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 328 毫秒
1.
碳前驱体CH3ArCH2NH2热解反应的热力学和动力学DFT研究   总被引:2,自引:0,他引:2  
在实验研究基础上,通过量子化学理论计算对碳前驱体CH3ArCH2NH2的热裂解机理作了进一步的研究.利用Gaussian98程序包中AM1方法及DFTUB3LYP/3-21G*方法,对化合物5种可能热裂解路径的热力学和动力学计算结果表明,CH3ArCH2NH2热裂解的主反应路径为生成自由基CH3ArCH2*和NH2*,其主反应路径AM1计算的活化能Ea=230.78kJ/mol,DFT计算的活化能Ea=321.18kJ/mol;比较键焓计算的数据与相应的实验数据,发现DFT计算结果与实验结果吻合得较好;通过分析优化的反应物及产物自由基的部分结构参数,了解了理论支持主反应的原因;计算的产物自由基的空间构型表明主反应路径生成的产物自由基相互间若进行稠环缩合反应,将获得分子平面取向性很好的稠环芳烃产物.  相似文献   

2.
甲苯热裂解机理的AM1研究(Ⅱ)动力学分析   总被引:1,自引:0,他引:1  
在(Ⅰ)报热力学的基础上,本文用Gaussian98程序包中AM1法UHF计算,对碳材料用碳前驱体甲苯热裂解机理进行了动力学研究,通过用QST2方法寻找过渡态并经过内禀反应坐标IRC验证。计算得到了甲苯5种热裂解路径的活化能;用过渡态理论,计算了得到了这些路径在298-1073K温度范围内的速率常数。动力学计算结果表明:甲苯在热解温度低于963K时的主反应路径为甲苯热裂解生成苄基自由基的反应;该主反应路径又是经过由反应物→中间体→产物而完成,速控步为反应物→中间体,速控步的活化能△E^O=E(TS1′)-E(R)=227.20kJ.mol^-1;当温度高于963K或1073K左右时,主反应路径转为苯环上脱甲基生成苯基和甲基自由基的路径。以上研究结果与实验结果相一致。  相似文献   

3.
碳源甲基苯热裂解机理的密度泛函动力学研究   总被引:9,自引:0,他引:9  
在热力学研究的基础上,用UB3LYP/3-21G^*方法对甲苯热裂解机理进行了动力学研究。计算得到了甲苯的5种热裂解路径的活化能。用过渡状态理论,计算得到了这些路径在298~1223K温度范围内的速率常数。动力学计算结果表明:甲苯在热解温度低于963K时的主反应路径为甲苯热裂解生成苄基自由基的反应,其速控步的活化能△E~0^θ^≠=402.27kJ/mol;当温度高于963K达1223K左右时,主反应路径转为苯环上脱甲基生成苯基和甲基自由基的路径,该路径的活化能△E~0^θ^≠=456.91kJ/mol。以上研究结果与实验结果相一致。  相似文献   

4.
碳前驱体CH3ArCH2NH2的热解性能及动力学研究   总被引:3,自引:1,他引:2  
通过密闭压力容器法、常压DSC、高压DSC及紫外分光光度定量分析法等实验手段,对液相沉积法制碳/碳复合材料用碳前驱体CH3ArCH2NH2的热裂解行为进行了研究,获得不同温度、不同压力下该碳前驱体的热分解温度和残碳率,用等温动力学和非等温动力学方法获得了热裂解反应的表观活化能,实验结果表明,常压热裂解温度大约为530.15-556.55K,1-3MPa的高压范围内的热裂解温度大约在618.34-675.49K,密闭压力容器中的残碳率为56.23%,常压下的残碳率为28.96%-36.47%,而高压下残碳率可达59.11%,根据基辛格等方法获得了等温条件下和非等温条件下热裂解反应的表观活化能Ea分别为206.78kJ/mol和183.93kJ/mol, 反应级数N~1.  相似文献   

