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 results of high-pressure variable-temperature and variable ionizing electron energy studies of gas-phase ion-molecule reactions of dimethyl ether in krypton are presented. Near the ionization threshold a series of peaks corresponding to (CH3OCH3)nH+ (n = 1-4) clusters are observed. At higher ionizing electron energies, two new series of peaks appear, corresponding to [CH3OCH2]+(CH3OCH3)n and [(CH3)3O]+ (CH3OCH3)n clusters. The onium ion, [(CH3)3O]+, has been previously reported at elevated temperatures under methane chemical ionization conditions. It was suggested that the onium ion is formed by reaction of (CH3)2OH+ with CH3OCH3 with subsequent elimination of methanel, i.e. by fragmentation of an adduct ion. The present results strongly suggest that, under our conditions, [CH3OCH2]+ rather than thermal (CH3)3OH+, is the precursor to [(CH3)3O]+.  相似文献   

2.
The positive ion–molecule reactions of OCS have been investigated in an ion cyclotron resonance spectrometer. A variety of reactions in OCS/hydrocarbon mixtures have been investigated for various C1? C4 hydrocarbons—alkanes, alkenes and alkynes. The formation of organosulfur ions is found in reactions in OCS/hydrocarbon (Cn) mixtures with n <4. Formation of organosulfur ions is observed from hydrocarbon ions reacting with OCS and [OCS]+˙ and S+˙ reacting with the hydrocarbons. The proton affinity of OCS has been determined to be 688.7±8 kJ mol?1 while that of CS2 is measured to be 712.1±8 kJ mol?1. Comparison with the proton affinity of CO2 shows that the proton affinity increases as sulfur is substituted for oxygen.  相似文献   

3.
The reactions of the cyclic molecules C6H6 (benzene), c-C3H6 (cyclopropane) and c-C6H12 (cyclohexane) with ArH+ (ArD+), H3+, N2H+, CH5+, HCO+, OCSH+, C2H3+, CS2H+ and H3O+ have been studied at 300 K using a SIFT apparatus. All the reactions except those of C2H3+ proceed via proton transfer and all are fast except the H3O+ and CS2H+ reactions with c-C6H12 which are endothermic and which establish that the proton affinity of c-C6H12 is 160 ± 1 kcal mol−1, which is considerably lower than the published value. In the c-C3H6 and the c-C6H12 reactions multiple products are observed and hence “breakdown curves” for the protonated molecules are constructed and the appearance energies of the various ion products are consistent with available thermochemical data. The reactions of C2H3+ with these cyclic molecules are atypical within this series of reactions in that they appear to proceed largely via hydride ion transfer. The implications of the results of this study to interstellar chemistry are alluded to.  相似文献   

4.
State-of-the-art ab initio studies demonstrate that the reaction Pd+ + CH3I → PdCH2I+ + H. is endothermic by ca. 20 kcal/mol, which translates into a bond dissociation energy (BDE) of ca. 83 kcal/mol for the Pd+? CH2I bond. This figure is in agreement with an experimental bracket of 68 kcal/mol < BDE(Pd+? CH2I) < 92 kcal/mol. Based on these findings, the previously studied Pd+/CH3I system was re-investigated, and double-resonance experiments demonstrate that the formation of PdCH2I+ occurs stepwise via PdCH as a reactive intermediate. Further, ion/molecule reactions of PdCH2I+ with unsaturated hydrocarbons are studied, which reveal the formation of carbon–carbon bonds in the gas phase.  相似文献   

5.
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   

6.
Carnitine inner salt, (CH3)3N+ CH2CH(OH)CH2COO?, and carnitine hydrochloride, (CH3)3N+CH2CH (OH)CH2COOH Cl?, in the solid state undergo ion-beam-induced intermolecular methyl transfer reactions as shown by (CH3)3N+ CH2CH(OH)CH2COOCH3 ions at m/z 176 in their positive ion spectra. In the case of carnitine HCl, the product ion is three times as abundant as the intact cation. For the inner salt however, the product is less than one-tenth as abundant as [M + H] +. In both cases, the reaction can be precluded by dissolution of the sample, supporting an intermolecular mechanism. The negative ion spectra for these compounds contain no [M ? CH3]? ions, suggesting that simple transmethylation does not occur. Rather it is proposed that the inner salt abstracts a methyl group from the intact carnitine cation to yield [M + CH3]+ and a neutral species, the driving force being a minimization of the total number of charges desorbed into the gas phase. Thermodynamic data favor this mechanism as do data for other carnitine salts. The reaction appears to be inhibited when one reactant is present in excess. This is the case for carnitine HNO3 and CH3SO3H, which tend to liberate the intact cation since the anions are large and polarizable. It is also the case for small, hard anions like fluoride, which appear to favor release of the inner salt, hence the cation at m/z 162 is of low abundance and the transmethylation product (m/z 176) is absent. The extent of the reaction is also dependent on the methods of preparation of the sample, and deposition of the salts from solution greatly reduces the extent of methyl transfer. [M ? CH3]? is observed when glycerol is used as a matrix, possibly due to a matrix-analyte methyl transfer reaction.  相似文献   

