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 15 毫秒
1.
Fragmentation of 13 compounds of the 4H-pyran-4-one and pyridin-4-one series under electron impact involves formation of rearrangement ions stabilized by multiple bonds and oxygen atoms (mostly [RC≡O]+ and RCH=OR′]+), as well of neutral molecules with low enthalpies of formation (CO, H2O, CH2O, CO2, CH2=C=O, C3O2, and RCOOH; R = H, Me, HC≡C, HOC≡C).  相似文献   

2.
The mass spectrometric investigation of specifically deuterium and 13C labelled 2-trimethylsilyl-l-phenoxyethanes proves that the dissociative ionization of β-silyl-substituted ethane derivatives (loss of PhO?; p-CH3C6H4O?; and C4H?9 from PhOCH2CH2SiMe3, p-MeC6H4OCH2CH2SiMe3 and CH3CH2CH(CH3)CH2-CH2SiMe3, respectively) yields the non-classical bridge ethylene trimethylsilanium ion and not the open-chain isomer. Other stable C5H13Si+? ions, characterised by collisional activation mass spectrometry, are the dimethyl n-propyl silicenium ion and the l-trimethylsilyl ethyl cation, both generated from the molecular ions of CH3CH2CH2Si(Cl)Me2 and CH3CH(Cl)SiMe3 via unimolecular loss of Cl?.  相似文献   

3.
17O, 29Si, and 13C NMR spectra of more than 100 mono-, di-, tri- and tetra-alkoxysilanes R4−nSi(OR′)n; R = CnH2n+1, Ph, CH2Cl, CH2Br; R′ = CnH2n+1, CH2Ph, CH2CH2Cl, CH2CHCH2, CH2CCH, CH2CF3. (CH2)3Cl, (CH2)3CN have been studied.Linear relationships between the chemical shifts of 17O, 29Si, 13C in alkoxysilanes and the inductive and steric constants of substituents R and R′ were observed. Different transmission of electronic effects along the SiO bond in various directions was revealed by means of 13C, 29Si, 17O NMR spectroscopy and correlation analysis. The results are discussed in terms of (pd)π-bonding between the oxygen and silicon atoms in compounds containing an SiO bond.  相似文献   

4.
Specific ion/molecule reactions are demonstrated that distinguish the structures of the following isomeric organosilylenium ions: Si(CH3) 3 + and SiH(CH3)(C2H5)+; Si(CH3)2(C2H5)+ and SiH(C2H5) 2 + ; and Si(CH3)2(i?C3H7)+, Si(CH3)2(n?C3H7)+, Si(CH3)(C2H5) 2 + , and Si(CH3)3(π?C2H4)+. Both methanol and isotopically labeled ethene yield structure-specific reactions with these ions. Methanol reacts with alkylsilylenium ions by competitive elimination of a corresponding alkane or dehydrogenation and yields a methoxysilylenium ion. Isotopically labeled ethene reacts specifically with alkylsilylenium ions containing a two-carbon or larger alkyl substituent by displacement of the corresponding olefin and yields an ethylsilylenium ion. Methanol reactions were found to be efficient for all systems, whereas isotopically labeled ethene reaction efficiencies were quite variable, with dialkylsilylenium ions reacting rapidly and trialkylsilylenium ions reacting much more slowly. Mechanisms for these reactions and differences in the kinetics are discussed.  相似文献   

5.
G. Meyer  P. Viout 《Tetrahedron》1977,33(15):1959-1961
The alkaline hydrolysis of p-nitrophenyl acetate and of CH3CO2(CH2)2N+(CH3)2C16H33, Br? was studied in the presence of micelles C16H33N+(CH3)2CH2CH2OH, Br? and CTAB, C16H33N+(CH3)3,Br?. A pathway involving an intermediate is suggested for the hydrolysis of the ester. Hydrolysis rate of the intermediate in the presence of micelles is the same as hydrolysis rate for the ester in the absence of micelles. Consequently, hydrolysis of p-nitrophenyl acetate is not catalysed by one type of micelle, while it is enhanced by another type of micelle.  相似文献   

