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 25 毫秒
1.
The mechanism of formation of an eta4-cyclobutadiene complex from a metallacycle, generated by oxidative coupling of two acetylenes with the fragments CpRuCl, [CpRu(PH3)]+, CpCo, and CpRh, was investigated by means of DFT/B3LYP calculations. Two distinct pathways can be envisaged. 1) A multistep reaction, which can be denoted the Vollhardt mechanism, proceeding via a cyclopropenyl carbene and a tetrahedrane-type intermediate. 2) A one-step transformation involving the formation of a third M-C bond with rearrangement of the metallacyclic ring. Although path 2 is definitely favored over path 1, both pathways are energetically prohibitive unless substituents are present on the acetylene. For the CpRuCl system with HC triple bond CR the barrier varies with R in the series H approximately Ph>Me>SiMe3. On going from H to SiMe3, the barrier for path 2 drops from 41.1 to 26.8 kcal mol(-1). This latter value is already reachable, in agreement with experiment. Whereas the reaction mechanisms involving the fragments CpCo, CpRh, and CpRuCl are very similar (but not identical owing to the additional ligand in CpRuCl), those of [CpRu(PH3)]+ reveal a modification with serious consequences. In both paths 1 and 2, the originally planar metallacycle experiences first a bending distortion induced by the sigma-donor strength of P (in contrast to Cl), which compensates the loss of electrons from the ring brought about by bending. The bent metallacycle is already electronically asymmetric and thus the further course of the reaction is facilitated.  相似文献   

2.
The low resolution spectra of φNCS and φNCO have been studied in the microwave R-band region. The spectroscopic constant (B + C) has been evaluated for both molecules. These data have been used to obtain information on the orientation of the linear NCS and NCO group relative to the bond from the phenyl ring. The results indicate an opening of the φ-N-C angle over that found in the hydrogen counterparts HNCS and HNCO. A modified INDO calculation yields structural parameters which are consistent with the trends observed, and indicate that the opening of the angle in going from HNCS (HNCO) to φNCS (φNCO) can be attributed to delocalization with the phenyl ring.  相似文献   

3.
曾艳丽  孟令鹏  郑世钧 《中国化学》2005,23(9):1187-1192
The reactions of HNCO to HOCN, HNCS to HSCN and HNCSe to HSeCN have been studied at MP2/6-311 + + G(2df, pd)//B3LYP/6-311 + +G(2df, pd) level. Geometries of the reactants, transition states and products have been optimized and geometries of the transition states are reported for the first time. The reasons why HNCO and HNCS instead of HOCN and HSCN were easily detected have been explained. It was predicted that HNCSe will be more easily detected than HSeCN. The breakage and formation of the chemical bonds in the reactions have been discussed by the topological analysis method of electronic density. The calculated results show that there are two kinds of structure transition states (STS) in reactions studied.  相似文献   

4.
The electronic structure of HOCN, HSCN, HNCO, and HNCS molecules and [OCN]? and [SCN]? anions has been studied by ab initio calculations at HF/6-31G(d), HF/6-31G(d, p), MP2/6-31G(d)//HF/6-31G(d), and MP2/6-31G(d, p)//HF/6-31G(d, p) levels of theory. The HNCO and HNCS molecules are shown to have higher thermodynamic stability than HOCN and HSCN, respectively. The protolyte strength series are substantiated: HSCN > HOCN, HNCS > HNCO, HOCN > HNCO, HSCN > HNCS. Computations including electron correlation [MP2/6-31G(d)//HF/6-31G(d) and MP2/6-31G(d, p)//HF/6-31G (d, p)] reproduce the general sequence of proton-donor properties: HSCN > HOCN > HNCS > HNCO, which coincides with the hydrophobicity series for the compounds. The relative proton-donor capacity of these acids in water solutions is generally governed by the electronic structure and by the size of their molecules and [OCN]? and [SCN]? anions, but not by medium effects.  相似文献   

5.
The co-adsorption of isocyanic acid (HNCO) and water (H2O) and their reaction to ammonia and carbon dioxide on the anatase phase of TiO2 were studied with ab initio density functional theory (DFT) calculations using a cluster model as well as with in situ DRIFTS investigations and kinetic experiments. We found that isocyanic acid can in principle adsorb both molecularly and dissociatively on the TiO2(1 0 1) surface, but the moment at which water gets involved in the process, is vital for determining the further course of the surface reaction. In the absence of water, it was found that HNCO can adsorb in molecular form on the TiO2 surface. Assuming this case to be the first step of the HNCO hydrolysis, the surface HNCO rearranges into an intermediate complex with a modified NCO skeleton. After decarboxylation water attacks the complex from the gas phase and ammonia is finally formed.

