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 234 毫秒
1.
方德彩*  陈彦梅 《化学学报》2014,72(2):253-256
采用多种密度泛函方法并辅以极化连续介质模型研究了四嗪与一些环烯烃发生[4+2]环加成反应的反应机理,得出反应过程按协同机理进行,即反应过程中只有一个过渡态. 主要使用的方法有:CAM-B3LYP,B3LYP,X3LYP,BMK,LC-wPBE,wB97x,wB97xd,M062x和M11. 这种过渡态的稳定性与前线分子轨道的相互作用有关,从过渡态电荷迁移的方向来看,环烯烃作为电子给体,而四嗪作为电子受体. 反应的难易取决于环烯烃的环的大小,环越大反应越难,这与环在形成过渡态时的变形能有关. 通过气相平动熵和溶液平动熵计算得到的活化自由能的数据,比较发现气相平动能得出的活化自由能离实验估计值相差很远. 而对于溶液中活化自由能垒数据对于大多数计算方法所得的结果比较接近,其中BMK,CAM-B3LYP和X3LYP的结果更合理一些. 同时也发现M062x和M11方法计算的反应速度常数与实验值差别较大,说明这些方法不太适合用来研究此类反应;而考虑色散作用的wB97xd也过高估计过渡态时分子之间的作用能,导致低自由能垒,具体产生的原因在文中进行了详细的讨论.  相似文献   

2.
过氧烷基自由基分子内氢迁移是低温燃烧反应中的一类重要基元反应. 本文用等键反应方法计算了该类反应的动力学参数. 所有反应物、过渡态、产物的几何结构均在B3LYP/6-311+G(d,p)水平下优化得到. 本文提出了用过渡态反应中心几何结构守恒作为反应类判据, 并将该分子内氢迁移反应分为四类, 包括(1,3)、(1,4)、(1,5)、(1,n) (n=6, 7, 8)氢迁移类. 分别将这4 类反应类中最小反应体系作为类反应的主反应, 并分别在B3LYP/6-311+G(d,p)低水平和CBS-QB3 高水平下得到其近似能垒和精确能垒. 其余氢迁移反应作为目标反应, 在B3LYP/6-311+G(d,p)低水下计算得到其近似能垒, 再采用等键反应方法校正得到目标反应的精确反应势垒和精确速率常数. 研究表明, 采用等键反应方法只需在低水平用从头算计算就可以得到大分子反应体系的高精度能垒和速率常数值, 且本文按等键反应本质的分类方法更能揭示反应类的本质, 并对反应类的定义给出了客观标准. 本文的研究为碳氢化合物低温燃烧模拟中重要的过氧烷基分子内氢迁移反应提供了准确的动力学参数.  相似文献   

3.
氢过氧自由基从烷烃分子中提取氢的反应是碳氢燃料中低温燃烧化学中非常重要的一类反应。本文用等键反应方法计算了这一类反应的动力学参数。所有反应物、过渡态、产物的几何结构均在HF/6-31+G(d)水平下优化得到。以反应中的过渡态反应中心的几何结构守恒为判据,该反应类可用等键反应处理。本文选取了乙烷和氢过氧自由基的氢提取反应为参考反应,其它反应作为目标反应,用等键反应方法对目标反应在HF/6-31+G(d)水平的近似能垒和反应速率常数进行了校正。为了验证方法的可靠性,选取C5以下的烷烃分子体系,对等键反应方法校正结果和高精度CCSD(T)/CBS直接计算结果进行了比较,最大绝对误差为5.58k J?mol~(-1),因此,采用等键反应方法只需用低水平HF从头算方法就可以再现高精度CCSD(T)/CBS计算结果,从而解决了该反应类中大分子体系的能垒的精确计算。本文的研究为碳氢化合物中低温燃烧模拟中重要的烷烃与氢过氧自由基氢提取反应提供了准确的动力学参数。  相似文献   

