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 O2–N2 and O2–Ar negative-ion chemical ionization mass spectra of aromatic amines show a series of unusual ions dominated by an addition appearing at [M + 14]. Other ions are observed at [M – 12], [M + 5], [M + 12], [M + 28] and [M + 30]. Ion formation was studied using a quadrupole instrument equipped with a conventional chemical ionization source and a Fourier transform ion cyclotron resonance (FTICR) mass spectrometer. These studies, which included the examination of ion chromatograms, measurement of positive-ion chemical ionization mass spectra, variation of ion source temperature and pressure and experiments with 18O2, indicate that the [M + 14] ion is formed by the electron-capture ionization of analytes altered by surfaceassisted reactions involving oxygen. This conversion is also observed under low-pressure conditions following source pretreatment with O2. Experiments with [15N]aniline, [2,3,4,5,6-2H5] aniline and [13C6]aniline show that the [M + 14] ion corresponds to [M + O ? 2H], resulting from conversion of the amino group to a nitroso group. Additional ions in the spectra of aromatic amines also result from surface-assisted oxidation reactions, including oxidation of the amino group to a nitro group, oxidation and cleavage of the aromatic ring and, at higher analyte concentrations, intermolecular oxidation reactions.  相似文献   

2.
The structures of the major adduct ions formed in ammonia chemical ionization of thirteen aliphatic and aromatic ketones have been studied by mass analysed ion kinetic energy spectrometry. The [M+NH3+H]+ ion is shown to have a protonated carbinolamine structure, [M+2NH3+H]+ to be a protonated carbinolamine with hydrogen-bonded ammonia and [2M+NH3+H]+ to be, at least in part, a protonated carbinolamine with hydrogen-bonded ketone. These structures may imply a nucleophilic addition of ammonia at the carbonyl of the ketone-ammonium ion complex. An unusual hydroxy migration is seen in the internal rearrangement of the [2M+NH3+H]+ ion leading to the formation of a protonated imine.  相似文献   

3.
Secondary ion mass spectra of singly substituted aromatic hydrocarbon/H2SO4 solutions showed intense aromatic molecular ion and protonated aromatic molecule peaks characteristic of dissolved aromatic compounds from a number of aromatic compound classes, including acids, aldehydes, ketones, nitriles and nitrogen heterocycles. The presence of simultaneously abundant peaks for molecular ions and protonated molecules in secondary ion mass spectra of each aromatic compound/sulfuric acid solution is consistent with known or expected gas-phase proton transfer chemistry. The ratio of intensities, M+˙:[M + H]+, appears to be determined by sulfuric acid solution chemistry of the compound. Spectra obtained from 1–2 μl samples were relatively free from chemical noise and persisted for up to 20 min. Detection limits for some substituted aromatic compounds are estimated to be 10?12.  相似文献   

4.
Under positive ion chemical ionization conditions with ammonla at relatively low pressure, aromatic nitro compounds do not form [M + H]+ ions but often form ionic clusters [M + NH4]+ and [M + N2H7]+. Nitrobenzene forms a cluster [2M + NH4]+ and aniline, formed by nucleophilic substitution, leads to a cluster [anilinium ion + nitrobenzene]+. The dinitrobenzenes form [M + NH4]+ clusters and show evidence of nitroaniline formation and clustering. 1,3,5-Trinitrobenzene gives little indication of clustering or of substitution. The six isomers of trinitrotoluene appear to be stabilized by the methyl group and form clusters up to [M + N3H10]+. Nucleophilic substitution leads to dinitrotoluidines, which also form clusters with ammonium ions.  相似文献   

