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 31 毫秒
1.
The present computational mechanistic study explores comprehensively the organoactinide‐mediated intramolecular hydroamination/cyclisation (IHC) of aminodienes by employing a reliable DFT method. All the steps of a plausible catalytic reaction course have been scrutinised for the IHC of (4E,6)‐heptadienylamine 1 t by [(CGC)Th(NMe2)2] precatalyst 2 (CGC=[Me2Si(η5‐Me4C5)(tBuN)]2?). For each of the relevant elementary steps the most accessible pathway has been identified from a multitude of mechanistic possibilities. The operative mechanism involves rapid substrate association/dissociation equilibria for the 3 t ‐S resting state and also for azacyclic intermediates 4 a , 4 s , easily accessible and reversible exocyclic ring closure, supposedly facile isomerisation of the azacycle’s butenyl tether prior to turnover‐limiting protonolysis. The following aspects are in support of this scenario: 1) the derived rate law is consistent with the experimentally obtained empirical rate law; 2) the accessed barrier for turnover‐limiting protonolysis does agree remarkably well with observed performance data; 3) the ring‐tether double‐bond selectivity is consistently elucidated, which led to predict the product distribution correctly. This study provides a computationally substantiated rationale for observed activity and selectivity data. Steric demands at the CGC framework appear to be an efficient means for modulating both performance and ring‐tether double‐bond selectivity. The careful comparison of (CGC)4f‐element and (CGC)5f‐element catalysts revealed that aminodiene IHC mediated by organoactinides and organolanthanides proceeds through a similar mechanistic scenario. However, cyclisation and protonolysis steps, in particular, feature a markedly different reactivity pattern for the two catalyst classes, owing to enhanced bond covalency of early actinides when compared to lanthanides.  相似文献   

2.
Reactive intermediates in the asymmetric Heck reaction between aryl electrophiles and 2,3‐dihydrofuran have been identified by NMR and mass spectrometry, with BINAP or the achiral diphosphanes dppp and dppf as ligands. The major cationic species observed is an alkylpalladium produced by addition of PdAr to the alkene followed by two further PdH‐mediated isomerisation steps. This last species has been characterised at −60° by 1H‐, 13C‐, and 31P‐NMR, including HMQC techniques. The regiochemistry of PdAr and PdH addition to the reactant are opposite as defined by the reaction with (2‐2H1)‐2,3‐dihydrofuran. DFT Calculations on the reaction pathway between [CH2(PH2)]PdPh+ and 2,3‐dihydrofuran reveal several structurally interesting intermediates involving agostic β‐H‐atom, ipso‐Ph or reactant O‐atom bonded to the Pd‐atom, and elucidate the isomerisation pathway.  相似文献   

3.
The course of hydrogenation of hirundigenin ( 7 ) (C21H30O5), either in acetic acid or ethanol, is strongly influenced by the solvent used. In acetic acid tow moles of hydrogen were taken up very quickly under saturation of the double bond and hydrogenolysis of the hydroxyl group. In ethanol, first the cyclohemiketal ring was slowly opened to a triol, and subsequently the double bond was saturated very slowly. In the case of anhydrohirundigenin ( 11 ) hydrogenation of the δ5-double bond occurred rapidly in acetic acid. Subsequently, the δ8, 14-ethylenic linkage was also slowly hydrogenated under formation of two stereomeric trans-derivatives. Vincetogenin consumed two moles of hydrogen in acetic acid, and this resulted in the formation of two isomers as well. Hirundigenin and anhydrohirundigenin are stable toward alkaline hydrolysis, but do rearrange in the presence of acids. One of the isomerisation products of dihydro-anhydrohirundigenin has been isolated in pure form.  相似文献   

