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.
Dissociative ionization of 1,2-epoxy n-alkanes gives rise to abundant [C4H7O]+ ions of structure [CH3OCHCHCH2]+. This conclusion is drawn from metastable ion analysis and from collisional activation spectra. This fragmentation involves the C? C ring opening and a 1,4-H migration leading to the corresponding enol ether [CH3OCHCHCH2R]+. precursor of [CH3OCHCHCH2]+ fragment. The same isomerization scheme applies to 1,2-epoxy methyl substituted alkanes and 2,3-epoxy n-alkanes.  相似文献   

2.
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   

3.
Gas-phase bimolecular and clustering reactions of acetonitrile in Xe, Kr, Ar, Ne and He were studied at high chemical ionization pressures in the new coaxial ion source at Auburn. With electron energies near the ionization threshold, the mass spectra are exceedingly simple and are comprised of [CH4CH]+ and clusters of [CH4CN]+ with various ligands such as H2O and CH3CN. At higher electron energies many other peaks appear. The intensities of the new peaks depend upon the ionization potential of the charge transfer gas, the ionizing electron energy and the ion source conditions, and are due to reactions of fragment ions. Residence time distributions at electron energies above the ionization threshold (∼ 30 eV) demonstrate that two molecular structures are present in the ion beam at m/z 42, one presumably is protonated acetonitrile ([CH3CNH]+) while the evidence indicates that the second species does not contain acidic hydrogens. With ionizing electron energies near threshold (∼ 10. 5 eV) only one structure is observed. Studies with electron energies near the ionization threshold under high-pressure chemical ionization conditions result in greatly simplified mass spectra and are possible only because of the coaxial geometry of the ion source.  相似文献   

4.
Due to neighbouring group effects the behaviour under electron impact of amino ethers of the type RO(CH2)nNR1R2 is distinctly different from that of both simple ethers and amines. In addition to α-cleavages two further primary fragmentations can often be observed, one leading to [M–aldehyde]+. or [M–ketone]+. ions, the other (with n≥3) giving [M–R]+ ions. Both processes involve hydrogen rearrangements which were investigated by means of deuterium labelling. Their importance is strongly dependent upon the nature of the substituents R, R1, R2 and upon the length (n) of the carbon chain between the heteroatoms. This dependence can easily be explained by the fragmentation mechanisms put forward.  相似文献   

5.
The chemistry of glycerol subjected to a high-energy particle beam was explored by studying the mass spectral fragmentation characteristics of gas-phase protonated glycerol and its oligomers by using tandem mass spectrometry. Both unimolecular metastable and collision-induced dissociation reactions were studied. Collision activation of protonated glycerol results in elimiation of H2O and CH3OH molecules. The resulting ions undergo further fragmentations. The origin of several fragment ions was established by obtaining their product and precursor ion spectra. Corresponding data for the deuterated analogs support those results. The structures of the fragment ions of compositions [C3H5O]+, [C2H5O]+, [C2H4O]+. and [C2H3O]+ derived from protonated glycerol were also identified. Proton-bound glycerol oligomers fragment principally via loss of neutral glycerol molecules. Dissociation of mixed clusters of glycerol and deuterated glycerol displays normal secondary isotope effects.  相似文献   

6.
(1) Thiocarboxylate anions [RCOS]? formed by dissociative secondary electron capture are either stable or fragment to yield [R]?. (2) Thiocarboxylate cations [RCOS]+ formed by charge stripping from [RCOS]?, fragment to form [R]+, [COS]+. and [RCO]+ (not when R=CF3). (3) Aryl hydrogen scrambling is observed in the case of the thiobenzoate cation. Aliphatic hydrogen scrambling is not detected for the thiopropionate cation.  相似文献   

7.
Abstract

Radialkationen der Zusammensetzung [H3PS]+ (m/z66) entstehen beim EI-Zerfall (70eV) von ionisiertem Diethylphosphinsulfid Et2HP(S). Aus der Analyse des Stoßaktivierungs(CA)-Massenspektrums der Ionen m/z 66 und aus thermochemischen Daten sowie den Ergebnissen semiempirischer MO-Rechnungen (MNDO) ergibt sich, daß bei ihnen die Struktur [H2P-SH]+. (Cs) gegenüber [H3PS]+. (C3v) bzw. [HPSH2]+. (Cs) bevorzugt ist. Mittels der jüngeren Technik der Neutralisations-Reionisations-Massenspektrometrie (NRMS) konnte gezeigt werden, daß Ionen der Struktur [H2PSH]+. durch Neutralisation in Neutralmoleküle H2P-SH übergeführt werden können, die in der verdünnten Gasphase stabil sind.  相似文献   

