程.  相似文献   

5.
利用同步辐射真空紫外光电离质谱和理论计算研究了胸腺嘧啶的光解离反应路径, 通过改变光子能量得到不同的质谱信号, 光子能量在12.0 eV时主要的碎片有m/z=98 (C4H6N24O+)、97 (C4H5N2O+)、84 (C3H4N2O+或C<  相似文献   

6.
本研究利用质谱和密度泛函理论计算研究了CuC3H-团簇阴离子与CO的反应. 实验结果指出CO与团簇CuC3H-中的C3H-部分偶联可生成唯一产物COC3H-. 此反应的活性和选择性远高于团簇CuC3-与CO的反应. 理论计算结果进一步明确了H辅助的C-C偶联反应.  相似文献   

7.
利用通用型交叉分子束研究了氟原子与1,2-丁二烯的反应,观测到了C4H5F+H反应通道. 测量产物C4H5FF在实验坐标下的角度分辨的飞行时间谱,获得了这个通道质心坐标下的产物角分布和动能分布. 实验结果表明,相对于氟原子束方向,产物C4H5FF主要是后向散射,同时也有大量的前向散射. 这表明反应通道主要通过长寿命的络合物形成机理进行的,同时也伴有直接的双分子亲核取代反应(SN2)机理.  相似文献   

8.
采用激光溅射法制备了同位素标记的氧化锰团簇正离子Mnm18On+,并研究了其在快速流动反应管中与硫化氢在热碰撞条件下的反应,氧化锰团簇正离子与硫化氢反应前后的质量分布与强度变化由飞行时间质谱仪检测.实验表明,绝大多数氧化锰团簇正离子可与硫化氢发生氧-硫交换反应产生水分子,反应通式为:Mnm18On++H2S→Mnm18On-1S++H218O.通过密度泛函理论计算了氧化锰团簇正离Mn2O2+、Mn2O3+和Mn2O4+与H2S反应的机理,结果显示,在这些反应体系中氧-硫交换反应通道同时具有热力学和动力学优势,印证了实验中观察到的现象.气相团簇研究发现的氧-硫交换反应与相关凝聚相体系反应结果一致  相似文献   

9.
利用同步辐射真空紫外光电离质谱技术,在不同光子能量下,研究了异补骨脂素(C11H6O3)的低压热解,探测了不同温度下异补骨脂素的热解产物及其与前驱体的比例. 实验结果表明,异补骨脂素的主要热解产物是CO及其依次消去CO的产物(C10H6O2和C9H6O). 利用密度泛函理论计算异补骨脂素的解离途径,并利用过渡态理论计算了竞争通道的反应速率常数. 通过实验和理论的结合,确定了异补骨脂素主要解离路径和相应产物的分子结构.  相似文献   

10.
用电子密度泛函理论研究了N-质子化corrole(H4Cor+)和meso位芳基取代质子化corroles(H4TPC+、H4TpFPC+和H4TdCPC+)的几何构型、内消旋反应机理以及电子光谱. 结果表明,这些化合物均有两种稳定构型(势能面极小),一个为C2对称性的S1(最稳定构型),另一为C1对称性的S2,其中S1的能量比S2低约15.8~18.5 kJ/mol.S1和S2的corrole环都呈现明显的面外扭曲变形. 手性S1的两个对映异构体之间的转化是一个以S2为中间态的多步过程. 用TDDFT计算了它们的紫外可见电子吸收光谱和圆二色谱(ECD). 与H4Cor+相比,H4TPC+、H4TpFPC+和H4TdCPC+的紫外可见吸收都发生了明显红移,且它们的Q带都因芳基取代基与corrole环之间的π-π共轭而明显增强. 计算表明,质子化corrole的若干相邻电子跃迁的旋转强度符号相反,表明ECD谱可能是研究其电子跃迁的有用工具.  相似文献   