4.
The [Ru(CO)H(PNN)] pincer complex based on a dearomatised PNN ligand (PNN: 2‐di‐tert‐butylphosphinomethyl‐6‐diethylaminomethylpyridine) was examined for its ability to isomerise alkenes. The isomerisation reaction proceeded under mild conditions after activation of the complex with alcohols. Variable‐temperature (VT) NMR experiments to investigate the role of the alcohol in the mechanism lend credence to the hypothesis that the first step involves the formation of a rearomatised alkoxide complex. In this complex, the hemilabile diethylamino side‐arm can dissociate, allowing alkene binding cis to the hydride, enabling insertion of the alkene into the metal–hydride bond, whereas in the parent complex only trans binding is possible. During this study, a new uncommon Ru0 coordination complex was also characterised. The scope of the alkene isomerisation reaction was examined.  相似文献   

5.
Some general guidelines for the generation of salts with high fluoride ion donor ability are discussed. A preliminary scale for a limited number of fluoride ion donors is presented, cations are classified according to their anion-cation interaction properties. Examples for fluoride ion transfer reactions are given and the influence of anion-cation interaction on the stabilisation of reactive anions is discussed. In our work mainly TAS fluoride (Me2N)3S+Me3SiF2 has been used as fluoride ion source: (a) for fluoride ion transfer to coordinatively unsaturated sulfur species, to SN and SO multiple bond systems, to SN, PN and CN heterocycles, (b) for the stabilisation of primary products of nucleophilic attacks and of intermediates in isomerisation processes or of intermediates in polymerisation processes, (c) for the generation of “naked” anions by silicon-element bond cleavage, and (d) for the activation of element-fluorine bonds by anion formation.  相似文献   

6.
Coenzyme B12-dependent glutamate mutase catalyses the interconversion of (2S)-glutamic acid and (2S, 3S)-3-methylaspartic acid. The enzyme is unable to accept alternative substrates for the rearrangement reaction but is inhibited by substrates analogues including (2S, 3R)- and (25, 3S)- and (2S, 3S)-3-methylglutamic acid, 2-bromo-2, 3-methanosuccinic acid, (2S)-homocysteic acid. The primary isotope effect uponV max andV/K for the isomerisation of the (2S)-glutamic acid is 3.7 ± 0.2 and 13.5 ± 1.0 respectively. There are two C-H bond breaking steps involved in the isomerization reaction. The relative sizes ofDV andD(V/K) indicate that neither of these steps are cleanly rate limiting.  相似文献   

7.
2-(1′-cis,3′-cis-)- and 2-(1′-cis,3′-trans-Penta-1′,3′-dienyl)-phenol (cis, cis- 4 and cis, trans- 4 , cf. scheme 1) rearrange thermally at 85–110° via [1,7 a] hydrogen shifts to yield the o-quinomethide 2 (R ? CH3) which rapidly cyclises to give 2-ethyl-2H-chromene ( 7 ). The trans formation of cis, cis- and cis, trans- 4 into 7 is accompanied by a thermal cis, trans isomerisation of the 3′ double bond in 4. The isomerisation indicates that [1,7 a] hydrogen shifts in 2 compete with the electrocyclic ring closure of 2 . The isomeric phenols, trans, trans- and trans, cis- 4 , are stable at 85–110° but at 190° rearrange also to form 7 . This rearrangement is induced by a thermal cis, trans isomerisation of the 1′ double bond which occurs via [1, 5s] hydrogen shifts. Deuterium labelling experiments show that the chromene 7 is in equilibrium with the o-quinomethide 2 (R ? CH3), at 210°. Thus, when 2-benzyl-2H-chromene ( 9 ) or 2-(1′-trans,3′-trans,-4′-phenyl-buta1′,3′-dienyl)-phenol (trans, trans- 6 ) is heated in diglyme solution at >200°, an equilibrium mixture of both compounds (~ 55% 9 and 45% 6 ) is obtained.  相似文献   