8.
The electron impact-induced fragmentation of 2,2-dimethyl- and 2-ethyl-1,3-indandione, 1 and 2, and their isomers, 3-isopropylidene- and 3-propylidenephthalide, 3 and 4, respectively, was studied in detail by mass-analysed ion kinetic energy (MIKE) and collision-induced dissociation (CID-MIKE) spectrometry, including 2H and 13C. labelled analogues of 1 and 2. In all regimes of internal energy, the molecular ions 1+. ? 4+. interconvert by up to seven consecutive, reversible isomerization steps prior to the main fragmentation processes, viz. loss of CH3. and C2H4. 1,3-Indandione and 3-methylenephthalide ions with identical alkylidene moieties (i.e. 1+.?3+. and 2+.?4+.) equilibrate rapidly and completely prior to fragmentation, whereas these pairs of isomers interconvert only slowly via a five-step rearrangement of the indandione ions 1+.?2+.. Distinct from the behaviour of simpler ionized carbonyl species, a 1,2-C shift of a (formally) neutral carbonyl group is found to occur along with that of a protonated one. Also distinct from simpler cases, methyl loss does not take place from the ionized enol intermediates formed within the interconversion 1+.?2+. of the diketone ions but rather from the n-propylidenephthalide ions 4+.. This follows from CID-MIKE spectrometry of the [M ? CH3]+ ions of 1–4 and two reference C10H7O2+ (m/z 159) ions of authentic structures (protonated 2-methylene-1,3-indandione and protonated 1,4-naphthoquinone). The characteristic CID fragmentation of the C10H7O2+ ions is rationalized. Finally, the multistep isomerization of ionized 1,3-indandiones apparently also extends to higher homologues [e.g. 5+. from 2-ethyl-2-methyl-1,3-indandione (5) and 6+. from 2,2-diethyl-1,3-indandione (6)]: the ionized phthaloyl group of 1,3-indandione radical cations 1+., 2+., 5+. and 6+., originally attached with its two acyl functionalities to the same carbon of the aliphatic chain, performs, in fact, a ‘multi-step migration’.  相似文献   

9.
The first trifluoromethylthiolation and [18F]trifluoromethylthiolation of alkyl electrophiles with in situ generated difluorocarbene in the presence of elemental sulfur and external (radioactive) fluoride ion is described. This transition‐metal‐free approach is high yielding, compatible with a variety of functional groups, and operated under mild reaction conditions. The conceptual advantage of this exogenous‐fluoride‐mediated transformation enables unprecedented syntheses of [18F]CF3S‐labeled molecules from most commonly used [18F]fluoride ions. The rapid radiochemical reaction time (≤1 min) and high functional‐group tolerance allow access to a variety of aliphatic [18F]CF3S compounds in high yields.  相似文献   

10.
For a number of hydrofluorocarbons (HFCs), EHez has been found to have a linear correlation with each of the following: (i) log (k/n); (ii) A/n; and (iii) Ea/R, where EH = HOMO energy of the molecule, z = average fractional positive charge on the abstractable hydrogen atom in the molecule, k = rate constant of the gas-phase H abstraction reaction of the molecule with OH radical at 298 K, n = number of abstractable H atoms in the molecule, A = preexponential factor, and Ea/R = activation temperature of the said reaction. These correlations have been used to estimate the temperature dependent rate constants for the reactions of OH radical with CF3CF2CH2CH2CF2CF3, CF3CH2CF2CH2CF3, CF3CF2CH2CH2F, CF3CH2CH3, CF3CH2CHF2, CF3CHFCH2F, and CHF2CHFCHF2 as {6.97 × 10−13 exp(1481/T)}, {5.43 × 10−13 exp(1754/T)}, {7.95 × 10−13 exp(l308/T)}, {8.0 × 10−13 exp(1300/T)}, {7.03 × 10−13 exp(1470/T)}, {7.33 × 10−13 exp(1417/T)}, and {8.09 × 10−13 exp(1285/T)}, respectively. These have not yet been measured experimentally. Linear correlation between EHez and log (k/n) has also been observed for nine halogen substituted acetaldehydes. On the other hand, EH is found to have a better linear correlation with log (k/n) than EHez in the case of fluorinated ethers and alcohols where the available experimental data are at present limited. © 1997 John Wiley & Sons, Inc. Int J Chem Kinet 29: 187–194, 1997.  相似文献   

11.
The vinylidene cation, [CH2C]+., was generated as a transient species by collisionally induced charge reversal of the corresponding anion. The collisionally induced mass spectrum of the transient [CH2C]+. is characteristically different from that of ionised acetylene.  相似文献   