5.
The study of ion chemistry involving the NO2+ is currently the focus of considerable fundamental interest and is relevant in diverse fields ranging from mechanistic organic chemistry to atmospheric chemistry. A very intense source of NO2+ was generated by injecting the products from the dielectric barrier discharge of a nitrogen and oxygen mixture upstream into the drift tube of a proton transfer reaction time‐of‐flight mass spectrometry (PTR‐TOF‐MS) apparatus with H3O+ as the reagent ion. The NO2+ intensity is controllable and related to the dielectric barrier discharge operation conditions and ratio of oxygen to nitrogen. The purity of NO2+ can reach more than 99% after optimization. Using NO2+ as the chemical reagent ion, the gas‐phase reactions of NO2+ with 11 aromatic compounds were studied by PTR‐TOF‐MS. The reaction rate coefficients for these reactions were measured, and the product ions and their formation mechanisms were analyzed. All the samples reacted with NO2+ rapidly with reaction rate coefficients being close to the corresponding capture ones. In addition to electron transfer producing [M]+, oxygen ion transfer forming [MO]+, and 3‐body association forming [M·NO2]+, a new product ion [M−C]+ was also formed owing to the loss of C═O from [MO]+.This work not only developed a new chemical reagent ion NO2+ based on PTR‐MS but also provided significant interesting fundamental data on reactions involving aromatic compounds, which will probably broaden the applications of PTR‐MS to measure these compounds in the atmosphere in real time.  相似文献   

6.
The crystal structures of 13 compounds of the form M[Al2Me6X]·aromatic and related have been examined in order to learn about the M+...aromatic approach. Four types of interactions have been discerned: (1) metal...aromatic, (2) metal...aromatic...metal, (3) aromatic...metal...aromatic, and (4) no metal...aromatic contact. It was found that the closest K+...C(aromatic) and Cs+...C(aromatic) separations are essentially equal after a correction for the difference in metal radii. The strength of the K+...aromatic attraction was found to be sufficient to move the K+ ion 0.3 Å out of the plane of the crown ether in two complexes of dibenzo-18-crown-6.  相似文献   

7.
The fragment spectra of protonated nitro‐substituted benzodiazepines show an unusual fragment [M + H ‐ 14]+, which is shown by accurate mass measurement to be due to the loss of a nitrogen atom. Our investigations show that this apparent loss of atomic nitrogen is rather an attachment of molecular oxygen to the [M + H ‐ NO2]+? ion, which is the main fragment ion in these spectra. The oxygen attachment is exothermic, and rate constants have been derived. MSn spectra show that it is not easily reversible upon fragmentation of the adduct ion and that it is also observed with some secondary and tertiary fragments, which allows to limit the attachment site to the aromatic ring annulated to the diazepine moiety. Fragments of the oxygen adduct ion indicate that the O2 molecule dissociates in the adduct formation process, and the two oxygen atoms are bound to different sites of the ion. Comparison with radical cations generated by fragmentation of non‐nitro‐substituted benzodiazepines, none of which showed an oxygen attachment, and the fragmentation mechanisms involved in their formation indicates that the [M + H ‐ NO2]+? ion is a distonic ion with the charge and radical site neighbored on the aromatic ring. From these results, we derive a proposal for the formation and structure of the [M + H ‐ NO2 + O2]+? ion, which explains the experimental observations. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

8.
Unexpected [M + 15]+ ions were formed during the analysis of aromatic aldehydes by use of methanol in positive‐ion electrospray ionization mass spectrometry. Aromatic aldehydes with electron‐withdrawing groups or electron‐donating groups were all tested to make sure the universality. All the aromatic aldehydes studied with methanol as the solvent could generate [M + 15]+ ion, and for most of them, the [M + 15]+ ion was more intense than the [M + H]+ ion. Deuterium‐labeling experiment, high‐performance liquid chromatography–MS experiment, collision‐induced dissociation experiment, and theoretical calculations were performed to identify the formation of [M + 15]+ ion. The proposed reaction mechanism is a gas‐phase aldol reaction between protonated aromatic aldehydes and methanol occurring in electrospray source. Understanding and using this unique gas‐phase ion/molecule reaction can indeed offer a novel and fast approach for the direct identification of aromatic aldehydes. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