8.
(6R,9′Z)-Neoxanthin: Synthesis, Physical Properties, Spectra, and Calculations of Its Conformation in Solution The synthesis of pure and crystalline (9′Z)-neoxanthin ( 6 ) is described. MnO2 Oxidation of (9Z)-C15-alcohol 7 at room temperature produces a mixture 8/9 of (9Z)- and (9E)-aldehydes. Predominant formation of the required (9Z)-aldehyde 8 is achieved by performing the oxidation at ? 10°. Condensation of 8 with the mono-Li salt of the symmetrical C10-diphosphonate 10 gave the (9Z)-C25-monophosphonate 11 . The Wittig-Horner condensation of 10 with the allenic C15-aldehyde 1b , under selected conditions allows the preparation of pure and crystalline (9′Z)-15,15′-didehydroneoxanthin ( 12 ) and, after subsequent semireduction, of crystalline (15Z,9′Z)-neoxanthin ( 13 ). Thermal isomerisation of a AcOEt solution of 13 at 95° yields preferentially (9′Z)-neoxanthin ( 6 ). Our crystalline sample shows the highest ?-values in the UV/VIS spectra ever recorded. The CD spectra display a pronounced similarity with those of corresponding violaxanthin isomers. In contrast to the (all-E)-isomer 5 , (9′Z)-neoxanthin undergoes very little isomerisation when heated to its melting point. For comparison purposes, a crystalline probe of 6 is also isolated from lawn mowings. Extensive 1H-and 13C-NMR investigations at 600 MHz of a (D6)benzene solution using 2D-experiments such as COSY, TOCSY, ROESY, HMBC, and HMQC techniques permit the unambiguous assignment of all signals. Force-field calculations of a model system of 6 indicate the presence of several interconverting conformers of the violaxanthin end group, 66% of which possess a pseudoequatorial and 34% a pseudoaxial OH? C(3′). The torsion angle (ω1) around the C(6′)? C(7′) bond, known to be of prime importance for the shape of the CD spectra, varies with values of 87° for 55% and 263° for 45% of the molecules. Therefore, the molecules clearly display a preference for the ‘syn’-position of the C(7′)?C(8′) bond and the epoxy group. Unexpectedly, the double bonds of C(7′)?C(8′) and C(9′)?C(10′) are not coplanar. The deviation amounts to ± 20°, both in the ‘syn’ - and the ‘anti’-conformation.  相似文献   

9.
Uranium(IV)–carbene–imido complexes [U(BIPMTMS)(NCPh3)(κ2N,N′‐BIPY)] ( 2 ; BIPMTMS=C(PPh2NSiMe3)2; BIPY=2,2‐bipyridine) and [U(BIPMTMS)(NCPh3)(DMAP)2] ( 3 ; DMAP=4‐dimethylamino‐pyridine) that contain unprecedented, discrete R2C=U=NR′ units are reported. These complexes complete the family of E=U=E (E=CR2, NR, O) metalla‐allenes with feasible first‐row hetero‐element combinations. Intriguingly, 2 and 3 contain cis‐ and trans‐C=U=N units, respectively, representing rare examples of controllable cis/trans isomerisation in f‐block chemistry. This work reveals a clear‐cut example of the trans influence in a mid‐valent uranium system, and thus a strong preference for the cis isomer, which is computed in a co‐ligand‐free truncated model—to isolate the electronic trans influence from steric contributions—to be more stable than the trans isomer by approximately 12 kJ mol?1 with an isomerisation barrier of approximately 14 kJ mol?1.  相似文献   

10.
The objective of this work is to propose a way to calculate approximate transition state geometries that can then be used as initial guesses in ab initio calculations. Transition state geometries are calculated for 26 hydrogen transfer reactions and 6 methyl transfer reactions at the MP2/6-31G* and MP2/6-311++G(d,p) levels. Selected cases are also done at other levels including CCSD(T)/6-311++G(d,p). The transition state geometry obeys an equation which arises from an extension of the Marcus equation proposed by Blowers and Masel [8]: In this equation, r B ,equ is the equilibrium bond length for the bond that breaks during the reaction, r F ,equ is the equilibrium bond length for the new bond which forms. r t B and r t F are the bond lengths at the saddle point in the potential energy surface. r t B and r t F are found to obey with an average error of 0.04 ?. In the last two equations above, ΔU is the heat of reaction, E 0 A is the intrinsic barrier, and C A is a constant that comes from the model of Blowers and Masel [8]. It is proposed that the above three equations are useful in generating initial guesses for transition state geometries in ab initio calculations. In the cases that were tried, rapid convergence was found when these guesses were used. Received: 23 February 2000 / Accepted: 13 May 2000 / Published online: 27 September 2000  相似文献   