6.
The [C4H6O] ion of structure [CH2?CHCH?CHOH] (a) is generated by loss of C4H8 from ionized 6,6-dimethyl-2-cyclohexen-1-ol. The heat of formation ΔHf of [CH2?CHCH?CHOH] was estimated to be 736 kJ mol?1. The isomeric ion [CH2?C(OH)CH?CH2] (b) was shown to have ΔHf, ? 761 kJ mol?1, 54 kJ mol?1 less than that of its keto analogue [CH3COCH?CH2]. Ion [CH2?C(OH)CH?CH2] may be generated by loss of C2H4 from ionized hex-1-en-3-one or by loss of C4H8 from ionized 4,4-dimethyl-2-cyclohexen-1-ol. The [C4H6O] ion generated by loss of C2H4 from ionized 2-cyclohexen-1-ol was shown to consist of a mixture of the above enol ions by comparing the metastable ion and collisional activation mass spectra of [CH2?CHCH?CHOH] and [CH2?C(OH)CH?CH2] ions with that of the above daughter ion. It is further concluded that prior to their major fragmentations by loss of CH3˙ and CO, [CH2?CHCH?CHOH]+˙ and [CH2?C(OH)CH?CH2] do not rearrange to their keto counterparts. The metastable ion and collisional activation characteristics of the isomeric allenic [C4H6O] ion [CH2?C?CHCH2OH] are also reported.  相似文献   

7.
The activation reactions of methane mediated by metal carbide ions MC3+ (M = Ir and Pt) were comparatively studied at room temperature using the techniques of mass spectrometry in conjunction with theoretical calculations. MC3+ (M = Ir and Pt) ions reacted with CH4 at room temperature forming MC2H2+/C2H2 and MC4H2+/H2 as the major products for both systems. Besides that, PtC3+ could abstract a hydrogen atom from CH4 to generate PtC3H+/CH3, while IrC3+ could not. Quantum chemical calculations showed that the MC3+ (M = Ir and Pt) ions have a linear M-C-C-C structure. The first C–H activation took place on the Ir atom for IrC3+. The terminal carbon atom was the reactive site for the first C–H bond activation of PtC3+, which was beneficial to generate PtC3H+/CH3. The orbitals of the different metal influence the selection of the reactive sites for methane activation, which results in the different reaction channels. This study investigates the molecular-level mechanisms of the reactive sites of methane activation.  相似文献   

8.
The elimination of ethene from CH3CH2NH=CH 2 + is characterized by ab initio procedures. This reaction occurs through several asynchronous stages, but without passing through formal intermediates. A potential energy barrier to hydrogen migration from the β carbon to N is largely determined by the energy required to cleave the CN bond, but is lowered slightly by H transfer from the β to the α carbon and then to N. The complex [C2H 5 + NH=CH2] is bypassed, even though that complex could exist at energies only slightly above that of the transition state for ethene elimination. Furthermore, conversion of a substantial reverse activation energy into energy of motion causes CH2=NH 2 + and CH2=CH2 to dissociate faster than they can form [CH2=NH 2 + CH2=CH2]. Comparison of results for CH3CH2NH=CH 2 + to ab initio ones for methane from CH3CH2CH 3 + and elimination of ethene from CH3CH2O=CH 2 + and CH3CH2CH=OH+ reveals that these dissociations occur in a similar but, in each case, a distinct series of asynchronous steps or stages, and that there is no sharp demarcation between concerted and stepwise eliminations as presently defined. In dissociations of CH3CH2NH=CH 2 + , loss of electron density at the C in the breaking N bond leads the transfer of electron density to that carbon by migration of a hydrogen from the adjacent C. We attribute this to a requirement for the moving H to be close to Cα before the moving H can start to develop covalent bonding to Cα. It is also concluded that elimination of ethene from CH3CH2NH=CH 2 + avoids a Woodward-Hoffmann symmetry-imposed barrier by H migrating sufficiently from the β to the α carbon on the way to N, so that the dissociation is essentially a 1,1 rather than a 1,2 elimination.  相似文献   

9.
In a 150-V methane discharge, the rate of polymerization is approximately 3 times greater on the electrodes than on the walls. The sum of the C1 and C2 ions is 2 to 3 times higher in the dark space adjacent to the electrodes than in the space adjacent to the walls. Ethylene and acetylene are present in about equal amounts in all regions of the discharge. It is concluded that the ions C2H3 +, C2H2 +, CH3 +CH8 +, and CH+ have the greatest influence on the rate of polymerization, while the neutrals CH4, C2H2, C2H4, and C2H6 have the least influence.  相似文献   