12.
The structure of squamocin polyhydroxybis(tetrahydrofuran)acetogenin from Annona squamosa L., was characterized by a combination of the chemical derivatization and precursor-ion scanning mass spectrometry. The lactone part of squamocin was modified with N,N-dimethylethylenediamine in the vapour phase to afford a strong positive charge at one end of the skeleton. The derivative was ionized by fast atom bombardment (FAB), and the product ion spectrum from [M + H]+ and the precursor ion spectrum from the m/z 72 ion [CH2?CHN+H(CH3)2] generated by low-energy collision were acquired. The structure of the derivative could be characterized by the precursor-ion spetrum from the m/z 72 ion, but little structural information could be obtained from the production spectrum from [M + H]+ of the derivative. By the attachment of a tertiary amino group to one end of the skeleton, charge-remote fragmentation was facilitated and abundant and unique fragment ions were observed in the precursor-ion spectrum. Structural characterization of six other isolated compounds was also carried out, and two novel compounds were found by this technique. A similar derivatization was applied in the peptide leu-enkephalin, and full sequencing from the C-terminus using precursor-ion scanning from m/z 72 could be achieved.  相似文献   

13.
New examples of the ortho effect in bisphenol A derivatives including interaction of the hydrogen of the ortho-hydroxy group with the neighbouring aromatic ring have been observed. The characteristic ions [M ? PhOH]+middot; (m/z = 134) and [M ? CH3 ? PhOH]+ (m/z = 119) were shown to form through the hydrogen transfer from hydroxy and isopropyl groups, respectively. The spectra of cyclic derivatives having ortho-hydroxy functions show [M ? 43]+, [M ? C8H9O]+, m/z = 147, m/z = 135 and [M ? C9H10O]+ ions. The proposed mechanims of the corresponding transformations were supported by mass spectra of deuterated analogues, methyl and trimethyl silyl ethers.  相似文献   

14.
[CH3CHNH3]+. and [CH2CH2NH3]+. ions exist as distinct, stable species in the gas-phase. These ions are formed from a variety of precursors, and they can be characterized by their unimolecular and collision-induced reactions. The properties of deuterium labeled analogs confirm the proposed structures. Evidence that stable C3H9N+. ions with unconventional structures also exist is presented; these and other amine ion isomers are formed from inter alia, alkylamine molecular ions in the ion source.  相似文献   

15.
Relative rate experiments using UV photolysis of F2 or Cl2 have been used to determine rate constant ratios for several hydrofluorocarbon (HFC) reactions with Cl or F atoms and for HFC alkyl radicals with molecular halogens. For mixtures with F2 present, dark reactions are, also, observed which are attributed to thermal dissociation of the F2 to form F atoms. At 296 K, the rate of reaction (1a) [CF2HCH3 + F → CF2CH3 + HF] relative to (1b) [CF2HCH3 + F → CF2HCH2 + HF] is k1a/k1b = 0.73 (±0.13) and is independent of T (= 262–348 K). At 296 K, the ratio of reaction (2a) [CF2HCH2F + F → products] to that of (k1a + k1b) is (k1a + k1b)/k2a = 2.7 (±0.4), and for reaction (2b) [CF3CH3 + F → products] (k1a + k1b)/k2b = 22 ± 12. The temperature dependence (263–365 K) of the rate constant of reaction (3) [CF3CFH2 + Cl → products] relative to reaction (4) [CF3CFClH + Cl → products] is k3/k4(±10%) = 1.55 exp(?300 K/T). For the alkyl radicals formed from HFC 152a (CF2HCH2 and CF2CH3) and from HFC 134a (CF3CFH), rate constants for the reactions with F2 and Cl2 were measured relative to their reactions with O2. The rate constant of reaction (5cl) [CF2CH3 + Cl2 → CF2ClCH3 + Cl] relative to (5o) [CF2CH3 + O2 → CF2(O2)CH3] is k5cl/k5o(±15%) = 0.3 exp(200 K/T). For reaction (5f) [CF2CH3 + F2 → CF3CH3 + F], k5f/k5o(±35%) = 0.23. The ratio for reaction (6f) [CF2HCH2 + F2 → CF2HCH2F + F] relative to (6o) [CF2HCH2 + O2 → CF2HCH2O2] is k6f/k6o(±40%) = 1.23 exp(?730 K/T). The rate constant ratio for reaction (8cl) [CF3CFH + Cl2 → CF3CFClH + Cl] relative to reaction (8o) [CF3CFH + O2 → CF3CFHO2] is k8cl/k8o(±18%) = 0.16 exp(?940 K/T). For reaction (8f) [CF3CFH + F2 → CF3CF2H + F], k8f/k8o(±35%) = 0.6 exp(?860 K/T). © 1993 John Wiley & Sons, Inc.  相似文献   