11.
The gas-phase isomerisation reaction of glutamine radical cation from [NH2CH (CH2CH2CONH2) COOH ]+• to [ NH2C (CH2CH2CONH2) C (OH)2]+• has been studied theoretically using the MPWB1K functional approach. The [ NH2 C (CH2CH2CONH2) C (OH)2]+• diol species has been found to be the most stable isomer for glutamine radical cation. Moreover, it has been observed that glutamine has a long enough side-chain with basic groups that acts as a solvent molecule favouring the proton-transfer from C α to COOH group. This fact reduces dramatically the isomerisation energy barriers compared to the same process for glycine radical cation in gas phase. Thus, this reaction can be considered as an example of gas-phase proton-transport catalysed reaction in which the proton-transport is carried out by the reactant molecule itself instead of any solvent. Contribution to the Serafin Fraga Memorial Issue.  相似文献   

12.
In pursuit of Platonic hydrocarbon dodecahedrane1, a retrosynthetic theme indicated in scheme 1, was formulated. The precursor tetraquinanedione synthon5 was first designed through a photo-thermal olefin metathesis approach. The tetraquinanedione 5 was further elaborated toexo, exo-tetraquinane diester15 through carbonyl homologation, oxidation, esterification sequence, scheme 5. Bis-cyclopentannulation ofexo,exo-diester15 by Greene methodology delivered a functionalised C20-hexaquinane44, havingexo-annulated cyclopentane rings. Cyclopentane inversion was achieved by a set of reactions involving enone generation, double bond isomerisation and hydrogenation to give spheroidal (C2v)-C20-hexaquinanedione-diester47, the penultimate precursor of dodecahedrane 1. Several interesting transformations and rearrangements of polyquinanes are also described. IUPAC nomenclature: Undecacyclo[9.9.0.02,9.03,7.04,20.05,18.06,16.08,15.010,14.012,19.013,17]eicosane.  相似文献   

13.
Linear isomers of C6N radical differ in the position of the nitrogen atom in the carbon chain of C6N. Reaction routes, involving intramolecular nitrogen atom insertion at varying position in the carbon chain of C6N, are analyzed for the isomerisation between linear isomers of C6N. Through an automated and systematic search performed with global reaction route mapping of the potential energy surface, thermal isomerisation pathways for C6N radical are proposed based on the computations carried out at CASSCF/aug‐cc‐pVTZ, and CCSD(T)/6‐311++G(d,p)//B3LYP/6‐311++G(d,p) levels of the theory. Notably, a high lying linear isomer, centrosymmetric with respect to the nitrogen atom, is observed to be stabilized by a unique crossover three center‐four electron π long bond between the carbon atoms that are spatially separated by a nitrogen atom in a natural bond orbital. This long bond is concluded to be responsible for the predicted thermal isomerisation to be more feasible than the dissociation during the isomerisation pathway of a linear isomer of C6N. © 2014 Wiley Periodicals, Inc.  相似文献   

14.
The dichloromethane solvates of the isomers tetrakis(μ‐1,3‐benzothiazole‐2‐thiolato)‐κ4N:S4S:N‐dipalladium(II)(PdPd), (I), and tetrakis(μ‐1,3‐benzothiazole‐2‐thiolato)‐κ6N:S2S:N‐dipalladium(II)(PdPd), (II), both [Pd2(C7H4NS2)4]·CH2Cl2, have been synthesized in the presence of (o‐isopropylphenyl)diphenylphosphane and (o‐methylphenyl)diphenylphosphane. Both isomers form a lantern‐type structure, where isomer (I) displays a regular and symmetric coordination and isomer (II) an asymmetric and distorted structure. In (I), sitting on an centre of inversion, two 1,3‐benzothiazole‐2‐thiolate units are bonded by a Pd—N bond to one Pd atom and by a Pd—S bond to the other Pd atom, and the other two benzothiazolethiolate units are bonded to the same Pd atoms by, respectively, a Pd—S and a Pd—N bond. In (II), three benzothiazolethiolate units are bonded by a Pd—N bond to one Pd atom and by a Pd—S bond to the other Pd atom, and the fourth benzothiazolethiolate unit is bonded to the same Pd atoms by, respectively, a Pd—S bond and a Pd—N bond.  相似文献   