5.
环己烷的热裂解机理   总被引:5,自引:0,他引:5  
用Gaussian 98程序包中AM1法和DFT方法,对液相沉积法制碳/碳(C/C)复合材料的碳源化合物环己烷的热解机理做了量子化学理论研究.通过对化合物6种可能的热裂解路径的热力学和动力学计算,找到了环己烷热裂解的主反应路径.结果表明:(1) AM1与DFT计算均显示,断裂C-C键,最终生成乙烯和2 丁烯的反应通道是环己烷的主要裂解通道,与质谱数据吻合; (2) 除主反应路径外,余下的由易到难生成化合物的顺序为甲基环戊烷 >环己烯 >4 甲基环戊烯 >1,3 丁二烯; (3) AM1方法可以很好地推测较大分子体系的热裂解机理,而DFT方法计算的热力学量更接近实验数据.  相似文献   

6.
以双环戊二烯为原料,经D-A反应及催化加氢合成了高密度烃燃料四氢环戊二烯三聚体(THTCPD).该三聚体的密度为1.082 g/cm3,体积热值为47.5 MJ/L,闪点为120℃,凝固点为48~49℃.采用裂解器与色谱-质谱联用技术,对THTCPD的热裂解进行了在线监测,结果表明温度对裂解反应影响较大.对裂解产物的结构进行了分析,产物以甲烷、乙烯、丙烯、环戊烯、环戊二烯、苯和甲苯为主.依据产物结构及单分子自由基反应模型,推测得到了9种路径的裂解机理.采用X3LYP法进行了各自由基的热力学计算,得到各反应路径的相对能量及路径比.通过不同温度下的裂解转化率,计算得到热裂解反应动力学方程,经线性拟合得到活化能Ea=6.67×104kJ/mol,指前因子A=133.75.  相似文献   

7.
采用密度泛函理论(DFT)的计算方法,研究了铂催化2-烯炔基苯甲醛水合环化反应的微观机理及化学选择性的根源.计算结果表明,首先炔基被催化活化而发生亲核环化生成吡喃铂中间体;接着吡喃铂中间体与烯烃双键发生[3+2]环加成生成铂-碳卡宾复合物;之后,反应将沿2条路径进行,得到产物3a或4a,其中4a的生成需经两步水分子辅助的质子转移过程.生成产物3a需要克服的活化能垒为146.5 k J/mol;对4a的生成,烯醇式和酮式互变异构是决速步聚,当一个水分子参与反应时,对应的能垒为185.8 k J/mol,当2个和3个水分子参与反应时,能垒分别降低到128.1和64.9 k J/mol.因此,水分子参与催化得到产物4a的路径是有利的.另外,反应的选择性与在异构化过程中水的共催化作用有关.以上结果很好地解释了实验现象,并为铂催化水环化反应提供新的见解.  相似文献   

8.
甲苯热解机理的AM1研究(Ⅰ)热力学分析   总被引:2,自引:0,他引:2  
在实验的基础上,本文用Gaussian98程序包中AM1法UHF计算,对碳材料用碳前驱体甲苯的热裂解反应机理进行了研究。在对反应物,产物自由基的结构进行能量梯度法全优化的同时,计算了不同温度下的标准热力学参数(298-1073K)。热力学计算结果表明:(1)当甲苯的热裂解温度相对较低时(773K左右),热力学计算结果首先支持苯环上甲基C-H键的断裂生成苯基自由基并继而生成联二甲苯的反应;随着温度的提高(达1073K时),生成苯自由基和甲基自由基的反应比例将大生成苄基自由基的比例;该反应机理与实验结果基本一致。(2)采用Gaussian98程序包中AM1法中的UHF计算,较适合低级芳香烃热裂解反应机理的理论研究。  相似文献   