However, when water is present at the beginning of the hydrolysis reaction, it immediately attacks the NCO group present at the surface, yielding a carbamic acid complex, which is further transformed into a carbamate complex. After decarboxylation an NH2 group remains at the surface. Finally, NH3 is formed by hydrogen transfer from molecularly adsorbed water at a neighboring titanium center and the hydrolysis reaction is completed.

Since water is always present in diesel exhaust gas, only the second mechanism is relevant under practical conditions. Moreover, the calculated energy barrier is lower for the second mechanism compared to the first reaction pathway. The comparison between the sum of the theoretical vibrational spectra of the reaction intermediates with the in situ DRIFT spectra also strongly supports the accuracy of the second reaction pathway. The experimental investigation of the kinetics of the HNCO hydrolysis on TiO2-anatase revealed a second order reaction—first order with respect to HNCO and first order with respect to water, which can only be reconciled with the second mechanism.  相似文献   


6.
张绍文  傅孝愿 《物理化学学报》1994,10(11):1004-1008
利用从头算RHF/3-21G方法研究了HNCO二聚后生成HNCNH和CO2的反应机理。计算表明,该反应是分步反应,由反应物经第一过渡态生成四元环中间体,再经过第二过渡态分解为产物,与实验得到的结论一致。反应的第一步是速度控制步骤,计算得到的活化位垒为172.55kJ·mol^-1,与实验上测得的176.40±16.30kJ·mol^-1相吻合。反应的第二位垒为83.68kJ·mol^-1,在实验条  相似文献   

7.
A complete catalytic cycle for the cyclotrimerization of acetylene with the CpRuCl fragment has been proposed and discussed based on DFT/B3LYP calculations, which revealed a couple of uncommon intermediates. The first is a metallacyclopentatriene complex RuCp(Cl)(C(4)H(4)) (B), generated through oxidative coupling of two alkyne ligands. It adds another alkyne in eta(2) fashion to give an alkyne complex (C). No less than three successive intermediates could be located for the subsequent arene formation. The first, an unusual five- and four-membered bicyclic ring system (D), rearranges to a very unsymmetrical metallaheptatetraene complex (E), which in turn provides CpRuCl(eta(2)-C(6)H(6)) (F) via a reductive elimination step. The asymmetry of E, including Cp ring slippage, removes the symmetry-forbidden character from this final step. Completion of the cycle is achieved by an exothermic displacement (21.4 kcal mol(-)(1)) of the arene by two acetylene molecules regenerating A. In addition to acetylene, the reaction of B with ethylene and carbon disulfide, the latter taken as a model for a molecule lacking hydrogen atoms, has also been investigated, and several parallels noted. In the case of the coordinated alkene, facile C-C coupling to the alpha carbon of the metallacycle is feasible due to an agostic assistance, which tends to counterbalance the reduced degree of unsaturation. Carbon disulfide, on the other hand, does not coordinate to ruthenium, but a C=S bond adds instead directly to the Ru=C bond. The final products of the reactions of B with acetylene, ethylene, and carbon disulfide are, respectively, benzene, cyclohexadiene, and thiopyrane-2-thione, the activation energies being lower for acetylene.  相似文献   