11.
Counterflow diffusion flame experiments and modeling results are presented for a fuel mixture consisting of N2, C2H2, and C2H4 flowing against decomposition products from a solid AP pellet. The flame zone simulates the diffusion flame structure that is expected to exist between reaction products from AP crystals and a hydrocarbon binder. Quantitative species and temperature profiles have been measured for one strain rate, given by a separation of 5 mm, between the fuel exit and the AP surface. Species measured include C2H2, C2H4, N2, CN, NH, OH, CH, C2, NO, NO2, O2, CO2, H2, CO, HCl, H2O, and soot volume fraction. Temperature was measured using a combination of a thermocouple at the fuel exit and other selected locations, spontaneous Raman scattering measurements throughout the flame, NO vibrational populations, and OH rotational population distributions. The burning rate of the AP was also measured for this flame’s strain rate. The measured eighteen scalars are compared with predictions from a detailed gas-phase kinetics model consisting of 105 species and 660 reactions. Model predictions are found to be in good agreement with experiment and illustrate the type of kinetic features that may be expected to occur in propellants when AP particles burn with the decomposition products of a polymeric binder.  相似文献   

12.
Shock-tube and flow-reactor experiments were used to study the thermal decomposition of diethyl carbonate (C2H5OC(O)OC2H5; DEC). The formation of CO2, C2H4, and C2H5OH was measured with gas chromatography/mass spectrometry (GC/MS) and high-repetition-rate time-of-flight mass spectrometry (HRR-TOF-MS) behind reflected shock waves. The same products were also detected by GC/MS in flow reactor experiments. All experiments combined span a temperature range of 663–1203 K at pressures between 1.0 and 2.0 bar. Time-resolved species concentration profiles from HRR-TOF-MS and product compositions from GC/MS measurements were simulated applying a detailed reaction mechanism for DEC combustion. A master-equation analysis was conducted based on computed energies from G4 calculations. Quantum chemical calculations confirm that DEC primarily decomposes by six-center elimination, C2H5OC(O)OC2H5 → C2H4 + C2H5OC(O)OH (1a), followed by rapid decomposition of the alkoxy acid, C2H5OC(O)OH → C2H5OH + CO2 (1b). Measured DEC decomposition rate constants k(T) at p ≈ 1.5 bar can be represented by the Arrhenius equation k(T) = 1013.64±0.12 exp(?204.24±1.95 kJ/mol/RT) s ? 1. Theoretical predictions for k1a were in good agreement with experimentally derived values. The theoretical analysis also included dipropyl carbonate (C3H7OC(O)OC3H7; DPC) decomposition and the reactivities of DEC and DPC are compared and discussed in the context of reactivity of dialkyl carbonates under pyrolytic conditions.  相似文献   

13.
The chemisorption of C2H2 and C2H4 on a clean or partly C- or O-covered Fe(111) surface was investigated with AES, TDS and HREELS. On the clean surface, both molecules adsorb under strong rehybridization close to sp3. Above 230 K, C2H2 reacts to form CH and presumably CH2 as the main products, which on further heating decompose to yield H2 desorption maxima at 580 and 490 K, leaving two carbon species on the surface which correspond to two loss peaks at 400 and 1290 cm?1 in the HREELS spectrum. C2H4 undergoes very rapid decomposition above 250 K; no intermediates have been detected. The presence of coadsorbed oxygen or carbon atoms only reduced the maximum uptake of C2H2, but led to the appearance of new molecular adsorption states of C2H4 and inhibited C2H4 decomposition.  相似文献   

14.
Six new lanthanide(III) complexes (i.e., [Ln(L)2(NA)1.5]·3H2O, where Ln=La(III), Pr(III), Nd(III), Sm(III), Gd(III), and Ce(III) and L and NA indicate N2H4 and C10H6(1-O)(2-COO), respectively) with 1-hydroxy-2-naphthoic acid [C10H6(1-O)(2-COOH)] and hydrazine (N2H4) as co-ligands were characterized by elemental, FTIR, UV-visible, and XRD techniques. In the FT-IR spectra, the N-N stretching frequency in the range of 981–949 cm−1 demonstrates evidence of the presence of coordinated N2H4, indicating the bidentate bridging nature of hydrazine in the complexes. These complexes show symmetric and asymmetric COO stretching from 1444 to 1441 cm−1 and 1582 to 1557 cm−1, respectively, indicating bidentate coordination. TG-DTA studies revealed that the compounds underwent endothermic dehydration from 98 to 110 °C. This was followed by the exothermic decomposition of oxalate intermediates to yield the respective metal oxides as the end products. From SEM images, the average size of the metal oxide particles prepared by thermal decomposition of the complexes was determined to be 39–42 nm. The powder X-ray and SEM coupled with energy dispersive X-ray (EDX) studies revealed the presence of the respective nano-sized metal oxides. The kinetic parameters of the decomposition of the complexes were calculated using the Coats-Redfern equation.  相似文献   

