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 cycloaddition chemistry of several representative unsaturated hydrocarbons (1,3-butadiene, benzene, ethylene, and acetylene) and a heterocyclic aromatic (thiophene) on a Si(111)-7x7 surface has been explored by means of density functional cluster model calculations. It is shown that (i) 1,3-butadiene, benzene, and thiophene can undergo both [4+2]-like and [2+2]-like cycloadditions onto a rest atom-adatom pair, with the former process being favored over the latter both thermodynamically and kinetically; (ii) ethylene and acetylene undergo [2+2] cycloaddition-like chemisorptions onto a rest atom-adatom pair; and (iii) all of these reactions adopt diradical mechanisms. This is in contrast to the [4+2] cycloaddition-like chemisorptions of conjugated dienes on a Si(100) surface and to the prototype [4+2] cycloadditions in organic chemistry, which were believed to adopt concerted reaction pathways. Of particular interest is the [4+2]-like cycloaddition of s-trans-1,3-butadiene, whose stereochemistry is retained during its chemisorption on the Si(111) surface.  相似文献   

2.
The gas‐phase reactions of chlorobenzene with all atomic lanthanide cations Ln+ (except Pm+) have been investigated by using Fourier transform ion cyclotron resonance mass spectrometry in conjunction with density functional theory calculations. According to the latter, a direct chlorine transfer to the lanthanide cation, which has been observed previously for fluorine abstraction from fluorobenzene, is not operative for the C6H5Cl/Ln+ couples; rather, chlorine transfer proceeds through an initial coordination of the lanthanide cation to the aromatic ring of the substrate. Both, the product distribution and the chlorine abstraction efficiencies are affected by the bond dissociation energy (BDE(Ln+?Cl)) as well as the promotion energies of Ln+ to attain a 4fn 5d1 6s1 configuration. In addition, mechanistic aspects of some C?H and C?C bond activations are presented. Where appropriate, comparison with the previously studied C6H5F/Ln+ systems is made.  相似文献   

3.
The probable fragmentation channels of hydroxymethyl radical cation were studied through the H‐and H2‐abstraction and C‐O bond breaking reactions including their related isomerization reactions. The energy barriers for hydroxymethyl cation undergoing isomerization reactions are generally higher than those undergoing the concerted 1,2‐elimination reactions to generate CHO+ and H2. The fragmentation reaction to form CHO+ and H2 through the 1,2‐elimination pathways is the major fragmentation channel for hydroxymethyl cation, consistent with the experimental observation. H abstraction from the hydroxyl group of CH2OH+ is more difficult than that from the methylene group. The feasible path to lose H is to generate CHOH2+ through hydrogen transfer reaction as the first step and then to undergo H‐elimination to generate trans‐CHOH+. Among all the reactions found in this study, the OH‐elimination to generate CH2+ has the highest energy barrier. Our calculation results indicate that the major signals contributed from the related species of hydroxymethyl cation found in the mass spectrum should be m/e 29, m/e 30.  相似文献   

4.
Reaction of thianthrene cation radical perchlorate (Th.+ClO4?) with 1,4-diphenylazomethane (DPAM) in MeCN and EtCN led to the formation of 1,2,4-triazoles. Triazoles formation is attributed to oxidative cycloaddition of benzaldehyde benzylhydrazone, the tautomer of DPAM, to the solvent nitriles. In confirmation, analogous cycloadditions were achieved by reaction of Th.+ClO4? with some benzaldehyde phenylhydrazones in the same solvents.  相似文献   

5.
Hetaryl dieneamines readily available from fused N-aryl azolium salts with secondary amines were found to undergo cycloadditions. With N-phenylmaleinimide and acetylene dicarboxylic ester, [4+2] cycloaddition took place to yield tetrahydroisoindolone (2) and phtalic acid (3) derivatives, respectively. Reaction with fulvene reagent4 resulted in a [6+4] cycloaddition followed by a spontaneous elimination and allowed a convenient route to hetaryl azulenes (5–7).  相似文献   