4.
采用量子化学方法研究了十氢化萘低温燃烧的动力学机理,获得了脱氢反应、自由基加氧反应及1,5氢迁移反应等反应的动力学参数,并在CBS-QB3水平下获得了相关物种的热力学参数,通过过渡态理论计算获得了具有紧致过渡态反应的高压极限速率常数,而无能垒反应的速率常数则由变分过渡态理论得到.基于此机理分析了十氢化萘低温反应的动力学规律和热力学机制.相比于链烷烃和单环烷烃,十氢化萘自由基加氧反应的速率常数随温度变化较快,1,5-氢迁移反应的能垒较高,揭示了物质结构对反应动力学的影响.热力学平衡常数分析结果表明,在低温下十氢化萘自由基加氧反应起主导作用.通过拟合获得了所有反应Arrhenius形式的速率常数,这些参数可用于双环烷烃低温燃烧机理的构建和优化.  相似文献   

5.
王红  何桥  谭凯 《化学学报》2013,71(12):1663-1667
采用MP2和密度泛函M06-2X方法,在6-31++G(d,p)基组水平上对烯丙基类不对称醚异构化反应机理进行了计算研究. 揭示了其可能的反应途径,预测了互变异构吉布斯自由能,活化能等性质. 计算结果表明,在没有金催化剂的条件下,尽管有醇溶剂时异构化活化能垒有所降低,异构化反应依然不容易进行. 相反,存在金催化剂并且有醇溶剂情况下,烯丙基类不对称醚异构化反应活化自由能大大降低,仅为7.5 kcal/mol. 通过比较有无醇溶剂和金催化剂对异构化的影响,揭示了金烯烃络合和醇分子参与反应以质子转移的异构化反应机理,很好解释了实验中观察的现象. 计算结果还表明:醇分子不仅参与反应提供质子转移,它还能与醚竞争金催化剂络合,因此在高浓度醇条件下会抑制异构化反应进行.  相似文献   

6.
采用密度泛函理论(DFT)对锰配合物催化二氧化碳加氢生成甲酸的反应进行了理论研究. 整个催化循环主要包括氢气活化和二氧化碳氢化2个阶段. 计算结果表明, 甲酸的参与明显降低了氢气活化的反应能垒; 二氧化碳的氢化过程遵循外层机理并且氢转移是分步进行的, 决速步骤为氢负离子的转移过程, 自由能垒为21.0 kJ/mol. 对配合物中硫原子上的取代基R进行了调变, 研究结果表明, 当R为吸电子基团时能降低氢气裂解和二氧化碳氢化过程中质子转移的能垒, 而当R为推电子基团时有利于氢负离子的转移,当R=CF3时整个反应的能量跨度(80.4 kJ/mol)最小.  相似文献   

7.
邓园  王思  丰海松  张欣 《高等学校化学学报》2023,44(2):20220486-151
糠醛催化加氢反应工艺主要分为气相、液相以及催化转移加氢等.相比于糠醛气相加氢,液相加氢为反应提供了更多的可持续性和自由度,但其中溶剂依赖现象对糠醛定向催化转化的影响机制尚不清晰.针对上述问题,本文选用3种溶剂(甲醇、水和环己烷)为研究对象,采用密度泛函方法,从理论计算角度探究了Pd催化糠醛加氢反应中溶剂效应对反应活性和选择性的重要作用.结果表明,在糠醛加氢反应过程中,溶剂一方面能够形成氢键网络促进质子穿梭,另一方面能够稳定反应物、中间体以及生成物,有效降低C=O加氢的能垒.自由能计算结果表明,在液态水、甲醇和环己烷中,随着溶剂极性的降低(水>甲醇>环己烷),第一步C=O氢化的能垒逐渐降低(0.70 eV>0.68 eV>0.44 eV).在水和甲醇介导的糠醛加氢反应过程中,第一步C=O加氢的反应势垒进一步降低为0.47和0.41 eV.差分电荷密度以及Bader电荷分析表明,反应过程中存在糠醛和Pd催化剂之间的电荷转移.分波态密度(PDOS)分析表明,溶剂的加入使d带中心向靠近费米能级的方向移动,表明Pd催化剂的催化活性得到提高.  相似文献   