15.
Single-pulse shock-tube experiments were used to study the thermal decomposition of selected oxygenated hydrocarbons: Ethyl propanoate (C2H5OC(O)C2H5; EP), propyl propanoate (C3H7OC(O)C2H5; PP), isopropyl acetate ((CH3)2HCOC(O)CH3; IPA), and methyl isopropyl carbonate ((CH3)2HCOC(O)OCH3; MIC) The consumption of reactants and the formation of stable products such as C2H4 and C3H6 were measured with gas chromatography/mass spectrometry (GC/MS). Depending on the considered reactant, the temperatures range from 716–1102 K at pressures between 1.5 and 2.0 bar. Rate-coefficient data were obtained from first-order analysis. All reactants primarily decompose by six-center eliminations: EP → C2H4 + C2H5COOH (propionic acid); PP → C3H6 + C2H5COOH; IPA → C3H6 + CH3COOH (acetic acid); MIC → C3H6 + CH3OC(O)OH (methoxy formic acid). Experimental rate-coefficient data can be well represented by the following Arrhenius expressions: k(EP → products) = 1013.49±0.16 exp(−214.95±3.25 kJ/mol/RT) s−1; k(PP → products) = 1012.21±0.16 exp(–191.21±2.79 kJ/mol/RT) s−1; k(IPA → products) = 1013.10±0.31 exp(–186.38±5.10 kJ/mol/RT) s−1; k(MIC → products) = 1012.43±0.29 exp(–165.25±4.46 kJ/mol/RT) s−1. The determination of rate coefficients was based on the amount of C2H4 or C3H6 formed. The potential energy surface (PES) of the thermal decomposition of these four reactants was determined with the G4 composite method. A master-equation analysis was conducted based on energies and molecular properties from the G4 computations. The results indicate that the length of a linear alkyl substituent does not significantly influence the rate of six-center eliminations, whereas the change from a linear to a branched alkyl substituent results in a significant reactivity increase. The comparison between rate-coefficient data also shows that alkyl carbonates have higher reactivity towards decomposition by six-center elimination than esters. The results are discussed in in the context of reactivity patterns of carbonyl compounds.  相似文献   

16.
Ablation of polyetheretherketone (PEEK), a high temperature thermoplastic, by XeCl laser radiation occurs at fluences in excess of 0.07±0.01 J cm–2. The volatile products of ablation are CO and C2H2 with smaller quantities of CH4, C4H2, C6H6 and other C3 and C4 hydrocarbons. At fluences close to the threshold ablation produces involatile material of relatively high molecular weight but at high fluences extensive disruption of the PEEK structure occurs with conversion of all of the oxygen in the polymer to carbon monoxide.  相似文献   

17.
The reactive scattering of formic acid from Ni(110) was studied over the temperature range of 175–920 K with MBRS in the millisecond time region by employing a modulation frequency of 36.8 Hz. The steady-state carbon and oxygen composition of the surface varied over the range of temperatures studied. For beam fluxes of 1013 molecules/cm2 sec the onset of decomposition on the steady-state surface occurred at 300 K. By 400 K decomposition was essentially complete, and the products CO2, CO, H2 and H2O were detected. All reaction events were prceded by a common step, and the products were then produced by a series/parallel mechanism. The rate constants measured for H2 and H2O formation indicated stringent limitations on the efficiency of second-order collisions on the surface for producing gaseous products. This study illustrates the use of MBRS for surface reaction mechanistic studies in the millisecond time scale.  相似文献   