7.
The methoxy cation, CH30+, formed by collision-induced charge reversal of methoxr anions with a kinetic energy of 8 keY, has been differentiated from the isomenric CH2OH+ ion by performing low kinetic energy ion-molecule reactions In the radiofrequency-only quadrupole of a reverse-geometry double-focusing quadrupole hybrid mass spectrometer. The methoxy cation reacts with CH3SH, CH3?CH=CH2, (CH3)2O, and CH3CH2Cl by electron transfer, whereas the CH2OH+ ion reacts by proton transfer with these substrates  相似文献   

8.
Product distributions and rate constants for the reaction of ground state C+ ions with O2, NO, HCl, CO2, H2S, H2O, HCN, NH3, CH4, H2CO, CH3OH, and CH3NH2 have been measured. Rate constants were obtained using ion cyclotron resonance trapped ion methods at JPL, and product distributions were obtained using a tandem (Dempster-ICR) mass spectrometer at the University of Utah. Rapid carbon isotope exchange has also been observed in C+-CO collisions.  相似文献   

9.
The reactions of CS(X 1Σ+), CS2(X 1Σ+g) and OCS(X 1Σ+) with O(3P) were studied at 298 K by means of a CO laser resonance absorption technique. The CO(ν) population distribution produced from the reaction O(3P) + CS(X 1Σ+) studied in a quartz flash photolysis tube (λ>/ 200 nm) is similar to distributions observed previously for ν> 7. For ν < 7 an energetically colder vibrational population was observed which is attributed to the reaction of O(3P) atoms with undissociated CS2(X 1Σ+g). Subsequent experiments carried out in a Pyrex flash photolysis tube (λ>/ 300 nm) in which the O(3P) + CS2(X 1Σ+g) reaction is the only one which can occur confirmed that the colder population observed is attributable to this process. The branching ratio for the reaction channel O(3P) + CS2(X 1Σ+g) → CO(X 1Σ+) + S2(3Σ?g) has been measured. We find that 1.4 ± 0.2% of the O + CS2 reaction proceeds through this channel, and that the rate constant for this reaction channel is, k = 3.5 (±0.5) × 1010 cm3/mole s. Isotope labeled experiments using 18O atoms show that the O(3P) + OCS(X 1Σ+) reaction takes place by a direct stripping mechanism, wherein CO(ν) is produced exclusively from the parent OCS molecule. The CO(ν) formed in this reaction carries about 9% of the total available energy.  相似文献   

10.
Fourier transform ion cyclotron resonance mass spectrometry has been used to measure the reaction rates for ions derived from methylamine with dimethylamine or trimethylamine. The use of the selective ion ejection technique greatly simplifies the elucidation of the ion-molecule reaction channels. The rate constants for proton transfer from protonated metwlamine, CH3NH 3 + (m/z 32), to dimethylamine and trimethylamine are 16.1 ± 1.6 × 10?10 and 9.3 ± 0.9 × 10?10 cm3 molec?1s?1, respectively. The rate constants for charge transfer from methylamine molecular ion, CH3NH 2 + (m/z 31), to dimethylamine and trimethylamine are 9.3 ± 1.8 x 10?10 and 15.0 ± 5 × 10?10 cm3molec?1s?1, respectively.  相似文献   