8.
HNCS与CH2(X2Π)反应微观动力学的理论研究   总被引:1,自引:0,他引:1  
用量子化学密度泛函理论的UB3LYP/6-311+G**方法和高级电子相关的UQCISD(T)/6-311+G**方法研究了异硫氰酸(HNCS)与乙炔基自由基(C2H(X2Π))反应的微观机理. 采用双水平直接动力学方法IVTST-M, 获取反应的势能面信息, 应用正则变分过渡态理论并考虑小曲率隧道效应, 计算了在250~2500 K温度范围内反应的速率常数. 研究结果表明, HNCS与C2H(X2Π)反应为多通道、多步骤的复杂反应, 共存在三个可能的反应通道, 主反应通道为通过分子间H原子迁移, 生成主要产物NCS+C2H2. 反应速率常数随温度升高而增大, 表现为正温度效应. 速率常数计算中变分效果很小. 在低温区隧道效应对反应速率的贡献较大, 反应为放热反应.  相似文献   

9.
刘朋军  潘秀梅  赵岷  孙昊  苏忠民  王荣顺 《化学学报》2002,60(11):1941-1945
用量子化学密度泛函理论的B3LYP方法,在6-31+G~*水平上按BERNY能量梯度解 析全参数优化了HNCO与CX(X=F,Cl,Br)反应势能面上各驻点的几何构型,通过 振动频率分析确认了中间体和过渡态,内禀反应坐标(IRC)对反应物、中间体、 过渡态和产物的相关性予以证实,对各驻点进行了零点能校正(ZPE)在此基础上 计算了反应能垒。研究结果表明,与HNCO和其它小分子自由基反应不同,HNCO与 CX自由基反应首先发生分子间H原子迁移,随后N与CX的C(1)原子相互靠近成键并 生成较稳定的中间体,再发生N-C(2)键的断裂,完成N向C(1)上的迁移并进一 步解离为产物。反应按反应物→TS1→IM→TS2→产物通道进行。反应为放热反应。  相似文献   

10.
It is shown that kinetic data on the polymerization of acetylene to vinyl acetylene and benzene can be reconciled with the formation of a 1,4 biradical which can isomerize by a 1-3, H-atom shift to the molecular product. Since the biradicals have a negligibly small life-time in the system the overall process appears to be a concerted bimolecular reaction. The labile isomer CH2 ? C: which had been suggested as being the reactive intermediate, is argued on energy considerations not to be a plausible intermediate. Data on the reverse pyrolysis of vinyl acetylene to acetylene are consistent with the model. Extending the model to butadiene explains the observed molecular nature of its decomposition to ethylene and acetylene. Reactions of other oligomers of acetylene are discussed.  相似文献   

11.
The mechanism of the catalytic formation of pyridines from the coupling of two alkynes and the nitriles NCR (R = H, Me, Cl, COOMe) with the fragments CpRuCl, CpCo, and CpRh has been investigated by means of DFT/B3LYP calculations. According to the proposed mechanism, the key reaction step is the oxidative coupling of two alkyne ligands to give metallacyclopentatriene (Ru, Rh) and metallacyclopentadiene (Co) intermediates. In the case of ruthenium, this process is thermodynamically clearly favored over the oxidative coupling between one alkyne and one nitrile ligand to afford an azametallacycle. This alternative pathway however cannot be dismissed in the case of Co and Rh. The rate determining step of the overall catalytic cycle is the addition of a nitrile molecule to the metallacyclopentatriene and metallacyclopentadiene intermediates, respectively, which has to take place in a side-on fashion. Competitive alkyne addition leads to benzene formation. Thus, also the chemoselectivity of this reaction is determined at this stage of the catalytic cycle. In the case of the RuCpCl fragment, the addition of nitriles R-CN and acetylenes RCCH has been studied in more detail. For R = H, Cl, and COOMe the side-on addition of nitriles is kinetically more favored than alkyne addition and, in accordance with experimental results, pyridine formation takes place. In the case of R = Me nitrile addition could not be achieved and the addition of alkynes to give benzene derivatives seems to be kinetically more favored. Once the nitrile is coordinated facile C-C bond coupling takes place to afford an unusual five- and four-membered bicyclic ring system. This intermediate eventually rearranges to a very unsymmetrical azametallaheptatriene complex which in turn provides CpRuCl(κ1-pyridine) via a reductive elimination step. Completion of the catalytic cycle is achieved by an exergonic displacement of the respective pyridine product by two acetylene molecules regenerating the bisacetylene complex.  相似文献   