10.
An account is given of the development of the proposal that ion–neutral complexes are involved in the unimolecular reactions of onium ions (R1R2C?Z+R3; Z = O, S, NR4; R1, R2, R3, R4 = H, CnH2n + 1), with particular emphasis on the informative C4H9O+ oxonium ion system (Z = O; R1, R2 = H; R3 = C3H7). Current ideas on the role of ion-neutral complexes in cation rearrangements, hydrogen transfer processes and more complex isomerizations are illustrated by considering the behaviour of isomeric CH3CH2CH2X+ and (CH3)2CHX+ species [X = CH2O, CH3CHO, H2O, CH3OH, NH3, NH2CH3, NH(CH3)2, CH2?NH, CH2?NCH3, CO, CH3˙, Br˙ and I˙]. Attention is focused on the importance of four energetic factors (the stabilization energy of the ion–neutral complex, the energy released by rearrangement of the cationic component, the enthalpy change for proton transfer between the partners of the ion neutral complex and the ergicity of recombination of the components) which influence the reactivity of the complexes. The nature and extent of the chemistry involving ion-neutral complexes depend on the relative magnitudes of these parameters. Thus, when the magnitude of the stabilization energy exceeds the energy released by cation rearrangement, the ergicity of proton transfer is small, and recombination of the components in a new way is energetically favourable, extensive complex-mediated isomerizations tend to occur. Loss of H2O from metastable CH2?O+C3H7 ions is an example of such a reaction. Conversely, if the stabilization energy is small compared with the magnitude of the energy released by eation rearrangement, the opportunities for complex-mediated processes to become manifest are decreased, especially if proton transfer is endoergic. Thus, CH3CH2CH2CO+ expels CO, with an increased kinetic energy release, after rate-limiting isomerization of CH3CH2CH2+? CO to (CH3)2CH+? CO has taken place. When proton transfer between the components of the complex is strongly exoergic, fragmentation corresponding to single hydrogen transfer occurs readily. The proton-transfer step is often preceded by cation rearrangement for CH3CH2CH2X+ species. In such circumstances, the involvement of ion–neutral complexes can be detected by the observation of unusual site selectivity in the hydrogen-transfer step. Thus, C3H6 loss from CH2?N+(R1)CH2CH2CH3 (R1 = H, CH3, C3H7) immonium ions is found by 2H-labelling experiments to proceed via preferential α-and γ-hydrogen transfer; this finding is explained if the incipient +CH2CH2CH3 ion isomerizes to CH3CH+CH3 prior to proton abstraction. In contrast, the isomeric CH2?N+(R1)CH(CH3)2 species undergo specific β-hydrogen transfer because the developing CH3CH+CH3 cation is stable with respect to rearrangements involving a 1,2-H shift.  相似文献   

11.
Fifteen [C6H5(X)FeCp]+ cations with substituents X having different electron-donating or electron-withdrawing effects were treated with NaBH4 in glyme or THF. The relative distributions of products from o-, m-, p- and ipso-additions of the hydride ion to the arene ring were determined by high resolution 1H NMR. For the η6-N,N-dimethylaniline-η5-cyclopentadienyliron cation with the most electron-donating of the substituents studied, only m- and p-hydride addition products were obtained, while in the reaction with the η6-nitrobenzene-η5-cyclopentadienyliron cation, which contained the most electron-withdrawing of the substituents investigated, only the o-addition product was formed. For the other 13 cations, with X  C6H5O, CH3O, p-CH3C6H4S, C6H5CH2, (CH3)3C, CH3, CH3CH2, C6H5, Cl, COOCH3, C6H5CO, CN and p-CH3C6H4SO2, o-, m- and p-hydride addition products were obtained in all cases, with a few instances also giving very minor amounts of ipso-adducts. The relative product distributions observed were interpreted by suggesting that while electronic effects played a major role, steric factors and free valency effects favoring o-addition as suggested by MO calculations [5] could also exert their influence in giving rise to the overall results.  相似文献   

12.
The goals of the present study were (a) to create positively charged organo‐uranyl complexes with general formula [UO2(R)]+ (eg, R═CH3 and CH2CH3) by decarboxylation of [UO2(O2C─R)]+ precursors and (b) to identify the pathways by which the complexes, if formed, dissociate by collisional activation or otherwise react when exposed to gas‐phase H2O. Collision‐induced dissociation (CID) of both [UO2(O2C─CH3)]+ and [UO2(O2C─CH2CH3)]+ causes H+ transfer and elimination of a ketene to leave [UO2(OH)]+. However, CID of the alkoxides [UO2(OCH2CH3)]+ and [UO2(OCH2CH2CH3)]+ produced [UO2(CH3)]+ and [UO2(CH2CH3)]+, respectively. Isolation of [UO2(CH3)]+ and [UO2(CH2CH3)]+ for reaction with H2O caused formation of [UO2(H2O)]+ by elimination of ·CH3 and ·CH2CH3: Hydrolysis was not observed. CID of the acrylate and benzoate versions of the complexes, [UO2(O2C─CH═CH2)]+ and [UO2(O2C─C6H5)]+, caused decarboxylation to leave [UO2(CH═CH2)]+ and [UO2(C6H5)]+, respectively. These organometallic species do react with H2O to produce [UO2(OH)]+, and loss of the respective radicals to leave [UO2(H2O)]+ was not detected. Density functional theory calculations suggest that formation of [UO2(OH)]+, rather than the hydrated UVO2+, cation is energetically favored regardless of the precursor ion. However, for the [UO2(CH3)]+ and [UO2(CH2CH3)]+ precursors, the transition state energy for proton transfer to generate [UO2(OH)]+ and the associated neutral alkanes is higher than the path involving direct elimination of the organic neutral to form [UO2(H2O)]+. The situation is reversed for the [UO2(CH═CH2)]+ and [UO2(C6H5)]+ precursors: The transition state for proton transfer is lower than the energy required for creation of [UO2(H2O)]+ by elimination of CH═CH2 or C6H5 radical.  相似文献   