8.
采用密度泛函理论方法研究了[Rh(R,R-DIOP)]+[DIOP=(1R,2R)-1,2-O-异丙叉-1,2-二醚氧基-1,2-双(二苯基磷基)乙烷]催化下苯并环丁酮手性聚稠环过程在气相、四氢乙呋喃(THF)及水中的反应机理.计算结果表明,在气相中反应容易进行,经TS2形成六元环的过程为决速步骤,但产物无明显的对映选择性.在THF中,S-和R-通道的C—C键活化能垒仅由79.5和69.3 kJ/mol提高到80.2和88.8 kJ/mol,未改变反应的实质;Rh与2个C原子的配位明显弱于气相,相对于催化剂和反应物自由能之和,S-和R-通道的反应总能垒分别提高到63.8和68.1 kJ/mol.对于S-通道,溶剂THF使经TS2的能垒升至112.0 kJ/mol,仍为整个过程的决速步骤;然而对R-通道,溶剂使经TS1形成五元环过程的能垒升至91.5 kJ/mol,但使经TS2的能垒由78.9 kJ/mol降至77.7 kJ/mol,IM1→TS1成为决速步骤.在以水为溶剂时,经TS1形成五元环的过程成为2个通道的决速步骤,其在S-和R-通道中的能垒分别为102.5和94.9 kJ/mol.因此,溶剂改变了反应的决速步骤及能垒.3种方法均预测R-通道为主反应路径,但THF能明显增加产物的对映选择性.采用自然键轨道(NBO)电荷分析了反应过程中电荷的变化.  相似文献   

9.
采用密度泛函理论的B3LYP方法, 在6- 311++G(d, p)基组水平上研究了CH3CHF自由基与HNCO的微观反应机理, 优化了反应过程中的反应物、中间体、过渡态和产物, 在QCISD(T)/6- 311++G(d, p)水平上计算体系在反应通道各驻点的能量. 振动分析结果和IRC分析结果证实了中间体和过渡态的真实性, 计算所得的成键临界点电荷密度变化也确认了该反应过程, 并找到了七条反应通道. 其中生成氟代烷基酰亚胺稳定分子的通道活化能垒最低, 在该反应体系中是与氢迁移平行竞争较易发生的一条反应通道.  相似文献   

10.
在B3LYP/6-311+G**计算水平上, 采用导体极化连续模型研究了溶剂化效应对6-亚甲基环戊二烯酮与HCN反应生成主要产物b类酸的反应机理的影响. 计算结果表明, 在溶剂中的反应机理与在气相中的反应机理一致. 溶剂化效应使反应路径中各驻点的自由能降低, 稳定化了各物质. 溶液中的活化自由能与气相相比也有所降低, 反应更容易发生, 其中CC进攻方式的活化自由能降低得更多.  相似文献   

11.
Several density functional theory (DFT) methods, such as CAM‐B3LYP, M06, ωB97x, and ωB97xD, are used to characterize a range of ene reactions. The Gibbs free energy, activation enthalpy, and entropy are calculated with both the gas‐ and solution‐phase translational entropy; the results obtained from the solution‐phase translational entropies are quite close to the experimental measurements, whereas the gas‐phase translational entropies do not perform well. For ene reactions between the enophile propanedioic acid (2‐oxo‐1,3‐dimethyl ester) and π donors, the two‐solvent‐involved explicit+implicit model can be employed to obtain accurate activation entropies and free‐energy barriers, because the interaction between the carbonyl oxygen atom and the solvent in the transition state is strengthened with the formation of C?C and O?H bonds. In contrast, an implicit solvent model is adequate to calculate activation entropies and free‐energy barriers for the corresponding reactions of the enophile 4‐phenyl‐1,2,4‐triazoline‐3,5‐dione.  相似文献   