11.
The mass spectra of the methyl-, trideuteromethyl-, ethyl- and pentadeuteroethylethers of 2,2′-bis-trimethylsilylbenzhydrol are reported. The most significant ions arise from the [M – CH3]+ ion, formed by loss of a methyl radical from one of the trimethylsilyl groups. After ring formation by interaction of the siliconium ion centre with an aromatic nucleus, the ion loses (CH3)3Si? OR (R = CH3, C2H5, CD3 and C2D5), giving ion m/e 223. The fragment (CH3)3Si? OCH3 is also eliminated in the four ethers investigated from the ion [M – R]+. Attack of the siliconium ion. Indications are found for a transannular hydrogen/deuterium rearrangement and a transannular elimination reaction. The intensity of some peaks in the spectra are discussed in relation to group R.  相似文献   

12.
This study compared the conversion of two malodorous substances, dimethyl sulfide (CH3SCH3, DMS) and methanethiol (CH3SH) in a cold plasma reactor. The DMS and CH3SH were successfully destroyed at room temperature. DMS decomposed less than CH3SH at the same conditions. In oxygen-free condition, CS2 and hydrocarbons were the major products, while SO2 and COx were main compounds in oxygen-rich environments. The DMS/Ar plasma yielded more hydrocarbons and less CS2 than that of CH3SH/Ar plasma. In the CH3SH/O2/Ar plasma, rapid formation of SO and CO resulted in the yields much more amounts of SO2 and CO2 than those in the DMS/O2/Ar plasma; and remained only a trace of total hydrocarbons, CH2O, CH3OH, CS2, and OCS. The major differences between the reaction mechanisms of DMS and CH3SH were also proposed and discussed.  相似文献   

13.
A very recent laser ablation‐molecular beam experiment shows that an Al+ ion can react with a single methylamine (MA, CH3NH2) or dimethylamine (DMA, (CH3)2NH) molecule to form a 1:1 ion–molecule complex Al+[CH3NH2] or Al+[(CH3)2NH)], whereas a dehydrogenated complex ion Cu+[CH3N] or Cu+[C2H5N] is detected, respectively, in the similar reaction for a Cu+ ion. Here, we show a comparative density functional theory study for the reactivities of the Al+ and Cu+ ions toward MA and DMA to reveal the intrinsic mechanism. It is found that the interactions of the Al+ ion with MA and DMA are mostly electrostatic, leading to the direct ion–molecule complexes, Al+? NH2CH3 and Al+? NH( CH3)2, in contrast to the non‐negligible covalent character in the corresponding Cu+‐containing complexes, Cu+? NH2CH3 and Cu+? NH( CH3)2. The general dehydrogenation mechanism for MA and DMA promoted by the Cu+ ion has been shown, and the preponderant structures contributing to the mass spectra of the product ions Cu+[CH3N] and Cu+[C2H5N] are rationalized as Cu+? NHCH2 and Cu+? N( CH2)( CH3). The presumed dehydrogenation reactions are also discussed for the Al+‐containing systems. However, the involved barriers are found to be too high to be overcome at low energy conditions. These results have rationalized all the experimental observations well. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

14.
Gas-phase clustering reactions of CoCp+ with H2 and with CH4 were investigated using temperature-dependent equilibrium experiments. In both systems, the CoCp+ ion was found to form strong interactions with two ligands. The first and second H2 groups cluster to CoCp+ with bond energies of 16.2 and 16.8 kcal/mol, respectively, while the first and second CH4 groups cluster to CoCp+ with bond energies of 24.1 and 12.1 kcal/mol, respectively. These bond energies are in good agreement with those determined by density functional theory (DFT). Molecular geometries for the four clusters determined with DFT are also presented. Weak experimental bond energies of 0.9 kcal/mol for the third H2 and 2.2 kcal/mol for the third CH4 clustering to CoCp+ suggest these ligands occupy the second solvation shell of the ion. In addition to clustering in the methane system, H2-elimination from CoCp(CH4)2+ was observed. The mechanism for this reaction was investigated by collision-induced dissociation experiments and DFT, which suggest the predominate H2-elimination product is (c-C5H6)Co+---C2H5. Theory indicates that dehydrogenation requires the active participation of the Cp ring in the mechanism. Transfer of H and CH3 groups to the C5-ring ligand allows the metal center to avoid the high-energy Co(IV) oxidation state required when it forms two covalent bonds in addition to its interaction with a C5-ring ligand.  相似文献   

