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 421 毫秒
1.
Methyl 2-oxocycIoalkane carboxylate structures are proposed lor the [M ? MeOH] ions from dimethyl adipate, pimelate, suberate and azelate. This proposal is based on a comparison of the metastable ion mass spectra and the kinetic energy releases for the major fragmentation reaction of these species with the same data for the molecular ions of authentic cyclic β-keto esters. The mass spectra of α,α,α′,α′-d4-pimelic acid and its dimethyl ester indicate that the α-hydrogens are involved only to a minor extent in the formation of [M ? ROH] and [M ? 2ROH] ions, while these α-hydrogens are involved almost exclusively in the loss of ROH from the [M ? RO˙]+ ions (R = H or CH3). The molecules XCO(CH2)7COOMe (X = OH, Cl) form abundant ions in their mass spectra with the same structure as the [M ? 2MeOH] ions from dimethyl azelate.  相似文献   

2.
The behavior of some substituted cyclopentadienylmanganese ions has been studied by tandem mass spectrometry. This metastable ion study showed that only C5H5Mn+ and (C5H4CN)Mn+ ions retain their nido-cluster structure (1), which is characterized by a simple metal-ligand bond cleavage. Other substituted ions, RXC5H4Mn+, rearrange to a different extent, depending on the nature of the substituent. The first rearrangement step is R radical migration to the central metal atom, leading to RMnC5H4X+-type ions (2). These ions decompose by elimination of X (for X=CO) or with formation of RMnX+, but further rearrangements can also occur. These are the reverse migration of R from the metal atom to the π-ligand (for R=H, Ph) and cyclopentadienyl ring expansion (for X=CH2). Collisional activation mass spectra contained an Mn+ ion peak, which can indicate the existence of stable type 1 structures for most cyclopentadienylmanganese ions. Carboxyl and hydroxymethyl derivatives exist, presumably as ions of type 2. The neutralization-reionization mass spectra of RXC5H4Mn+ ions are also discussed.  相似文献   

3.
The mass spectra of 30 sulfinamide derivatives (RSONHR', R' alkyl or p-XC6H4) are reported. Most of the spectra had peaks attributable to thermal decomposition products. For some compounds these were identified by pyrolysis under similar conditions to be: RSO2NHR', RSO2SR, RSSR and NH2R' (in all kinds of sulfinyl amides); RSNHR' (in the case of arylsulfinyl arylamides); RSO2C6H4NH2, RSOC6H4NH2 and RSC6H4NH2 (in the case of arylsulfinyl arylamides of the type of X = H) The mass spectra of the three thermally stable compounds showed that there are several kinds of common fragment ions. The mass spectra of the thermally labile compounds had two groups of ions; (i) characteristic fragment ions of the intact molecules and (ii) the molecular ions of the thermal decomposition products. It was concluded that the sulfinamides give the following ions after electron impact: [M]+, [M ? R]+, [M ? R + H]+, [M ? SO]+, [RS]+, [NHR']+, [NHR' + H]+, [RSO]+, [RSO + H]+, [R]+, [R + H]+, [R']+ and [M ? OH]+, and that the thermal decomposition products give the following ions: [RSO2SR]+, [RSSR]+, [M ? O]+, [M + O]+ and [RSOC6H4NH2]+.  相似文献   

4.
Organophosphate esters (OPEs) are chemical compounds incorporated into materials as flame‐proof and/or plasticizing agents. In this work, 13 non‐halogenated and 5 halogenated OPEs were studied. Their mass spectra were interpreted and compared in terms of fragmentation patterns and dominant ions via various ionization techniques [electron ionization (EI) and chemical ionization (CI) under vacuum and corona discharge atmospheric pressure chemical ionization (APCI)] on gas chromatography coupled to mass spectrometry (GC‐MS). The novelty of this paper relies on the investigation of APCI technique for the analysis of OPEs via favored protonation mechanism, where the mass spectra were mostly dominated by the quasi‐molecular ion [M + H]+. The EI mass spectra were dominated by ions such as [H4PO4]+, [M–R]+, [M–Cl]+, and [M–Br]+, and for some non‐halogenated aryl OPEs, [M]+● was also observed. The CI mass spectra in positive mode were dominated by [M + H]+ and sometimes by [M–R]+, while in negative mode, [M–R] and more particularly [X] and [X2]‐● were mainly observed for the halogenated OPEs. Both EI and APCI techniques showed promising results for further development of instrumental method operating in selective reaction monitoring mode. Instrumental detection limits by using APCI mode were 2.5 to 25 times lower than using EI mode for the non‐brominated OPEs, while they were determined at 50‐100 times lower by the APCI mode than by the EI mode, for the two brominated OPEs. The method was applied to fish samples, and monitored transitions by using APCI mode showed higher specificity but lower stability compared with EI mode. The sensitivity in terms of signal‐to‐noise ratio varying from one compound to another. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

5.
Homoadamantane derivatives can be divided into two groups according to their mass spectra. To the first group belong compounds with electron attracting substituents (COOH, CI, COOCH3, Br); compounds with electron releasing substituents (OCH3, OH, NH3, NHCOCH3) constitute the second group. The most characteristic feature of the first group compounds is the splitting off of the substituent. The hydrocarbon fragment [C11H17]+ thus formed then loses olefin molecules with the formation of corresponding ionic species C11?nH17?2n. The 3-substituted compounds of this group undergo thermal Wagner-Meerwein type rearrangements into adamantane derivatives, resulting in the [C10H15]+ (m/e 135) ion formation; this is the main difference between 1- and 3-substituted homoadamantanes. The series of [CnH2n?6X]+ ions (where X = OCH3, OH, NH2, NHCOCH3, n = 6 to 10) are characteristic of the mass spectra of the second group compounds, the ion [C6H6X]+, [M ? C5H11]+ being the most abundant. The intensity ratio of [M ? C5H11]+ to [M ? C4H9]+ ions is 10:1 for 1-substituted and 3:1 for 3-substituted compounds of this group, allowing the location of the substituent. Some individual features of the spectra are also reported.  相似文献   