12.
应用密度泛函理论(DFT), 通过CpRu(PH3)2SH(Cp=环戊二烯基)与HNCS的模型化反应, 探讨了CpRu-(PPh3)2SH与RNCS(R=Ph, 1-naphthyl)反应生成CpRu(PPh3)S2CNHR的两种可能的反应机理. 一种可能的机理是, 一个PH3配体先从反应物CpRu(PH3)2SH解离出来, 得到一个16e中间体, 然后经过一个氢转移反应, 得到产物; 另一种可能的机理是, 先经过一个氢转移反应, 然后一个PH3配体再从金属中心解离出来, 得到产物. 通过分析两种机理的势能曲线发现, 反应的决速步骤为从硫原子到氮原子的氢迁移过程. 第一种反应机理中反应的最高活化能明显比第二种反应机理的最高活化能高. 因此, 我们预测反应倾向于先发生氢迁移, 然后配体PH3再从金属中心上解离出来. 在该反应机理中, 尽管和产物相连的中间体稳定性稍高于产物, 由于熵效应致使最终产物仍然是实验中所得到的产物.  相似文献   

13.
Copper(I)‐catalyzed addition of limited amounts of azides to multiple alkynes, which led to statistical mixtures of triazole/acetylene derivatives or, in other cases, resulted in preferred formation of multiple triazoles, was studied at pre‐organizable calixarene platforms bearing up to four propargyl groups. Depending on calixarene structures and reaction conditions, the unprecedented specific or selective formation of exhaustively triazolated calixarenes or a complete loss of the selectivity were observed. Both autocatalytic copper activation and a local copper(I) concentration increase due to copper–triazole complexation were thoroughly studied as the most expected reasons for the selectivity and both were disproved. Mixed triazolated/propargylated calixarenes and their copper(I) complexes proved not to be involved in the cascade‐like process that was modeled to be driven by an intramolecular transfer of two copper(I) ions from a just‐formed binuclear copper intermediate to the adjacent acetylene unit.  相似文献   

14.
在QCISD(T)/6-311++G(d,p)和B3LYP/6-311++G(d,p)级别上研究了HNCS与Cl原子的反应机理. 并应用经典过渡态理论和正则变分过渡态理论结合小曲率隧道效应, 计算了200-2500 K温度范围内各反应通道的速率常数. 结果表明, HNCS与Cl原子反应存在3个反应通道. 当温度低于294 K时, 生成HCl+NCS的夺氢反应(a)是优势通道, 温度高于294 K时, 生成HNC(Cl)S的加成反应(c)为主反应通道, Cl进攻N的反应通道(b)因能垒较高而难以进行.  相似文献   

15.
Cyclic ortho esters undergo stereoselective and regioselective reaction with phenols when treated with BF(3) x OEt(2) at low temperatures. Attack of the phenol on the ortho ester occurs at an open carbon para to electron-donating groups on the phenol ("C-addition") or at the phenolic hydroxyl group ("O-addition") depending on the nature of the cation formed from reaction of the ortho ester and BF(3) x OEt(2). Products resulting from O-addition undergo reversion to a mixture of starting phenol, C-addition product, and O-addition product if treated with BF(3) x OEt(2) at room temperature, but C-addition products are stable under the same conditions. X-ray structural analysis of the C-addition compound indicates that its stereochemistry is opposite to that observed in reaction of similar ortho esters with chloride from TMSCl. However, the stereochemistry of the reaction can be rationalized by the ability of the ortho ester to isomerize via an intermediate benzylic cation and examination of the preferred trajectory of attack of the nucleophile on the intermediate oxonium ion.  相似文献   

16.
Reaction of acetylene with tris(acetylacetonato)titanium(III) and diethylaluminum chloride system leads to formation of benzene, a trace of ethylbenzene, and a small amount of polyacetylene. The isotopic composition of products obtained from cyclotrimerization of acetylene-d2 and an equimolar mixture of acetylene and acetylene-d2 is investigated to elucidate the mechanism of the cyclotrimerization. The results suggest a mechanism in which an acetylene inserts into the metal—ethyl bond formed by reaction of Ti(acac)3 and Al(C2H5)2Cl, followed by insertion of two acetylene molecules and elimination of a hydrogen atom from the first inserted acetylene to yield an ethylbenzene and a metal hydride intermediate. The metal hydride intermediate catalyzes acetylene cyclotrimerization to give benzene. During the reaction, the hydrogen atom in the metal hydride intermediate does not exchange with the hydrogen atom in the inserted acetylene molecules.  相似文献   