9.
Summary Negative ion mass spectra for 3 aliphatic and 4 aromatic isocyanates have been obtained by low pressure chemical ionization, using CH4, CO2 and N2O as reagent gases. All compounds furnished intense anions at m/z 42. With CH4, quasi-molecular anions were observed at m/z M+1 for aliphatic and m/z M+1 and M–1 for aromatic isocyanates. With N2O, anionic substitution products at m/z M+15 and M+30 were observed, and with CO2 and N2O, peaks at m/z M–12 could be detected for all aromatic isocyanates. Studies with 13CO2 and C18O2 as reagent gases showed that the anions at m/z M–12 and M+15 correspond to [M–CO+O] and [M–H+O], respectively.
Negativionen-Massenspektrometrie mit chemischer Ionisierung von einigen Isocyanaten
Zusammenfassung Die Negativionen-Massenspektren von 3 aliphatischen und 4 aromatischen Isocyanaten wurden mittels chemischer Ionisation bei tiefem Quellendruck aufgenommen, und zwar mit den Reagensgasen CH4, CO4 und N2O. Alle Verbindungen lieferten intensive Anionen mit m/z 42. Mit CH4 erhielten wir die quasi-molekularen Anionen M+1 für aliphatische sowie M+1 und M–1 für aromatische Isocyanate. Das Reagens N2O ergab die anionischen Substitutionsprodukte M+15 und M+30. Sowohl CO2 als auch N2O führten mit aromatischen Isocyanaten zur Bildung von M–12 Anionen. Versuche mit 13CO2 und mit C18O2 als Reagensgase zeigten, daß die Anionen M–12 und M+15 den Ionen [M–CO+O] und [M–H+O] entsprechen.
  相似文献   

10.
Low molecular weight polyisobutylenes (PIB) with chlorine, olefin and succinic acid end‐groups were studied using direct analysis in real time mass spectrometry (DART‐MS). To facilitate the adduct ion formation under DART conditions, NH4Cl as an auxiliary reagent was deposited onto the PIB surface. It was found that chlorinated adduct ions of olefin and chlorine telechelic PIBs, i.e. [M + Cl]? up to m/z 1100, and the deprotonated polyisobutylene succinic acid [M? H]? were formed as observed in the negative ion mode. In the positive ion mode formation of [M + NH4]+, adduct ions were detected. In the tandem mass (MS/MS) spectra of [M + Cl]?, product ions were absent, suggesting a simple dissociation of the precursor [M + Cl]? into a Cl? ion and a neutral M without fragmentation of the PIB backbones. However, structurally important product ions were produced from the corresponding [M + NH4]+ ions, allowing us to obtain valuable information on the arm‐length distributions of the PIBs containing aromatic initiator moiety. In addition, a model was developed to interpret the oligomer distributions and the number average molecular weights observed in DART‐MS for PIBs and other polymers of low molecular weight. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

11.
Mass spectra of 1-phenylethanol-1 and its analogues, specifically deuterated in the aliphatic chain, suggest that the [M? CH3]+ ion is represented partly by an α-hydroxybenzyl fragment. Moreover, the molecular ion loses successively—after scrambling of all hydrogen atoms, except those of CH3? a hydrogen atom and C6H6, generation the CH3CO+ ion. Diffuse peaks, found in the spectra of of 2-phenylethanol-1 and its analogues, specifically deuterated in the aliphatic chain and in the phenyl ring, show that the molecular ion loses C2H4O, possibly via a four-center mechanism, after an exchange of aromatic and hydroxylic hydrogens. Mass spectra of 1-phenylpropanol-2 and its analogues, specifically, deuterated in the aliphatic chain, demonstrate that in the molecular ion exclusively the hydroxyl hydrogen atom is transferred to one of the ortho-positions of the phenyl ring via a McLafferty rearrangement, generating the [M ? C2H4O]+ ion. Furtherore, an eight-membered ring structure is proposed for the [M ? CH3]+ ion to explain the loss of H2O and C2H2O from this ion after an extensive scrambling of hydrogen atoms.  相似文献   