13.
The principal fragmentation reactions of metastable [C3H7S]+ ions are loss of H2S and C2H4. These reactions and the preceding isomerizations of [C3H7S]+ ions with six different initial structures were studied by means of labelling with 13C or D. From the results it is concluded that the loss of H2S and C2H4 both occur at least mainly from ions with the structure [CH3CH2CH? SH]+ or from ions with the same carbon sulfur skeleton, with the exception of the ions with the initial structure [CH3CH2S? CH2]+, which partly lose C2H4 without a preceding isomerization. For all ions, more than one reaction route leads to [CH3CH2CH?SH]+. It is concluded that the loss of H2S is at least mainly a 1,3-elimination from the [CH3CH2CH?SH]+ ions. Both decomposition reactions are preceded by extensive but incomplete hydrogen exchange.  相似文献   

14.
The gas-phase clustering reactions of proton in propanol and acetone, and chloride ions in acetone were investigated. The −ΔHn−1,n values obtained for clustering reactions (n−1,n) were as follows: H+ (C3H7OH)n−1 + C3H7OH ⇄ H+ (C3H7OH)n, (2, 3) 18.9 kcal mol−1, (3, 4) 14.2 kcal mol−1, (4, 5) 11.7 kcal mol−1; H+ (CH3COCH3)2 + CH3COCH3 ⇄ H+ (CH3COCH3)3, 14.2 kcal mol−1; and Cl + CH3COCH3 ⇄ Cl (CH3COCH3), 12.4 kcal mol.−1. For clustering reactions, Cl (CH3COCH3n−1 + CH3COCH3 ⇄ Cl (CH3COCH3)n where n ≥ 2, the equilibria could not be established; probably due to the isomerization of ligand acetone molecules from the keto to enol form.  相似文献   

15.
The major metal-containing species formed upon fast atom bombardment of amino acid/Ni+2 mixtures is the [M + Ni]+ adduct, involving reduction of the Ni+2 to the +1 oxidation state. By contrast, electrospray ionization of amino acid/Ni+2 mixtures produces predominantly [Ni(M ? H)M]+; this species, on collisional activation, produces predominantly [M + Ni]+ by elimination of [M - H], presumably a carboxylate radical. The unimolecular fragmentation reactions occurring on the metastable ion time scale for the [M + Ni]+ adducts of a variety of α-amino acids have been recorded. The adducts with phenylalanine, α-aminoisobutyric acid and α-aminobutyric acid fragment by elimination of H2O, H2O + CO and, to a minor extent, by elimination of CO2. These reactions are similar to those observed for the [M + Cu]+ adducts of α-amino acids. A reaction distinctive for the [M + Ni]+ adducts involves formation of the immonium ion RCH=NH 2 + . By contrast, the [M + Ni]+ adducts with leucine, isoleucine, and norleucine show extensive metastable ion fragmentation by elimination of H2, CH4, C2H4, C3H6, and C4H8, with the relative importance of the different fragmentation channels depending on the configuration of the C4H9 side chain. These results are interpreted in terms of C-C and C-H bond activation of the C4H9 side chain by the Ni+. The adducts with valine and norvaline fragment in a fashion similar to the adduct with phenylalanine, except that minor elimination of C3H6 is observed.  相似文献   