18.
The adsorption/decomposition kinetics/dynamics of thiophene has been studied on silica-supported Mo and MoSx clusters. Two-dimensional cluster formation at small Mo exposures and three-dimensional cluster growth at larger exposures would be consistent with the Auger electron spectroscopy (AES) data. Thermal desorption spectroscopy (TDS) indicates two reaction pathways. H4C4S desorbs molecularly at 190–400 K. Two TDS features were evident and could be assigned to molecularly on Mo sites, and S sites adsorbed thiophene. Assuming a standard preexponential factor (ν = 1 × 1013/s) for first-order kinetics, the binding energies for adsorption on Mo (sulfur) sites amount to 90 (65) kJ/mol for 0.4 ML Mo exposure and 76 (63) kJ/mol for 2 ML Mo. Thus, smaller clusters are more reactive than larger clusters for molecular adsorption of H4C4S. The second reaction pathway, the decomposition of thiophene, starts at 250 K. Utilizing multimass TDS, H2, H2S, and mostly alkynes are detected in the gas phase as decomposition products. H4C4S bond activation results in partially sulfided Mo clusters as well as S and C residuals on the surface. S and C poison the catalyst. As a result, with an increasing number of H4C4S adsorption/desorption cycles, the uptake of molecular thiophene decreases as well as the H2 and H2S production ceases. Thus, silica-supported sulfided Mo clusters are less reactive than metallic clusters. The poisoned catalyst can be partially reactivated by annealing in O2. However, Mo oxides also appear to form, which passivate the catalyst further. On the other hand, while annealing a used catalyst in H/H2, it is poisoned even more (i.e., the S AES signal increases). By means of adsorption transients, the initial adsorption probability, S0, of C4H4S has been determined. At thermal impact energies (Ei = 0.04 eV), S0 for molecular adsorption amounts to 0.43 ± 0.03 for a surface temperature of 200 K. S0 increases with Mo cluster size, obeying the capture zone model. The temperature dependence of S0(Ts) consists of two regions consistent with molecular adsorption of thiophene at low temperatures and its decomposition above 250 K. Fitting S0(Ts) curves allows one to determine the bond activation energy for the first elementary decomposition step of C4H4S, which amounts to (79 ± 2) kJ/mol and (52 ± 4) kJ/mol for small and large Mo clusters, respectively. Thus, larger clusters are more active for decomposing C4H4S than are smaller clusters.  相似文献   

19.
The decomposition of ethylene by pulsed, unfocussed CO2-laser radiation has been studied at pressures from 500 to 3000 Torr, using the P(14) line of the 10.6m band (v=949.48cm–1) at incident fluences from about 0.1 to 1.0J/cm2. Major products in order of decreasing importance were 1,3-butadiene, acetylene, ethane, propane, 1-butene and methane. These are known products of the thermal free-radical chain decomposition, and it is concluded that the laser-induced decomposition under our conditions is a transient bulk thermal reaction occurring in a thin disc of heated gas close to the entrance window of the reaction vessel at temperatures ranging from about 1000 to 1500K. As in the thermal decomposition, cyclobutane was observed to be a minor product, which in a sequence of laser pulses approached a final constant concentration. The possibility that this corresponded to an equilibrium concentration at some effective reaction temperature was explored. Computer simulation was used to model the accumulation of cyclobutane in the system, both in a single pulse and in a sequence of pulses, and predictions of this model were compared with experiment. It was concluded that cyclobutane could be used in this way as an approximate internal thermometer, within certain limits. Mechanisms of formation of the free-radical chain products are discussed. It is concluded that the chains are initiated by the bimolecular disproportionation reaction, 2C2H4 C2H3+C2H5, and that secondary initiation by dissociation of the product, 1-butene, becomes increasingly important as the reaction proceeds, leading to autocatalysis. It is further concluded that the radical chain decomposition in this system is a transient process occurring in a brief time interval following the short laser pulse (FWHM=110ns), and is far from steady-state conditions.  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