6.
Kinetics of cycloadditions of phenyl azides to methyl 3-pyrrolidinoacrylate (2) produced Hammett ? = 2.2. An Ea of 13.5 kcal mole ?1 and Δ S* of ?37.4 cal.K?l mole?1 were calculated for the cycloaddition of p-O2NC6H4N3 to 2. The cycloadditions are concerted, non-synchronous, and are controlled by LUMO (azide) - HOMO (dipolarophile) interactions.  相似文献   

7.
Crystal Structure and Raman Spectrum of Trifluoro(methyl)arsoniumHexafluoroarsenate CH3AsF3+AsF6 Trifluoro(methyl)arsonium hexafluoroarsenate was prepared and characterized by Raman spectroscopy and its crystal structure, determined by X-ray diffraction. Trifluoro(methyl)arsonium hexafluoroarsenate crystallises in the monoclinic space group P21/n (no. 14) with Z = 4 and a = 787.5(1), b = 892.9(1), c = 1061.4(1) pm, β = 96.96(1)°, V = 0.74083(14) nm3. The final R factor is 0.0343 for 2009 reflections with I > 2σ(I). The cation has C3v symmetry and the anion shapes a distorted octahedron.  相似文献   

8.
S. Fukuzumi  J.K. Kochi 《Tetrahedron》1982,38(8):1035-1049
The Diels-Alder cycloaddition of anthracene to tetracyanoethylene (TCNE) is quantitatively compared to alkylmetal insertion under the same reaction conditions. In both systems, the observation of transient charge transfer (CT) absorption bands is related to the presence of 1:1 electron donor-acceptor complexes of anthracene (Ar) and alkylmetal (RM) donors with the TCNE acceptor. The activation free energies ΔG3 for anthracene cycloaddition and alkylmetal insertion are found to be equal to the energies of ion-pair formation, i.e. [Ar+TCNE?] and RM+TCNE?], which are evaluated from the CT transition energies hνCT. Indeed, the differences in the rates of alkylmetal insertion and anthracene cycloaddition by a factor of more than 109, are shown quantitatively to arise from the differences in ion-pair solvation ΔGs. The same differences in ΔGs also apply quantitatively to the free ions, [Ar+] and [RM+], independently derived from the electrochemical and iron(III) oxidations of alkylmetals and aromatic compounds, respectively, by outer- sphere electron transfer. The charge transfer formulation of the activation process but provides a unifying basis for comparing such diverse processes as Diels-Alder cycloadditions and organometal cleavages, when a common electron-deficient acceptor is employed. The relationship to the concerted mechanisms of the Diels-Alder reaction is discussed.  相似文献   

9.
Hydride abstraction from diarylamines with the trityl ion is explored in an attempt to generate a stable diarylnitrenium ion, Ar2N+. Sequential H-atom abstraction reactions ensue. The first H-atom abstraction leads to intensely colored aminium radical cations, Ar2NH.+, some of which are quite stable. However, most undergo a second H-atom abstraction leading to ammonium ions, Ar2NH2+. In the absence of a ready source of H-atoms, a unique self-abstraction reaction occurs when Ar=Me5C6, leading to a novel iminium radical cation, Ar=N.+Ar, which decays via a second self H-atom abstraction reaction to give a stable iminium ion, Ar=N+HAr. These products differ substantially from those derived via photochemically produced diarylnitrenium ions.  相似文献   

10.
Quantum chemical calculations using DFT at the B3LYP level have been carried out for the reaction of ethylene with the group-7 compounds ReO2(CH3)(CH2) (Re1), TcO2(CH3)(CH2) (Tc1) and MnO2(CH3)(CH2) (Mn1). The calculations suggest rather complex scenarios with numerous pathways, where the initial compounds Re1-Mn1 may either engage in cycloaddition reactions or numerous addition reactions with concomitant hydrogen migration. There are also energetically low-lying rearrangements of the starting compounds to isomers which may react with ethylene yielding further products. The [2 + 2]Re,C cycloaddition reaction of the starting molecule Re1 is kinetically and thermodynamically favored over the [3 + 2]C,O and [3 + 2]O,O cycloadditions. However, the reaction which leads to the most stable product takes place with initial rearrangement to the dioxohydridometallacyclopropane isomer Re1a that adds ethylene with concomitant hydrogen migration yielding Re1a-1. The latter reaction has a slightly higher barrier than the [2 + 2]Re,C cycloaddition reaction. The direct [3 + 2]C,O cycloaddition becomes more favorable than the [2 + 2]M,C reaction for the starting compounds Tc1 and Mn1 of the lighter metals technetium and manganese but the calculations predict that other reactions are kinetically and thermodynamically more favorable than the cycloadditions. The reactions with the lowest activation barriers lead after rearrangement to the ethyl substituted dioxometallacyclopropanes Tc1a-1 and Mn1a-1. The manganese compound exhibits an even more complex reaction scenario than the technetium compounds. The thermodynamically most stable final product of ethylene addition to Mn1 is the ethoxy substituted metallacyclopropane Mn1a-2 which has, however, a high activation barrier.  相似文献   