9.
采用广义梯度近似(GGA)的密度泛函理论(DFT)(DFT-GGA)并结合平板模型, 研究了甲胺在清洁及磷(P)改性的Mo(100)表面(P-Mo(100))发生C—N键断裂的反应历程(CH3NH2→CH3+NH2). 优化了裂解过程中反应物、过渡态和产物的几何构型, 获得了反应路径上各物种的吸附能及反应的活化能数据. 计算结果表明, 在清洁和磷改性的Mo(100)表面, 甲胺均稳定吸附在顶位, 甲基和氨基最稳定的吸附位置均为桥位. 甲胺的C—N键在P-Mo(100)表面裂解的活化能为2.39 eV, 高于其在清洁表面的活化能(1.99 eV). 这表明Mo(100)表面被预吸附的P原子钝化了. 电子结构分析表明, 改性P原子使得金属Mo的供电子能力减弱, 导致它的d带中心下移, 从而降低了该表面的反应活性, 提高了甲胺的C—N键裂解的活化能. 活化能的分解表明, C—N键在P-Mo(100)与Mo(100)表面裂解的活化能的差异主要体现在初态到过渡态时甲胺的结构变化引起的能量变化(△EdefCH3NH2)、过渡态仅有甲基存在时的吸附能(ETSCH3)和过渡态甲基和氨基的相互作用(EintCH3…NH2). △EdefCH3NH2和ETSCH3使活化能升高幅度大于EintCH3…NH2使活化能降低幅度, 最终导致甲胺的C—N键在P-Mo(100)表面裂解的活化能要高于在Mo(100)表面裂解的活化能.  相似文献   

10.
CH4和CO2合成乙酸中CO2与·H及·CH3相互作用的理论计算   总被引:1,自引:0,他引:1  
章日光  黄伟  王宝俊 《催化学报》2007,28(7):641-645
采用量子化学密度泛函方法,对CO2与H及CH3自由基反应进行了理论计算,给出了CO2与H及CH3自由基相互作用的反应机理,提出了CO2与H及CH3自由基作用的4条反应路径.其中以H和CH3自由基进攻CO2的C原子反应为优先路径,主要产物为乙酸,而甲酸甲酯为动力学禁阻产物.计算结果与实验结果相当吻合.为CH4和CO2两步反应合成含氧化合物提供了理论解释和指导.  相似文献   

11.
The bond dissociation energies (BDEs) and radical stabilization energies (RSEs) which result from 166 reactions that lead to carbon-centered radicals of the type ˙CH(2)X, ˙CHXY and ˙CXYZ, where X, Y and Z are any of the fourteen substituents H, F, Cl, NH(2), OH, SH, CH[double bond, length as m-dash]CH(2), C[triple bond, length as m-dash]CH, BH(2), CHO, COOH, CN, CH(3), and CF(3), were calculated using spin-restricted and -unrestricted variants of the double-hybrid B2-PLYP method with the 6-311+G(3df,2p) basis set. The interactions of substituents X, Y, and Z in both the radicals (˙CXYZ) and in the precursor closed-shell molecules (CHXYZ), as well as the extent of additivity of such interactions, were investigated by calculating radical interaction energies (RIEs), molecule interaction energies (MIEs), and deviations from additivity of RSEs (DARSEs) for a set of 152 reactions that lead to di- (˙CHXY) and tri- (˙CXYZ) substituted carbon-centered radicals. The pairwise quantities describing the effects of pairs of substituents in trisubstituted systems, namely pairwise MIEs (PMIEs), pairwise RIEs (PRIEs) and deviations from pairwise additivity of RSEs (DPARSEs), were also calculated for the set of 61 reactions that lead to trisubstituted radicals (˙CXYZ). Both ROB2-PLYP and UB2-PLYP were found to perform quite well in predicting the quantities related to the stabilities of carbon-centered radicals when compared with available experimental data and with the results obtained from the high-level composite method G3X(MP2)-RAD. Particular selections of substituents or combinations of substituents from the current test set were found to lead to specially stable radicals, increasing the RSEs to a maximum of +68.2 kJ mol(-1) for monosubstituted radicals ˙CH(2)X (X = CH[double bond, length as m-dash]CH(2)), +131.7 kJ mol(-1) for disubstituted radicals ˙CHXY (X = NH(2), Y = CHO), and +177.1 kJ mol(-1) for trisubstituted radicals ˙CXYZ (X = NH2, Y = Z = CHO).  相似文献   