15.
The process of catalytic isomerisation of various allylic alcohols (alk-1-en-3-ols) into saturated ketones under mild conditions is reported. The water-soluble Na4[{RuCl2(mtppms)2}2] complex, previously reported by us as a precursor to very active hydrogenation catalysts was also found an active catalyst of the redox isomerisation of allylic alcohols in aqueous media. The new Na[Ru(CO)Cp(mtppms)2] as well as Na4[{RuCl(μ-Cl)(CCCPh2)(mtppms)2}2] and Na2[RuClCp(mtppms)2] also showed good to excellent catalytic activities for redox isomerisations in aqueous systems at 50-80 °C under inert atmosphere.  相似文献   

16.
Even systems in which strong electron correlation effects are present, such as the large near-degeneracy correlation in a dissociating electron pair bond exemplified by stretched H2, are represented in the Kohn–Sham (KS) model of non-interacting electrons by a determinantal wavefunction built from the KS molecular orbitals. As a contribution to the discussion on the status and meaning of the KS orbitals we investigate, for the prototype system of H2 at large bond distance, and also for a one-dimensional molecular model, how the electron correlation effects show up in the shape of the KS σ g orbital. KS orbitals φHL and φFCI obtained from the correlated Heitler-London and full configuration interaction wavefunctions are compared to the orbital φLCAO, the traditional linear combination of atomic orbitals (LCAO) form of the (approximate) Hartree-Fock orbital. Electron correlation manifests itself in an essentially non-LCAO structure of the KS orbitals φHL and φFCI around the bond midpoint, which shows up particularly clearly in the Laplacian of the KS orbital. There are corresponding features in the kinetic energy density t s of the KS system (a well around the bond midpoint) and in the one-electron KS potential v s (a peak). The KS features are lacking in the Hartree-Fock orbital, in a minimal LCAO approximation as well as in the exact one. Received: 11 December 1996 / Accepted: 10 January 1997  相似文献   

17.
Catalytic water oxidation at Ir (OH)+ ( Ir =IrCp*(Me2NHC), where Cp*=pentamethylcyclopentadienyl and Me2NHC=N,N′‐dimethylimidazolin‐2‐ylidene) can occur through various competing channels. A potential‐energy surface showing these various multichannel reaction pathways provides a picture of how their importance can be influenced by changes in the oxidant potential. In the most favourable calculated mechanism, water oxidation occurs via a pathway that includes four sequential oxidation steps, prior to formation of the O?O bond. The first three oxidation steps are exothermic upon treatment with cerium ammonium nitrate and lead to formation of Ir V(?O)(O . )+, which is calculated to be the most stabile species under these conditions, whereas the fourth oxidation step is the potential‐energy‐determining step. O?O bond formation takes place by coupling of the two oxo ligands along a direct pathway in the rate‐limiting step. Dissociation of dioxygen occurs in two sequential steps, regenerating the starting material Ir (OH)+. The calculated mechanism fits well with the experimentally observed rate law: v=kobs[ Ir ][oxidant]. The calculated effective barrier of 24.6 kcal mol?1 fits well with the observed turnover frequency of 0.88 s?1. Under strongly oxidative conditions, O?O bond formation after four sequential oxidation steps is the preferred pathway, whereas under milder conditions O?O bond formation after three sequential oxidation steps becomes competitive.  相似文献   