12.
Benchmark calculations of proton affinities and gas-phase basicities of molecules most relevant to biological phosphoryl transfer reactions are presented and compared with available experimental results. The accuracy of proton affinity and gas-phase basicity results obtained from several multi-level model chemistries (CBS-QB3, G3B3, and G3MP2B3) and density-functional quantum models (PBE0, B1B95, and B3LYP) are assessed and compared. From these data, a set of empirical bond enthalpy, entropy, and free energy corrections are introduced that considerably improve the accuracy and predictive capability of the methods. These corrections are applied to the prediction of proton affinity and gas-phase basicity values of important biological phosphates and phosphoranes for which experimental data does not currently exist. Comparison is made with results from semiempirical quantum models that are commonly employed in hybrid quantum mechanical/molecular mechanical simulations. Data suggest that the design of improved semiempirical quantum models with increased accuracy for relative proton affinity values is necessary to obtain quantitative accuracy for phosphoryl transfer reactions in solution, enzymes, and ribozymes.  相似文献   

13.
Marton A  Kocsis E  Inczédy J 《Talanta》1983,30(9):709-712
The adsorption of water vapour on anion-exchangers of various degrees of cross-linking ( x 2, x 4, x 8, x 10) and in different ionic forms (Cl(-), Br(-), I(-)) was studied by the isopiestic technique. The calculated integral free-energy changes were independent of the degree of cross-linking of the resins. With increase in the number of adsorbed water molecules the free-energy functions approached limiting values which were characteristic for the counter-ions. The free-energy change was combined with the enthalpy of water sorption (obtained from direct calorimetric measurements) to obtain the entropy change due to the water uptake. Both the enthalpy and the entropy functions indicated the existence of several processes during the adsorption of water, among which the most relevant are hydration, swelling of the matrix, and dilution of the internal electrolyte of the ion-exchanger.  相似文献   

14.
We present a new theoretical strategy, ab initio rate constants plus integration of rate equations, that is used to characterize the role of entropy in driving high-temperature/low-pressure hydrocarbon chemical kinetics typical of filament-assisted diamond growth environments. Twelve elementary processes were analyzed that produce a viable pathway for converting methane in a feed gas to acetylene. These calculations clearly relate the kinetics of this conversion to the properties of individual species, demonstrating that (1) loss of translational entropy restricts addition of hydrogen (and other radical species) to unsaturated carbon-carbon bonds, (2) rotational entropy determines the direction of the rate-limiting abstraction reactions, and (3) the overall pathway is enhanced by high beta-scission reaction rates driven by translational entropy. These results suggest that the proposed strategy is likely applicable to understand gas-phase chemistry occurring in the systems of combustion and other chemical vapor depositions.  相似文献   

15.
The accuracy of quantum chemical predictions of structures and thermodynamic data for metal complexes depends both on the quantum chemical methods and the chemical models used. A thermodynamic analogue of the Eigen-Wilkins mechanism for ligand substitution reactions (Model A) turns out to be sufficiently simple to catch the essential chemistry of complex formation reactions and allows quantum chemical calculations at the ab initio level of thermodynamic quantities both in gas phase and solution; the latter by using the conductor-like polarizable continuum (CPCM) model. Model A describes the complex formation as a two-step reaction: 1. [M(H2O)x](aq) + L(aq) <==>[M(H2O)x], L(aq); 2. [M(H2O)x], L(aq) <==>[M(H2O)(x-1)L],(H2O)(aq). The first step, the formation of an outer-sphere complex is described using the Fuoss equation and the second, the intramolecular exchange between an entering ligand from the second and water in the first coordination shell, using quantum chemical methods. The thermodynamic quantities for this model were compared to those for the reaction: [M(H2O)x](aq) + L(aq) <==>[M(H2O)(x-1)L](aq) + (H2O)(aq) (Model B), as calculated for each reactant and product separately. The models were tested using complex formation between Zn(2+) and ammonia, methylamine, and ethylenediamine, and complex formation and chelate ring closure reactions in binary and ternary UO(2)(2+)-oxalate systems. The results show that the Gibbs energy of reaction for Model A are not strongly dependent on the number of water ligands and the structure of the second coordination sphere; it provides a much more precise estimate of the thermodynamics of complex formation reactions in solution than that obtained from Model B. The agreement between the experimental and calculated data for the formation of Zn(NH(3))(2+)(aq) and Zn(NH(3))(2)(2+)(aq) is better than 8 kJ/mol for the former, as compared to 30 kJ/mol or larger, for the latter. The Gibbs energy of reaction obtained for the UO(2)(2+) oxalate systems using model B differs between 80 and 130 kJ/mol from the experimental results, whereas the agreement with Model A is better. The errors in the quantum chemical estimates of the entropy and enthalpy of reaction are somewhat larger than those for the Gibbs energy, but still in fair agreement with experiments; adding water molecules in the second coordination sphere improves the agreement significantly. Reasons for the different performance of the two models are discussed. The quantum chemical data were used to discuss the microscopic basis of experimental enthalpy and entropy data, to determine the enthalpy and entropy contributions in chelate ring closure reactions and to discuss the origin of the so-called "chelate effect". Contrary to many earlier suggestions, this is not even in the gas phase, a result of changes in translation entropy contributions. There is no simple explanation of the high stability of chelate complexes; it is a result of both enthalpy and entropy contributions that vary from one system to the other.  相似文献   