12.
Alkylamines (RCH(2)NH(2), R = H, CH(3), C(2)H(5), C(3)H(7), i-C(3)H(7)) have been investigated by dissociative photoionization by threshold photoelectron photoion coincidence spectroscopy (TPEPICO). The 0 K dissociation limits (9.754 +/- 0.008, 9.721 +/- 0.008, 9.702 +/- 0.012, and 9.668 +/- 0.012 eV for R = CH(3), C(2)H(5), C(3)H(7), i-C(3)H(7), respectively) have been determined by preparing energy-selected ions and collecting the fractional abundances of parent and daughter ions. All alkylamine cations produce the methylenimmonium ion, CH(2)NH(2)+, and the corresponding alkyl free radical. Two isodesmic reaction networks have also been constructed. The first one consists of the alkylamine parent molecules, and the other of the alkyl radical photofragments. Reaction heats within the isodesmic networks have been calculated at the CBS-APNO and W1U levels of theory. The two networks are connected by the TPEPICO dissociation energies. The heats of formation of the amines and the alkyl free radicals are then obtained by a modified least-squares fit to minimize the discrepancy between the TPEPICO and the ab initio values. The analysis of the fit reveals that the previous experimental heats of formation are largely accurate, but certain revisions are suggested. Thus, Delta(f)Ho(298K)[CH(3)NH(2)(g)] = -21.8 +/- 1.5 kJ mol-1, Delta(f)Ho(298K)[C(2)H(5)NH(2)(g)] = -50.1 +/- 1.5 kJ mol(-1), Delta(f)Ho(298K)[C(3)H(7)NH(2)(g)] = -70.8 +/- 1.5 kJ mol(-1), Delta(f)Ho(298K)[C(3)H(7)*] = 101.3 +/- 1 kJ mol(-1), and Delta(f)Ho(298K)[i-C(3)H(7)*] = 88.5 +/- 1 kJ mol(-1). The TPEPICO and the ab initio results for butylamine do not agree within 1 kJ mol-1; therefore, no new heat of formation is proposed for butylamine. It is nevertheless indicated that the previous experimental heats of formation of methylamine, propylamine, butylamine, and isobutylamine may have been systematically underestimated. On the other hand, the error in the ethyl radical heat of formation is found to be overestimated and can be decreased to +/- 1 kJ mol(-1); thus, Delta(f)Ho(298K)[C(2)H(5).] = 120.7 +/- 1 kJ mol(-1). On the basis of the data analysis, the heat of formation of the methylenimmonium ion is confirmed to be Delta(f)Ho(298K)[CH(2)NH(2)+] = 750.3 +/- 1 kJ mol(-1).  相似文献   

13.
The mechanism of the cycloaddition reaction between singlet dichloro‐germylene carbene and aldehyde has been investigated with MP2/6‐31G* method, including geometry optimization and vibrational analysis for the involved stationary points on the potential energy surface. The energies of the different conformations are calculated by zero‐point energy and CCSD (T)//MP2/6‐31G* method. From the potential energy profile, it can be predicted that the reaction has two competitive dominant reaction pathways. The channel (A) consists of four steps: (1) the two reactants (R1, R2) first form an intermediate INT2 through a barrier‐free exothermic reaction of 142.4 kJ/mol; (2) INT2 then isomerizes to a four‐membered ring compound P2 via a transition state TS2 with energy barrier of 8.4 kJ/mol; (3) P2 further reacts with aldehyde (R2) to form an intermediate INT3, which is also a barrier‐free exothermic reaction of 9.2 kJ/mol; (4) INT3 isomerizes to a germanic bis‐heterocyclic product P3 via a transition state TS3 with energy barrier of 4.5 kJ/mol. The process of channel (B) is as follows: (1) the two reactants (R1, R2) first form an intermediate INT4 through a barrier‐free exothermic reaction of 251.5 kJ/mol; (2) INT4 further reacts with aldehyde (R2) to form an intermediate INT5, which is also a barrier‐free exothermic reaction of 173.5 kJ/mol; (3) INT5 then isomerizes to a germanic bis‐heterocyclic product P5 via a transition state TS5 with an energy barrier of 69.4 kJ/mol. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