16.
The bimolecular reaction of the CH2CHOH.+ enol ion (m/z 44) with acetaldehyde gives a strongly dominant product,m/z 45, formed mainly by proton transfer from the ion to the molecule. The abundance of the product coming from a H. abstraction reaction from the neutral, albeit more exothermic, is negligible. In order to explain this result, the long lived [CH2CHOH.+, CH3CHO] solvated ion was generated by reaction of the CH2CHOH.+ enol ion with (CH3CHO) n in the cell of a Fourier transform ion cyclotron resonance mass spectrometer. The structure of this solvated ion was clearly established. Labeling indicates that [CH2CHOH.+, CH3CHO], upon low energy collisions, reacts by H. abstraction more rapidly than by H+ transfer to the neutral moiety. This shows that the entropic factors are determinant when the enol ion reacts directly with acetaldehyde.  相似文献   

17.
Free, unsolvated phenylium ions formed by the spontaneous β decay of [1,4-3H2]benzene have been allowed to react with gaseous alcohols (ROH: R = Et, CF3CH2, Pr, or i-Pr; partial pressure: 3–56 Torr) and alkyl chlorides (R′Cl: R′ = Pr, i-Pr, or Bu; partial pressure: 20–450 Torr), in the presence of a thermal radical scavenger (O2: 4 Torr). Phenylium ion confirms its considerable site selectivity, demonstrated by the distinct preference toward the n-centre of the substrate (46–100%), although significant insertion into the alkyl group of alcohols is observed as well. Phenylium ion displays significant positional selectivity even between different n-type sites in a bidentate molecule such as CF3CH2OH. An affinity F < O < Cl trend is observed, which indicates a direct relationship between the polarizability of the n-centre of the molecule and its orienting properties toward phenylium ion. The stability features of the ionic intermediates from addition of phenylium ion with ROH or R′Cl have been evaluated, as well as their fragmentation and isomerization mechanisms. The behaviour of phenylium ion toward the selected substrates in the gas phase is discussed and compared with previous mechanistic hypotheses from related nuclear-decay studies.  相似文献   

18.
The behaviour under electron impact (70 eV) which includes some rearrangement processes of some tetraorganodiphosphanedisulfides R2P(S)-P(S)R2 (R ? CH3, C2H5, n-C3H7, n-C4H9, C3H5, C6H5) and CH3RP(S)–P(S)CH3R (R ? C2H5, n-C3H7, n-C4H9, C6H5, C6H5, C6H5,CH2) is reported and discussed. Fragmentation patterns which are consistent with direct analysis of daughter ions and defocusing metastable spectra are given. The atomic composition of many of the fragment ions was determined by precise mass measurements. In contrast to compounds R3P(S) loss of sulphur is not a common process here. The first step in the fragmentation of these compounds is cleavage of one P–C bond and loss of a substituent R?. The second step is elimination of RPS leading to [R2PS]+ from which the base peaks in nearly all the spectra arise. The phenyl substituted compounds give spectra with very abundant [(C6H5)3P]+. and [(C6H5)2CH3P]+. ions respectively, resulting from [M]+. by migration of C6H5. Rearrangement of [M]+. to a 4-membered P-S ring system prior to fragmentation is suggested.  相似文献   

19.
2-Fluoro-1,3-thiazoles were rapidly and efficiently labeled with no-carrier-added fluorine-18 (t1/2 = 109.7 min) by treatment of readily prepared 2-halo precursors with cyclotron-produced [18F]fluoride ion. The [18F]2-fluoro-1,3-thiazolyl moiety constitutes a new and easily-labeled structural motif for prospective molecular imaging radiotracers.  相似文献   

20.
The serotonin receptor 5-HT1A ligand 4-[18F]fluoro-N-[2-[1-(2-methoxyphenyl)-1-piperazinyl]ethyl-N-2-pyridinyl-benzamide (p-[18F]MPPF) was produced by a simplified method of Le Bars et al. Traditional oil bath heating was compared to microwave heating. Various radioanalytical methods, radio-Thin Layer Chromatography (TLC), High Pressure Liquid Chromatography (HPLC) and Mass Spectrometry (MS), were compared in the evaluation of the labelled product(s). The crude reaction mixture consisted of p-[18F]MPPF and 2–4 radioactive by-products eluting after the product fraction, and the reverse-phase HPLC method failed occasionally to separate p-[18F]MPPF from the radioactive by-product with close retention time. The heating method had no significant effect on the composition of labelled by-products. In LC-(ESI)-MS analysis of p-[18F]MPPF the labelled product was identified with m/z ratio of 435 ([M + H+]). The other HPLC fractions were measured to have following m/z ratios: (1) 327; 349; (675) (2) 402; 407/408; (791) and (3) 436, suggesting different kind of decomposition of the labelled product and/or the inactive precursor. The ion trap mass spectrometer was sufficient for the qualitative analysis of p-[18F]MPPF. However, differentiation of by-products arising from the decomposition of p-[18F]MPPF or from its precursor p-MPPNO2 proved to be challenging.  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