17.
The reaction of a {W(CO)5}‐stabilized phosphinophosphonate 1 , (CO)5WPH(Ph)? P(O)(OEt)2, with ethynyl‐ ( 2 a – f ) and diethynylketones ( 7 – 11 , 18 , and 19 ) in the presence of lithium diisopropylamide (LDA) is examined. Lithiated 1 undergoes nucleophilic attack in the Michael position of the acetylenic ketones, as long as this position is not sterically encumbered by bulky (iPr)3Si substituents. Reaction of all other monoacetylenic ketones with lithiated 1 results in the formation of 2,5‐dihydro‐1,2‐oxaphospholes 3 and 4 . When diacetylenic ketones are employed in the reaction, two very different product types can be isolated. If at least one (Me)3Si or (Et)3Si acetylene terminus is present, as in 7 , 8 , and 19 , an anionic oxaphosphole intermediate can react further with a second equivalent of ketone to give cumulene‐decorated oxaphospholes 14 , 15 , 24 , and 25 . Diacetylenic ketones 10 and 11 , with two aromatic acetylene substituents, react with lithitated 1 to form exclusively ethenyl‐bridged bisphospholes 16 and 17 . Mechanisms that rationalize the formation of all heterocycles are presented and are supported by DFT calculations. Computational studies suggest that thermodynamic, as well as kinetic, considerations dictate the observed reactivity. The calculated reaction pathways reveal a number of almost isoenergetic intermediates that follow after ring opening of the initially formed oxadiphosphetane. Bisphosphole formation through a carbene intermediate G is greatly favored in the presence of phenyl substituents, whereas the formation of cumulene‐decorated oxaphospholes is more exothermic for the trimethylsilyl‐containing substrates. The pathway to the latter compounds contains a 1,3‐shift of the group that stems from the acetylene terminus of the ketone substrates. For silyl substituents, the 1,3‐shift proceeds along a smooth potential energy surface through a transition state that is characterized by a pentacoordinated silicon center. In contrast, a high‐lying transition state TS(E′–F′)R=Ph of 37 kcal mol?1 is found when the substituent is a phenyl group, thus explaining the experimental observation that aryl‐terminated diethynylketones 10 and 11 exclusively form bisphospholes 16 and 17 .  相似文献   

18.
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   

19.
The reaction mechanisms of HNCS with NH(X3∑) were theoretically investigated. The minimum energy paths (MEP) of the reaction were calculated by using the density functional theory(DFT) at the B3LYP/6-311 G** level. The equilibrium structural parameters, the harmonic vibrational frequencies, the total energies, and the zeropoint energies(ZPE) of all the species were calculated. The single-point energies along the MEP were further refined at the QCISD(T)/6-311 G** level. It was found that the mechanisms of the HNCS NH(X3∑) reaction involve two channels producing the HNC HNS and the N2H2 CS products. Channel 1 plays a dominant role and the HNC HNS are the main products. The reaction is exothermic.  相似文献   

20.
The adsorption of HCN and the reaction of HCN with NO(2) over Na-, and Ba-Y,FAU zeolite catalysts were investigated using in situ FTIR and TPD/TPR spectroscopies. Both catalysts adsorb HCN molecularly at room temperature, and the strength of adsorption is higher over Ba-Y than Na-Y. Over Na-Y, the reaction between HCN and NO(2) is slow at 473 K. On Ba-Y, HCN reacts readily with NO(2) at 473K, forming N(2), CO, CO(2), HNCO, NO, N(2)O, and C(2)N(2). The results of this investigation suggest that initial step in the HCN + NO(2) reaction over these catalysts is the hydrogen abstraction from HCN, and the formation of ionic CN- and NC- species. The formation of N(2) can proceed directly from these ionic species upon their interaction with NO+. Alternatively, these cyanide species can be oxidized to isocyanates which then can be further transformed to N(2), N(2)O and CO(x) in their subsequent reaction with NO(x).  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