12.
When using tetrachloromethane as the reagent gas in gas chromatography-ion trap mass spectrometry equipped with hybrid ionization source, the cation CCl3+ was generated in high abundance and further gas-phase experiments showed that such an electron-deficient reagent ion CCl3+ could undergo interesting ion–molecule reactions with various volatile organic compounds, which not only present some informative gas-phase reactions, but also facilitate qualitative analysis of diverse volatile compounds by providing unique mass spectral data that are characteristic of particular chemical structures. The ion–molecule reactions of the reagent ion CCl3+ with different types of compounds were studied, and results showed that such reactions could give rise to structurally diagnostic ions, such as [M + CCl3 – HCl]+ for aromatic hydrocarbons, [M – OH]+ for saturated cyclic ether, ketone, and alcoholic compounds, [M – H]+ ion for monoterpenes, M·+ for sesquiterpenes, [M – CH3CO]+ for esters, as well as the further fragment ions. The mechanisms of ion–molecule reactions of aromatic hydrocarbons, aliphatic ketones and alcoholic compounds with the reagent ion CCl3+ were investigated and proposed according to the information provided by MS/MS experiments and theoretical calculations. Then, this method was applied to study volatile organic compounds in Dendranthema indicum var. aromaticum and 20 compounds, including monoterpenes and their oxygen-containing derivatives, aromatic hydrocarbon and sesquiterpenes were identified using such ion–molecule reactions. This study offers a perspective and an alternative tool for the analysis and identification of various volatile compounds.  相似文献   

13.
Electron-impact studies of diazadiphosphetidines,[YF2PNMe]2(Y? F,Me, Ph, MeO,2,5-Me2C6H3, and m-CF3C6H4) are reported, the most abundant fragments corresponding to m/e [M/2–1]+, [M/2]+ and [M/2–1]+. It is concluded from metastable data that formation of the noval rearrangement ion, [M]+→[M/2+1]+is predominantly due to an electron-impact process. Variable temperature spectra of(F3PNMe)2, (i.e. for Y=F), suggest that ions of m/e [M/2-1]+are formed, in part, by a thermal process. For the compound [(m-CF3C6H4)F2PNMe]2 a well resolved negative ion spectrum has been obtained, with the molecular ion present in 100% abundance.  相似文献   

14.
A reinvestigation of the mechanism of formation of the [M – 1]+ ion in a series of N,N-dialkylbenzamides suggests that previous mechanisms put forward to account for the formation of the [M – 1]+ ion are deficient. A new mechanism is proposed which accounts for the data observed previously, as well as our results for a series of N,N-dialkyl-2-chlorobenzamides, 4-substituted N,N-dimethylbenzamides and some related compounds. For the N,N-dialkyl-2-chlorobenzamides, comparison of the abundances of the [M – 1]+ ion with the [M – 35]+ ion suggests that a concurrent reaction is occurring, besides loss of the ortho aromatic hydrogen atom. A study of substituent effects on the intensity ratio [M – 1]+/[M]+ shows an upward concave plot of this against σ+, suggesting that two competing mechanisms occur for the formation of the [M – 1]+ ion.  相似文献   

15.
The N2 negative ion chemical ionization (NICI) mass spectra of aniline, aminonaphthalenes, aminobiphenyls and aminoanthracenes show an unexpected addition appearing at [M + 11]. This addition is also observed in the N2 positive chemical ionization (PCI) mass spectra. An ion at [M – 15]? is found in the NICI spectra of aminoaromatics such as aniline, 1- and 2-aminonaphthalene and 1- and 2-aminoanthracene. Ion formation was studied using labeled reagents, variation of ion source pressure and temperature and examination of ion chromatograms. These experiments indicate that the [M + 11], [M – 15] and [M + 11] ions result from the ionization of analytes altered by surface-assisted reactions. Experiments with 15N2, [15N] aniline, [2,3,4,5,6-2H5] aniline and [13C6] aniline show that the [M + 11] ion corresponds to [M + N – 3H]. The added nitrogen originates from the N2 buffer gas and the addition occurs with loss of one ring and two amino group hydrogens. Fragmentation patterns in the N2 PCI mass spectrum of aniline suggest that the neutral product of the surface-assisted reaction is 1,4-dicyanobuta-1,3-diene. Experiments with diamino-substituted aromatics show analogous reactions resulting in the formation of [M – 4H] ions for aromatics with ortho-amino groups. Experiments with methylsubstituted aminoaromatics indicate that unsubstituted sites ortho to the amino group facilitate nitrogen addition, and that methyl groups provide additional sites for nitrogen addition.  相似文献   