11.
B3LYP/6-311 + G(d,p) has been used to calculate the relative energies and geometrical parameters of the respective reactants, transition states, and cycloadducts from the cycloadditions of azomethine ylide and ethene, (Z)-2-butene, and (E)-2-butene. The half-chair (envelope) transition state structures are consistent with a synchronous concerted cycloaddition mechanism.  相似文献   

12.
Whereas the cluster [Mo3S4(acac)3(py)3]+ ([ 1 ]+, acac=acetylacetonate, py=pyridine) reacts with a variety of alkynes, the cluster [W3S4(acac)3(py)3]+ ([ 2 ]+) remains unaffected under the same conditions. The reactions of cluster [ 1 ]+ show polyphasic kinetics, and in all cases clusters bearing a bridging dithiolene moiety are formed in the first step through the concerted [3+2] cycloaddition between the C?C atoms of the alkyne and a Mo(μ‐S)2 moiety of the cluster. A computational study has been conducted to analyze the effect of the metal on these concerted [3+2] cycloaddition reactions. The calculations suggest that the reactions of cluster [ 2 ]+ with alkynes feature ΔG values only slightly larger than its molybdenum analogue, however, the differences in the reaction free energies between both metal clusters and the same alkyne reach up to approximately 10 kcal mol?1, therefore indicating that the differences in the reactivity are essentially thermodynamic. The activation strain model (ASM) has been used to get more insights into the critical effect of the metal center in these cycloadditions, and the results reveal that the change in reactivity is entirely explained on the basis of the differences in the interaction energies Eint between the cluster and the alkyne. Further decomposition of the Eint values through the localized molecular orbital‐energy decomposition analysis (LMO‐EDA) indicates that substitution of the Mo atoms in cluster [ 1 ]+ by W induces changes in the electronic structure of the cluster that result in weaker intra‐ and inter‐fragment orbital interactions.  相似文献   

13.
Cyclic ketene N,X‐acetals 1 are electron‐rich dipolarophiles that undergo 1,3‐dipolar cycloaddition reactions with organic azides 2 ranging from alkyl to strongly electron‐deficient azides, e.g., picryl azide ( 2L ; R1=2,4,6‐(NO2)3C6H2) and sulfonyl azides 2M – O (R1=XSO2; cf. Scheme 1). Reactions of the latter with the most‐nucleophilic ketene N,N‐acetals 1A provided the first examples for two‐step HOMO(dipolarophile)–LUMO(1,3‐dipole)‐controlled 1,3‐dipolar cycloadditions via intermediate zwitterions 3 . To set the stage for an exploration of the frontier between concerted and two‐step 1,3‐dipolar cycloadditions of this type, we first describe the scope and limitations of concerted cycloadditions of 2 to 1 and delineate a number of zwitterions 3 . Alkyl azides 2A – C add exclusively to ketene N,N‐acetals that are derived from 1H‐tetrazole (see 1A ) and 1H‐imidazole (see 1B , C ), while almost all aryl azides yield cycloadducts 4 with the ketene N,X‐acetals (X=NR, O, S) employed, except for the case of extreme steric hindrance of the 1,3‐dipole (see 2E ; R1=2,4,6‐(tBu)3C6H2). The most electron‐deficient paradigm, 2L , affords zwitterions 16D , E in the reactions with 1A , while ketene N,O‐ and N,S‐acetals furnish products of unstable intermediate cycloadducts. By tuning the electronic and steric demands of aryl azides to those of ketene N,N‐acetals 1A , we discovered new borderlines between concerted and two‐step 1,3‐dipolar cycloadditions that involve similar pairs of dipoles and dipolarophiles: 4‐Nitrophenyl azide ( 2G ) and the 2,2‐dimethylpropylidene dipolarophile 1A (R, R=H, tBu) gave a cycloadduct 13 H , while 2‐nitrophenyl azide ( 2 H ) and the same dipolarophile afforded a zwitterion 16A . Isopropylidene dipolarophile 1A (R=Me) reacted with both 2G and 2 H to afford cycloadducts 13G , J ) but furnished a zwitterion 16B with 2,4‐dinitrophenyl azide ( 2I) . Likewise, 1A (R=Me) reacted with the isomeric encumbered nitrophenyl azides 2J and 2K to yield a cycloadduct 13L and a zwitterion 16C , respectively. These examples suggest that, in principle, a host of such borderlines exist which can be crossed by means of small structural variations of the reactants. Eventually, we use 15N‐NMR spectroscopy for the first time to characterize spirocyclic cycloadducts 10 – 14 and 17 (Table 6), and zwitterions 16 (Table 7).  相似文献   