18.
An ab initio study of the structure of Mn(thd)2, Fe(thd)2, and Co(thd)2 complexes in different electronic states is carried out. Quantum chemical calculations are performed using the PC GAMESS program with relativistic effective core pseudopotentials and Gaussian valence triple-zeta basis sets. Calculation methods: DFT/ROB3LYP and CASSCF followed by the inclusion of dynamic electron correlation through multiconfiguration quasi-degenerate second order perturbation theory (MCQDPT2). All three complexes are shown to have a low-spin electronic ground state with a planar structure of the bicyclic fragment at D 2h molecular symmetry. The M—O bond is mainly ionic, and M(thd)2 molecules can be considered as an M2+ cation coordinated by two negatively charged bidentate ligands.  相似文献   

19.
1-Alkyl-2-(arylazo)imidazole (RaaiR/) exists in trans configuration about the –NN- bond. Upon optical excitation in UV region the trans-configuration isomerises to cis-RaaiR/. The photochromism is very susceptible to internal substituents and external environment like solvent polarity, viscosity and presence of innocent foreign species. In this work, the trans-to-cis photoisomerisation of RaaiR/has been studied in DMF solution of cyclodextrin (α/β/γ-CD). The rate of trans-to-cis isomerisation is decreased in presence of CD compared to native form of RaaiR/. The quantum yield of the photoisomerisation is decreased by 35–55% in inclusion phase, [email protected]/, than free photochrome and follows the rate sequence: free state > γ-cyclodextrin > β-cyclodextrin > α-cyclodextrin. The cis-to-trans isomerisation proceeds slowly in visible light irradiation while it is appreciably fast with increasing temperature. The activation energy (Ea) of cistrans thermal isomerisation is also diminished compared to free state of photochrome. The absorption spectral studies have been used in case of Pai-C18H37 with β-CD and inclusion constant is Kb ?= ?0. 121 M?1. The 1H NMR spectral measurement also suggests interaction of aryl protons with cavity protons of β-CD. DFT computation is also attempted to examine the inclusion of RaaiR/with CD and the negative values of binding energy show that the process of inclusion is spontaneous and complexes formed are stable.  相似文献   

20.
The catalytic activity of ruthenium(IV) ([Ru(η33‐C10H16)Cl2L]; C10H16=2,7‐dimethylocta‐2,6‐diene‐1,8‐diyl, L=pyrazole, 3‐methylpyrazole, 3,5‐dimethylpyrazole, 3‐methyl‐5‐phenylpyrazole, 2‐(1H‐pyrazol‐3‐yl)phenol or indazole) and ruthenium(II) complexes ([Ru(η6‐arene)Cl2(3,5‐dimethylpyrazole)]; arene=C6H6, p‐cymene or C6Me6) in the redox isomerisation of allylic alcohols into carbonyl compounds in water is reported. The former show much higher catalytic activity than ruthenium(II) complexes. In particular, a variety of allylic alcohols have been quantitatively isomerised by using [Ru(η33‐C10H16)Cl2(pyrazole)] as a catalyst; the reactions proceeded faster in water than in THF, and in the absence of base. The isomerisations of monosubstituted alcohols take place rapidly (10–60 min, turn‐over frequency=750–3000 h?1) and, in some cases, at 35 °C in 60 min. The nature of the aqueous species formed in water by this complex has been analysed by ESI‐MS. To analyse how an aqueous medium can influence the mechanism of the bifunctional catalytic process, DFT calculations (B3LYP) including one or two explicit water molecules and using the polarisable continuum model have been carried out and provide a valuable insight into the role of water on the activity of the bifunctional catalyst. Several mechanisms have been considered and imply the formation of aqua complexes and their deprotonated species generated from [Ru(η33‐C10H16)Cl2(pyrazole)]. Different competitive pathways based on outer‐sphere mechanisms, which imply hydrogen‐transfer processes, have been analysed. The overall isomerisation implies two hydrogen‐transfer steps from the substrate to the catalyst and subsequent transfer back to the substrate. In addition to the conventional Noyori outer‐sphere mechanism, which involves the pyrazolide ligand, a new mechanism with a hydroxopyrazole complex as the active species can be at work in water. The possibility of formation of an enol, which isomerises easily to the keto form in water, also contributes to the efficiency in water.  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