16.
The loss of X· radical from [M + Cu + X]+ ions (copper reduction) has been studied by the so called in-source fragmentation at higher cone voltage (M = crown ether molecule, X = counter ion, ClO4, NO3, Cl). The loss of X· has been found to be affected by the presence/lack of aromatic ring poor/rich in electrons. Namely, the loss of X· occurs with lower efficiency for the [NO2-B15C5 + Cu + X]+ ions than for the [B15C5 + Cu + X]+ ions, where NO2-B15C5 = 3-nitro-benzo-15-crown-5, B15C5 = benzo-15-crown-5. A reasonable explanation is that Anion-π interactions prevent the loss of X· from the [NO2-B15C5 + Cu + X]+ ions. The presence of the electron-withdrawing NO2 group causes the aromatic ring to be poor in electrons and thus its enhances its interactions with anions. For the ion containing the aromatic ring enriched in electrons, namely [NH2-B15C5 + Cu + ClO4]+ where NH2-B15C5 = 3-amino-benzo-15-crown-5, the opposite situation has been observed. Because of Anion-π repulsion the loss of X· radical proceeds more readily for [NH2-B15C5 + Cu + X]+ than for [B15C5 + Cu + X]+. Iron reduction has also been found to be affected by Anion-π interactions. Namely, the loss of CH3O· radical from the ion [B15C5 + Fe + NO3 + CH3O]+ proceeds more readily than from [NO2B15C5 + Fe + NO3 + CH3O]+.  相似文献   

17.
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   

18.
The laser desorption mass spectrometry of the oxocarbon squaric acid (3,4-dihydroxy-3-cyclobutene-1,2-dione) and its salts of the form A2C4O4 (A = cation) is described. Both positive and negative ion spectra were obtained. The positive ion spectrum of the acid is characterized by an ion corresponding to loss of CO from [M + H]+. The negative ion spectrum shows an intense [M ? H]? peak in addition to a dimer species. The alkali salt spectra contain [M + A]+ in the positive mode and [M ? A]? and an intense [C4HO4]? in the negative mode. The smaller alkali salts also have an [M + H]+ adduct ion. Unlike the alkali squarates, the ammonium salt shows ions corresponding to losses of neutrals from the molecular adduct in the positive ion spectrum and a dimer species in the negative ion spectrum. Molecular weight information was obtained in all cases. A (bis) dicyanomethylene derivative of potassium squarate was also studied. Some field desorption mass spectrometry results are presented for comparison.  相似文献   

19.
Aminomonosaccharides (glucosamine, galactosamine, and mannosamine) in H2O and D2O were ionized by atmospheric pressure chemical ionization (APCI) and their fragmentation patterns were investigated to identify them. All the aminomonosaccharides showed the same fragment ions but their relative ion intensities were different. Major product ions generated in H2O were [M + H]+, [M + H – H2O]+, and [2M + H – 3H2O]+, while in D2O were [MD6 + D]+, [MD6 + D – D2O]+, and [2MD6 + D – D2O – 2HDO]+. At a high fragmentor voltage above 120 V, the relative ion intensities of the major product ions showed different trends according to the aminomonosaccharides. For the use of H2O as solvent and eluent, the order of the ion intensity ratio of [M + H – H2O]+/[2M + H – 3H2O]+ was galactosamine > mannosamine > glucosamine. When using D2O as solvent and eluent, the order of the ion intensity ratios of [MD6 + D – D2O]+/[MD6 + D]+ and [2MD6 + D – D2O – 2HDO]+/[MD6 + D]+ was mannosamine > galactosamine > glucosamine. It was found that glucosamine, galactosamine, and mannosamine could be distinguished by the specific trends of the major product ion ratios in H2O and D2O. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

20.
The metastable ions [M]2+, [M – H]2+· and [M – H2]2+ from malononitrile fragment by loss of [CH]+, [C]+· and [C]+·, respectively. The reaction of the molecular ion involves the methylene and nitrile carbon atoms in the statistical probability ratio, while that of [M – H]2+· involves exclusively the nitrile carbon and that of [M ? H2]2+ involves an approximately equal contribution, from both sources. It is suggested that the metastable molecular ion fragments through a bipyrimidal intermediate.  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