16.
Mass spectra of substituted benchrotrenyls RC6H5Cr(CO)3 where R?H, F, CI, I, CH3, OCH3, COOCH3, C2H5, N(CH3)2, NH2, C6H5, C(CH3)3, p-C6H4NH2, CH2C6H5, CH2CH2C6H5), 1,3,5-(CH3)3C6H3Cr(CO)3 and 1,2,3,5-(CH3)4C6H2Cr(CO)3 have been studied. It has been found that for monosubstituted benchrotrenyls there is a linear dependence of the parameter log [Cr]+/[RC6H5Cr]+) on the number of degrees of freedom of the [RC6H5Cr]+ ion. Decarbonylation of the molecular ions is not affected by the nature of the substituent R. The results are interpreted in terms of the quasi-equilibrium theory of mass spectra.  相似文献   

17.
Summary Temperature-programmed desorption (TPD) of CH4, C2H6, C2H4, and CO and temperature-programmed pulse surface reactions (TPSR) of CH4, C2H6, C2H4, CO, and CO/H2 over a Co/MWNTs catalyst have been investigated. The TPD results indicated that CH4 and C2H6 mainly exist as physisorbed species on the Co/MWNTs catalyst surface, whilst C2H4 and CO exist as both physisorbed and chemisorbed species. The TPSR results indicated that CH4 and C2H6 do not undergo reaction between room temperature and 450oC. Pulsed C2H4 can be transformed into CH4 at 400 oC whilst pulsed CO can be transformed into CO2 at 100 or 150oC. In gaseous mixtures of CO and H2 containing excess CO, the products of pulsed reaction were CH3CHO and CH3OH. When the ratio of CO and H2 was 1:2, pulsed CO and H2 were transformed into CH3CHO, CH3OH and CH4. In H2 gas flow, pulsed CO was transformed into a mixture of CH3CHO and CH4 between 200 and 250oC and was transformed into CH4 only above 250oC.  相似文献   

18.
Differences between SiH+5 and CH+5 are more significant than the similarities. The proton affinity of SiH4 exceeds than of CH4 by ≈25 kcal/mol. but the heat of hydrogenation of SiH+3 is smaller than that of CH+3 by nearly the same amount. Like CH+5 the C5 structures of SiH+5 are preferred, but SiH+5 is best regarded as a weaker SiH+3—H2 complex. D3h, C2v, and C4v forms are much higher in energy and SiH+5 should not undergo hydrogen scrambling (pseudorotation) readily, as does CH+5 The neutral BH5 is only weakly bound toward loss H2, and the D3h. C2v, and C4v forms are also high in energy. The contral-atom electronegativities, C+ > B > Si+, control this behavior. The electronegativities also determine the ability to bear positive charges. Thermodynamically. SiH+5 and SiH+3 are more stable than CH+5 and CH+3, respectively; hydride transfer occurs from SiH4 to CH+3 and proton transfer from CH+5 to SiH4.  相似文献   

19.
Unstable 2-hydroxpropene was prepared by retro-Diels-Alder decomposition of 5-exo-methyl-5-norbornenol at 800°C/2 × 10?6 Torr. The ionization energy of 2-hydroxypropene was measured as 8.67±0.05 eV. Formation of [C2H3O]+ and [CH3]+ ions originating from different parts of the parent ion was examined by means of 13C and deuterium labelling. Threshold-energy [H2C?C(OH)? CH3] ions decompose to CH3CO++CH3˙ with appearance energy AE(CH3CO+) = 11.03 ± 0.03 eV. Higher energy ions also form CH2?C?OH+ + CH3 with appearance energy AE(CH2?C?OH+) = 12.2–12.3 eV. The fragmentation competes with hydrogen migration between C(1) and C(3) in the parent ion. [C2H3O]+ ions containing the original methyl group and [CH3]+ ions incorporating the former methylene and the hydroxyl hydrogen atom are formed preferentially, compared with their corresponding counterparts. This behaviour is due to rate-determining isomerization [H2C?C(OH)? CH3] →[CH3COCH3], followed by asymmetrical fragmentation of the latter ions. Effects of internal energy and isotope substitution are discussed.  相似文献   

20.
A mass spectrometer fast atom bombardment source has been used to synthesize, in the gas phase, the ion-molecule complexes of transition-metal ions (Ni+, CO+, Fe+, and Mn+) with α- or β-unsaturated alkenenitriles, RCH=CHCN (R=H, CH3, and C2H5) and CH3CH=CHCH2CN, and 2-methyl glutaronitrile. The metastable ion fragmentations of the complexes are monitored in the first held-free region by B/E linked scans. Surprisingly, an intense HCN loss via an intermediate (C n H2n ?2)?M+?(HCN) is observed for the complexes of the alkenenitriles. The metal ions significantly affect the fragmentation processes. The coexistence of both end-on and side-on coordination modes is suggested to explain the fragmentations.  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