14.
Single crystal X-ray diffraction studies of [Mn(OS(CH3)2)6](ClO4)2 have shown that the low temperature phase transition, detected by differential scanning calorimetry (DSC) at about 223 K, is associated with the crystal symmetry's reduction from an orthorhombic crystallographic system (Fdd2, No. 43) to a monoclinic one (Cc, No. 9). The analysis of the full width at half maximum of the bands connected with: δd(OClO)F2 and ρ(CH3) vibrational modes in the FT-IR and FT-RS spectra, respectively, registered in the function of temperature, proved that the reorientational motions of ClO4- anions and CH3 groups from (CH3)2SO ligands, began to slow down at temperatures below the phase transition at about 223K. Mean values of activation energy for ClO4- reorientation in the high temperature phase I and low temperature phase II are: Ea(I)≈14 kJ mol(-1) and Ea(II)≈10 kJ mol(-1), respectively. Analogous values for CH3 reorientation are: Ea(I)≈23 kJ mol(-1) and Ea(II)≈1 kJ mol(-1), respectively.  相似文献   

15.
Xiuhui Lu  Xin Che  Leyi Shi  Junfeng Han 《中国化学》2010,28(10):1803-1809
The mechanism of the cycloaddition reaction of forming germanic hetero‐polycyclic compound between singlet germylene carbene and formaldehyde has been investigated with MP2/6‐31G* method, including geometry optimization and vibrational analysis for the involved stationary points on the potential energy surface. The energies of the different conformations are calculated by CCSD (T)//MP2/6‐31G* method. From the potential energy profile, we predict that the cycloaddition reaction of forming germanic hetero‐polycyclic compound between singlet germylene carbene and formaldehyde has two competitive dominant reaction pathways. First dominant reaction pathway consists of four steps: (1) the two reactants (R1, R2) first form an intermediate (INT1) through a barrier‐free exothermic reaction of 117.5 kJ/mol; (2) intermediate (INT1) then isomerizes to a four‐membered ring compound (P2) via a transition state (TS2) with an energy barrier of 25.4 kJ/mol; (3) four‐membered ring compound (P2) further reacts with formaldehyde (R2) to form an intermediate (INT3), which is also a barrier‐free exothermic reaction of 19.6 kJ/mol; (4) intermediate (INT3) isomerizes to a germanic bis‐heterocyclic product (P3) via a transition state (TS3) with an energy barrier of 5.8 kJ/mol. Second dominant reaction pathway is as follows: (1) the two reactants (R1, R2) first form an intermediate (INT4) through a barrier‐free exothermic reaction of 197.3 kJ/mol; (2) intermediate (INT4) further reacts with formaldehyde (R2) to form an intermediate (INT5), which is also a barrier‐free exothermic reaction of 141.3 kJ/mol; (3) intermediate (INT5) then isomerizes to a germanic bis‐heterocyclic product (P5) via a transition state (TS5) with an energy barrier of 36.7 kJ/mol.  相似文献   

16.
应用量子化学从头计算和密度泛函理论(DFT)对HO2+C2H2反应体系的反应机理进行了研究.在B3LYP/6-311G**和CCSD(T)/6-311G**水平上计算了HO2+ C2H2反应的二重态反应势能面.计算结果表明,主要反应方式为自由基HO2的H原子和C2H2分子中的C原子结合,经过一系列异构化,最后分解得到主要产物P1 (CH2O+ HCO).此反应是放热反应,化学反应热为-321.99 kJ·mol-1.次要产物为P2 (CO2 +CH3),也是放热反应.  相似文献   

17.
We have calculated the heats of formation (HOFs) for a series of polyazidocubanes by using the density functional theory (DFT), Hartree-Fock, and MP2 methods with 6-31G* basis set as well as semiempirical methods. The cubane skeleton was chosen for a reference compound, that is, the cubane skeleton was not broken in the process of designing isodesmic reactions. There exists group additivity for the HOF with respect to the azido group. The semiempirical AM1 method also produced reliable results for the HOFs of the title compounds, but the semiempirical MINDO3 did not. The relationship between HOFs and molecular structures was discussed. It was found that the HOF increases 330-360 kJ/mol for each additional number of the azido group being added to the cubane skeleton. The distance between azido groups slightly influences the values of HOFs. The interacting energies of neighbor azido groups in polyazidocubanes are in the range of 2.3 approximately 6.6 kJ/mol, which are so small and less related to the substituent numbers. The average interaction energy between nearest neighbor --N3 groups in the most stable conformer of octaazidocubane is 2.29 kJ/mol at the B3LYP/6-31G* level. The relative stability related to the number of azido groups of the title compounds was assessed based on the calculated HOFs, the energy gaps between the frontier orbitals, and the bond orders of the C--N3 and C--C bonds. The predicted detonation velocity of hepta- and octa-derivatives is over 9 km/s, and the detonation pressure of them is ca. 40 GPa or over.  相似文献   