14.
Thiobenzophenone and diazomethane afford thiadiazoline 1 at −78°. By elimination of N2 from 1 at −45° (t1/2 ca. 1 h), (diphenylmethylidenesulfonio)methanide ( 2 ), which cannot be isolated but is interceptible by dipolarophiles, is set free. The nucleophilic 1,3‐dipole 2 undergoes cycloadditions with electrophilic C,C multiple bonds; the structures of 16 cycloadducts were elucidated. One‐step and two‐step cycloaddition pathways are discussed in the light of the steric course observed for (E)/(Z)‐isomeric ethylene derivatives. Competition experiments with pairs of dipolarophiles at −45° and HPLC analysis of the adducts provided relative rate constants of 26 dipolarophiles, involving 2 C≡C, 13 C=C, 9 C=S, and 2 N=N bonds. In accordance with Sustmann`s reactivity model of concerted cycloadditions, 2 shows the highest selectivity of all known 1,3‐dipoles, i.e., the largest spread of rate constants (krel=1 for methyl propiolate and 33×106 for TCNE). As a consequence of low LU energies, thiones are very active dipolarophiles, and fluorene‐9‐thione (krel=79×106) stands at the top.  相似文献   

15.
Recent mass spectrometric studies of simple inorganic species (both charged and neutral) of main-group elements are reviewed, focusing attention on radicals and ions of interest to the chemistry of the atmosphere and its pollution. The examples illustrated concern the detection of HO3, hydrogen trioxide, the O2/O3 isotope exchange and its charged intermediate O5+, the reactions promoted by ionization of ozone/halocarbon mixtures in atmospheric gases, the ion chemistry of NOx oxides, and that of elemental chlorine and chlorine fluoride. Among the results of specific interest to gas-phase ion chemistry, the examples illustrated concern the intracluster ligand-switching reactions in ternary NO+ complexes, the NO2+ reactivity towards ethylene and acetylene, the gas-phase basicity of Cl2, the formation and characterization of Cl2X+ ions (X = Cl, F) and of [H3C-Cl-Cl]+, a new isomer of protonated dichloromethane.  相似文献   

16.
Oxirane chemical ionization (CI) gives numerous ions, including C2H3O+ and C2H5O+. These ions react with organic molecules through various specific ion–molecule reactions such as hydride abstraction, protonation, additions or cycloadditions. Oxirane CI allows discrimination between unsaturated compounds with [M + 43]+ and [M + 57]+ adduct ions and heteroatom functions with [M + 45]+ adduct ion. All are diagnostic ions. Oxirane CI permits selectivity during the ionization process of a mixture and discrimination of isomers.  相似文献   