16.
Calculations by the B3LYP density functional method with various basis sets and by the QCISD(T)/6-31G(d) ab initio method showed that the main pathway of monomolecular gas-phase decomposition of nitroethylene is that involving a cyclic intermediate, 4H-1,2-oxazete 2-oxide; the barrier of its formation (201.9, 203.9, and 216.5 kJ mol–1, as estimated by various methods) reasonably agrees with the experimental value (191.9 kJ mol–1). The barriers of alternative pathways of gas-phase decomposition of nitroethylene are considerably higher. The barriers of reactions involving radical cations are considerably lower than those of the similar reactions involving molecules. Among all the considered pathways of nitroethylene decomposition, bimolecular pathways are the most favorable energetically.Translated from Zhurnal Obshchei Khimii, Vol. 74, No. 8, 2004, pp. 1327–1342.Original Russian Text Copyright © 2004 by Shamov, Nikolaeva, Khrapkovskii.This revised version was published online in April 2005 with a corrected cover date.  相似文献   

17.
We present here a cell model for evaluating Gibbs energy barriers corresponding to bimolecular reactions (or processes of larger molecularity) in which a loss of translational degrees of freedom takes place along the reaction coordinate. With this model, we have studied the Walden inversion processes: Xa- + H3CXb --> XaCH3 + Xb- (X = F, Cl, Br, and I). In these processes, our model yields an increase of about 2.3-3.4 kcal/mol in Gibbs energy in solution corresponding to the loss of the translational degrees of freedom when passing from separate reactants to the TS in good agreement with experimental data. The corresponding value in the gas phase is about 6.7-7.1 kcal/mol. When the difference between these two figures is used to correct the results obtained by the standard UAHF implementation of the continuum model, the theoretical results are brought significantly closer to the experimental ones. This seems to indicate that for these reactions the parametrization used does not adequately introduce the increase in Gibbs energy corresponding to the constriction of the translational motion of the species along the reaction coordinate when passing from the gas phase to solution. Therefore, we believe that continuum models could perform much better if we released the parametrization process from the task of taking into account the constriction in translation motion in solution, which could be more adequately evaluated using the cell model proposed here, thus allowing it to focus on better reproducing all the remaining solvation effects.  相似文献   