18.
Diaminohydroxymethyl (1) and triaminomethyl (2) radicals were generated by femtosecond collisional electron transfer to their corresponding cations (1+ and 2+, respectively) and characterized by neutralization-reionization mass spectrometry and ab initio/RRKM calculations at correlated levels of theory up to CCSD(T)/aug-cc-pVTZ. Ion 1+ was generated by gas-phase protonation of urea which was predicted to occur preferentially at the carbonyl oxygen with the 298 K proton affinity that was calculated as PA = 875 kJ mol-1. Upon formation, radical 1 gains vibrational excitation through Franck-Condon effects and rapidly dissociates by loss of a hydrogen atom, so that no survivor ions are observed after reionization. Two conformers of 1, syn-1 and anti-1, were found computationally as local energy minima that interconverted rapidly by inversion at one of the amine groups with a <7 kJ mol-1 barrier. The lowest energy dissociation of radical 1 was loss of the hydroxyl hydrogen atom from anti-1 with ETS = 65 kJ mol-1. The other dissociation pathways of 1 were a hydroxyl hydrogen migration to an amine group followed by dissociation to H2N-C=O* and NH3. Ion 2+ was generated by protonation of gas-phase guanidine with a PA = 985 kJ mol-1. Electron transfer to 2+ was accompanied by large Franck-Condon effects that caused complete dissociation of radical 2 by loss of an H atom on the experimental time scale of 4 mus. Radicals 1 and 2 were calculated to have extremely low ionization energies, 4.75 and 4.29 eV, respectively, which belong to the lowest among organic molecules and bracket the ionization energy of atomic potassium (4.34 eV). The stabilities of amino group containing methyl radicals, *CH2NH2, *CH(NH2)2, and 2, were calculated from isodesmic hydrogen atom exchange with methane. The pi-donating NH2 groups were found to increase the stability of the substituted methyl radicals, but the stabilities did not correlate with the radical ionization energies.  相似文献   

19.
Peptides and proteins may contain post-translationally modified phosphorylated amino acid residues, in particular phosphorylated serine (pSer), threonine (pThr) and tyrosine (pTyr). Following earlier work by Lehmann et al., the [M-H]- anions of peptides containing pSer and pThr functionality show loss of the elements of H3PO4. This process, illustrated for Ser (and using a model system), is CH3CONH-C(CH2OPO3H2)CONHCH(3) --> [CH3CONHC(==CH2)CONHCH3 (-OPO3H2)] (a) --> [CH3CONHC(==CH2)CONHCH3-H]- + H3PO4, a process endothermic by 83 kJ mol(-1) at the MP2/6-31++G(d,p)//HF/6-31++G(d,p) level of theory. In addition, intermediate (a) may decompose to yield CH3CONHC(==CH2)CONHCH3 + H2PO4 - in a process exothermic by 3 kJ mol(-1). The barrier to the transition state for these two processes is 49 kJ mol(-1). Characteristic cleavages of pSer and pThr are more energetically favourable than the negative ion backbone cleavages of peptides described previously. In contrast, loss of HPO3 from [M-H]- is characteristic of pTyr. The cleavage [NH2CH(CH2-C6H4-OPO3H-)CO2H] --> [NH2C(CH2-C6H4-O-)CO2H (HPO3)] (b) --> NH2CH(CH2-C6H4-O-)CO2H + HPO3 is endothermic by 318 kJ mol(-1) at the HF/6-31+G(d)//AM1 level of theory. In addition, intermediate (b) also yields NH2CH(CH2-C6H4-OH)CO2H + PO3 - (reaction endothermic by 137 kJ mol(-1)). The two negative ion cleavages of pTyr have a barrier to the transition state of 198 kJ mol(-1) (at the HF/6-31+G(d)//AM1 level of theory) comparable with those already reported for negative ion backbone cleavages.  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