17.
The participation of alkynylboronates in [4 + 2] cycloadditions has been investigated using both kinetic and DFT studies. Kinetic studies of the cycloaddition of tetrazine 1 with alkynylboronate 2 strongly suggest that a concerted cycloaddition mechanism is in operation. This mechanism has been confirmed by DFT calculations; moreover, a highly synchronous transition state appears to operate in this process. The experimentally observed poor reactivity of electron-rich dienes with alkynylboronates has also been confirmed by theoretical studies by analyzing the transition states of the cycloadditions with bis-2,5-trimethylsilyloxyfuran. The surprising conclusion has been made that alkynylboronates are relatively electron rich and have a cycloaddition reactivity that resembles that of acetylene. In contrast, the related dichloroalkynylborane cycloaddition reactivity resembles that of dimethylacetylene dicarboxylate.  相似文献   

18.
Ion-molecule reactions of free phenyl cations with diethylamine in the gas phase were studied radiochemically. The reaction practically totally follows the pathway of proton transfer, which occurs in the intermediate complex not only from the incoming cation (C6H5 +) but also from the ethyl substituent of the amine. Also, products of the reaction of diethylamine with benzyne (C6H4) generated by proton abstraction from the phenyl cation were detected.  相似文献   

19.
The reactions of the indenyl radicals with acetylene (C2H2) and vinylacetylene (C4H4) is studied in a hot chemical reactor coupled to synchrotron based vacuum ultraviolet ionization mass spectrometry. These experimental results are combined with theory to reveal that the resonantly stabilized and thermodynamically most stable 1-indenyl radical (C9H7.) is always formed in the pyrolysis of 1-, 2-, 6-, and 7-bromoindenes at 1500 K. The 1-indenyl radical reacts with acetylene yielding 1-ethynylindene plus atomic hydrogen, rather than adding a second acetylene molecule and leading to ring closure and formation of fluorene as observed in other reaction mechanisms such as the hydrogen abstraction acetylene addition or hydrogen abstraction vinylacetylene addition pathways. While this reaction mechanism is analogous to the bimolecular reaction between the phenyl radical (C6H5.) and acetylene forming phenylacetylene (C6H5CCH), the 1-indenyl+acetylene→1-ethynylindene+hydrogen reaction is highly endoergic (114 kJ mol−1) and slow, contrary to the exoergic (−38 kJ mol−1) and faster phenyl+acetylene→phenylacetylene+hydrogen reaction. In a similar manner, no ring closure leading to fluorene formation was observed in the reaction of 1-indenyl radical with vinylacetylene. These experimental results are explained through rate constant calculations based on theoretically derived potential energy surfaces.  相似文献   

20.
Lithium phenylacetylide ( 1a ) and N2O4 ( 2 ) at −78° yield diphenylbutadiyne ( 6a ) by oxidative coupling, phenylacetylene ( 7a ) by oxidation and then solvent H‐abstraction, and benzoyl cyanide ( 8 ) by dimerizative‐rearrangement of nitroso(phenyl)acetylene ( 23 ). Nitro(phenyl)acetylene ( 3 , R=Ph) is not obtained. Benzonitrile ( 9 ), a further product, possibly results from hydrolytic decomposition of nitroso(phenyl)ketene ( 27 ) generated from phenylacetylenyl nitrite ( 26 ). Phenylacetylene ( 7a ) and 2 give, along with (E)‐ and (Z)‐1,2‐dinitrostyrenes ( 34 and 35 , resp.), 3‐benzoyl‐5‐phenylisoxazole ( 10 ), presumably as formed by cycloaddition of benzoyl nitrile oxide ( 40 ) to 7a . Further, 2 reacts with other lithium acetylides ( 1b – 1e ), and with sodium, magnesium, zinc, copper, and copper lithium phenylacetylides, 1f – 1l , to yield diacetylenes 6a – 6c and monoacetylenes 7a – 7c . Conversions of metallo acetylide aggregates to diacetylenes are proposed to involve generation and addition reactions of metallo acetylide radical cationic intermediates in cage, further oxidation, and total loss of metal ion. Loss of metal ions from metallo acetylide radical cations and H‐abstraction by non‐caged acetylenyl radicals will give terminal acetylenes. The principal reactions (75–100%) of heavy metal acetylides phenyl(trimethylstannyl)acetylene ( 44 ) and bis(phenylacetylenyl)mercury ( 47 ) with 2 are directed nitrosative additions (NO+) and loss of metal ions to give nitroso(phenyl)ketene ( 27 ), which converts to benzoyl cyanide ( 8 ).  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