18.
Ab initio molecular orbital calculations have been used to study the base-catalyzed hydrogenation of carbonyl compounds. It is found that these hydrogenation reactions share many common features with S(N)2 reactions. Both types of reactions are described by double-well energy profiles, with deep wells and a low or negative overall energy barrier in the gas phase, while the solution-phase profiles show very shallow wells and much higher barriers. For the hydrogenation reactions, the assembly of the highly ordered transition structure is found to be a major limiting factor to the rate of reaction. In the gas phase, the overall barriers for reactions catalyzed by Group I methoxides increase steadily down the group, due to the decreasing charge density on the metal. On the other hand, for Group II and Group III metals, the overall barriers decrease down the group, which is attributed to the increasing ionic character of the metal-oxygen bond. The reaction with B(OCH(3))(3) has an exceptionally high barrier, which is attributed to pi-electron donation from the oxygen lone pairs of the methoxy groups to the formally vacant p orbital on B, as well as to the high covalent character of the B-O bonds. In solution, these reactivity trends are generally the opposite of the corresponding gas-phase trends. While similar barriers are obtained for reactions catalyzed by methoxides and by tert-butoxides, reactions with benzyloxides have somewhat higher barriers. Aromatic ketones are found to be more reactive than purely aliphatic ketones. Moreover, comparison between catalytic hydrogenation of 2,2,5,5-tetramethylcyclopentanone and pivalophenone shows that factors such as steric effects may also be important in differentiating their reactivity. Solvation studies with a wide range of solvents indicate a steady decrease in barrier with decreasing solvent dielectric constant, with nonpolar solvents generally leading to considerably lower barriers than polar solvents. In practice, a good balance between polarity and catalyst solubility is required in selecting the most suitable solvent for the base-catalyzed hydrogenation reaction.  相似文献   

19.
Pseudorotation reactions of biologically relevant oxyphosphoranes were studied by using density functional and continuum solvation methods. A series of 16 pseudorotation reactions involving acyclic and cyclic oxyphosphoranes in neutral and monoanionic (singly deprotonated) forms were studied, in addition to pseudorotation of PF5. The effect of solvent was treated by using three different solvation models for comparison. The barriers to pseudorotation ranged from 1.5 to 8.1 kcal mol(-1) and were influenced systematically by charge state, apicophilicity of ligands, intramolecular hydrogen bonding, cyclic structure and solvation. Barriers to pseudorotation for monoanionic phosphoranes occur with the anionic oxo ligand as the pivotal atom, and are generally lower than for neutral phosphoranes. The OCH3 groups were observed to be more apicophilic than OH groups, and hence pseudorotations that involve axial OCH3/equatorial OH exchange had higher reaction and activation free energy values. Solvent generally lowered barriers relative to the gas-phase reactions. These results, together with isotope 18O exchange experiments, support the assertion that dianionic phosphoranes are not sufficiently long-lived to undergo pseudorotation. Comparison of the density functional results with those from several semiempirical quantum models highlight a challenge for new-generation hybrid quantum mechanical/molecular mechanical potentials for non-enzymatic and enzymatic phosphoryl transfer reactions: the reliable modeling of pseudorotation processes.  相似文献   

20.
We observe chlorine radical dynamics in solution following two-photon photolysis of the solvent, dichloromethane. In neat CH(2)Cl(2), one-third of the chlorine radicals undergo diffusive geminate recombination, and the rest abstract a hydrogen atom from the solvent with a bimolecular rate constant of (1.35 +/- 0.06) x 10(7) M(-1) s(-1). Upon addition of hydrogen-containing solutes, the chlorine atom decay becomes faster, reflecting the presence of a new reaction pathway. We study 16 different solutes that include alkanes (pentane, hexane, heptane, and their cyclic analogues), alcohols (methanol, ethanol, 1-propanol, 2-propanol, and 1-butanol), and chlorinated alkanes (cyclohexyl chloride, 1-chlorobutane, 2-chlorobutane, 1,2-dichlorobutane, and 1,4-dichlorobutane). Chlorine reactions with alkanes have diffusion-limited rate constants that do not depend on the molecular structure, indicating the absence of a potential barrier. Hydrogen abstraction from alcohols is slower than from alkanes and depends weakly on molecular structure, consistent with a small reaction barrier. Reactions with chlorinated alkanes are the slowest, and their rate constants depend strongly on the number and position of the chlorine substituents, signaling the importance of activation barriers to these reactions. The relative rate constants for the activation-controlled reactions agree very well with the predictions of the gas-phase structure-activity relationships.  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