15.
Reactions that proceed within mixed ethylene–methanol cluster ions were studied using an electron impact time-of-flight mass spectrometer. The ion abundance ratio, [(C2H4)n(CH3OH)mH+]/[(C2H4)n(CH3OH)m+], shows a propensity to increase as the ethylene/methanol mixing ratio increases, indicating that the proton is preferentially bound to a methanol molecule in the heterocluster ions. The results from isotope-labelling experiments indicate that the effective formation of a protonated heterocluster is responsible for ethylene molecules in the clusters. The observed (C2H4)n(CH3OH)m+ and (C2H4)n(CH3OH)m–1CH3O+ ions are interpreted as a consequence of the ion–neutral complex and intracluster ion–molecule reaction, respectively. Experimental evidence for the stable configurations of heterocluster species is found from the distinct abundance distributions of these ions and also from the observation of fragment peaks in the mass spectra. Investigations on the relative cluster ion distribution under various conditions suggest that (C2H4)n(CH3OH)mH+ ions with n + m ≤ 3 have particularly stable structures. The result is understood on the basis of ion–molecule condensation reactions, leading to the formation of fragment ions, $ {\rm CH}_2=\!=\mathop {\rm O}\limits^ + {\rm CH}_3 $ and (CH3OH)H3O+, and the effective stabilization by a polar molecule. The reaction energies of proposed mechanisms are presented for (C2H4)n(CH3OH)mH+(n + m ≤ 3) using semi-empirical molecular orbital calculations.  相似文献   

16.
A method is introduced by which mass-analysed ion kinetic energy spectra free from Z-discrimination can be obtained for both collisionally activated (CA) and metastable decomposition reactions. The method, performed on a ZAB-E instrument fitted with a collision cell, but applicable also to the ZAB-2F, involves summation of the ‘height resolved’ contributions (formed by beam collimation in the Z-axis and selected by electrostatic deflection of the incident beam) using the signal averaging facility normally available. Representative results (at 8 or 10 keV energy) are given for the CA (Ar target) reactions [CS2]2+ → [CS]+; [CS2]+ → S+ and [CH3OH]+ → [m/z = 12–31]+, and for the metastable reaction [m/z 45]+ → [m/z 29]+ in ethanol.  相似文献   

17.
Cross section measurements for the proton transfer reactions of NH+4, CH3NH+3, and PH+4 with Ca(g) have been obtained over a range of low ion kinetic energies. For all reactions studied the cross sections drop sharply with increase in ion kinetic energy, indicating exothermic behavior. The results show that Ca(g) is an unusually strong base with a proton affinity in excess of 9.2 eV. Cross sections for the PH+4Ca reaction are an order to magnitude higher than those for the NH+4Ca reaction for ion energies between one and three eV. This effect is not explained by simple theories of ion-induced dipole interactions. It is suggested that the enhanced rate of the PH+4Ca reaction may be due to d-orbital participation.  相似文献   

18.
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   

19.
The ion-trap mass spectrometer has several features which make it a useful device for the study of ion/molecule reactions, viz., the ability to store ions for long periods, mass-selective storage, access to time and pressure-resolved data, and MS/MS capabilities in which the fragmentation behavior of selected ions may give insight into ion structure. These capabilities are used to study the gas-phase halomethylation of a variety of organic compounds with CH2Cl+ as the reagent ion. The ion/molecule reaction of greatest interest involves addition of CH2Cl+, followed by the elimination of HCl, resulting in a net addition of methyne. This methyne-addition reaction is observed in many aromatic compounds as well as such compounds as cycloheptatriene and cyclo-octene. The structures of the product ions were probed using collision-activated dissociation.  相似文献   

20.
《Chemical physics》2001,263(2-3):449-457
Photoelectron–photoion coincidence spectroscopy has been used to examine dissociative ionisation of CS2 from electronic states of CS2+ up to 27 eV, including the satellite states 3, 4, 6 and 10 whose decay has not been studied before. Branching ratios to the ions S+, CS+, S2+ and C+ have been determined throughout the range and kinetic energy release distributions have been deduced from peak shapes, allowing inferences on the states of the fragments. The choice of product channel is not strongly dependent on initial parent ion state identity. The products are formed in many different final states, but kinetic energy releases less than 3 eV are favoured, corresponding to formation of highly excited states of the products. In confirmation, optical emission has been found in coincidence with photoelectrons from formation of several inner valence states of the ions. Formation of S2+ occurs from several initial states of the parent ion and possible mechanisms are considered. It is concluded that a “quasi-statistical” model may best describe the dissociation of CS2+ from the inner valence states.  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