6.
The fast atom bombardment (FAB) mass spectra of telluronium salts were studied. The spectra exhibit the intact cation (C+) and cluster ions ([M + C]+). The principal fragment ions in the FAB mass spectra of telluronium salts are [RTe]+, [R2Te]+˙, [R2Te − H]+, [RTeR′]+˙, and [RTeR′ + H]+. When the anion was [BPh4], interesting cluster ions such as [M + C − BPh3]+ appeared.  相似文献   

7.
The chemical ionization mass spectra of several hydroxy steroids were obtained using methane as the reactant gas. The spectra are much less complex than the electron ionization spectra and little fragmentation of the steroid nucleus is observed. The major fragment ions involve the loss of water from [M + H]+. A 3-keto group in the steroids was characterized by an abundant [M + C2H5]+ ion. 5α- and 5β-Dihydrotestosterone could be distinguished by their spectra, with H2 as the reactant gas by marked differences in amounts of [M + H]+, [M + H ? H2O]+ and [M + H ? 2H2O]+. Substituted 3α-X-, 17 β-ol compounds, (X = Cl, Br) were also studied to obtain relative amounts of protonation at these sites.  相似文献   

8.
The mass spectral fragmentation of methyl esters of E and Z isomers of 2,3-dichloro-, 2-bromo-3-chloro-, 3-bromo-2-chIoro- and 2,3-dibromopropenoic acids have been investigated. The M peak is shown with all isomers, the [M ? OCH3]+, [M ? X]+, [M ? OCH3 ? CO]+, [M ? OCH3 ? CO ? X] and [M ? OCH3 ? CO ? X ? X]+ ions constituting abundant peaks in all spectra. The results, particularly from the bromochloro isomers, show that a halogen atom is eliminated from the 3- rather than the 2- position and from the Z rather than the E isomer. Bromine as a bulky atom is preferentially lost.  相似文献   

9.
Field desorption mass spectra are reported for a range of [M(CO)3(η-arene)]X (MMn or Re, XBF4 or PF6) salts. In most cases the spectra are simple, being dominated by molecular, [M]+·, [M + 1]+, and [MCO]+ ions for the cationic part of their structure. However, with the π-chloroarene complexes [Mn(CO)3(η-ClC6H5)]PF6 and [Mn(CO)3(η-1-Cl, 4-MeC6H4)]PF6, facile loss of the chloro substituent and further fragmentation leads to unusually complex spectra, which include strong peaks arising from recombination of fragment species. Cluster ions are also noted in several cases, allowing identification of the anion.  相似文献   

10.
《Polyhedron》1987,6(10):1885-1899
The synthesis and characterization, chiefly as salts of the anions [X(ONO2)2] (X = H+ or Ag+) (by analysis, X-ray powder photography, vibrational spectra and thermogravimetry) of adducts of the nitrates trans-[M(L)4X2](NO3) (M =Rh or Ir; L = pyridine, perdeuteriopyridine or 4-methylpyridine; X = Cl or Br) with hydrogen nitrate and silver nitrate are described.  相似文献   

11.
N-Acetylcysteine and nine N-acetylcysteine conjugates of synthetic origin were characterized by positive- and negative-ion plasma desorption mass Spectrometry. For sample preparation the electrospray technique and the nitrocellulose spin deposition technique were applied. The fragmentation of these compounds, which are best seen as S-substituted desaminoglycylcysteine dipeptides, shows a similar behaviour to that of linear peptides. In the positive-ion mass spectra intense protonated molecular ion peaks are observed. In addition, several sequence-specific fragment ions (A+, B+, [Y + 2H]+, Z+), immonium ions (I+) and a diagnostic fragment ion for mercap-turic acids (RM+) are detected. The negative-ion mass spectra exhibit deprotonated molecular ions and in contrast only one fragment ion corresponding to side-chain specific cleavage ([RXS]?) representing the xenobiotic moiety. In the case of a low alkali metal concentration on the target, cluster molecular ions of the [nM + H]+ or [nM - H]? ion type (n = 1-3) are observed. The analysis of an equimolar mixture of eight N-acetylcysteine conjugates shows different quasi-molecular ion yields for the positive- and negative-ion spectra.  相似文献   

12.
Herein we report a reversed‐phase high‐performance liquid chromatography tandem mass spectrometry (RP‐HPLC/MS/MS) method for the analysis of positional isomers of triacylglycerols (TAGs) in vegetable oils. The fragmentation behavior of [M + X]+ ions (X = NH4, Li, Na or Ag) was studied on a quadrupole‐time‐of‐flight (Q‐TOF) mass spectrometer under low‐energy collision‐induced dissociation (CID) conditions. Mass spectra that were dependent on the X+ ion and the nature and position of the acyl substituents were observed for four pairs of 'AAB/ABA'‐type TAGs, namely PPO/POP, OOP/OPO, LLO/LOL and OOL/OLO (where P is 16:0, palmitic acid; O is 18:1, oleic acid; and L is 18:2, linoleic acid). For the majority of [M + X]+ adducts, the loss of the fatty acid in the outer positions (sn‐1 or sn‐3) was favored over the loss in the central position (sn‐2), which enabled the determination of the fractional abundance of the isomers. Ratios of the intensity of fragment ions at various AAB/ABA compositions produced linear calibration curves with positive slopes, comparable to those obtained traditionally by ESI‐MS/MS of [M + NH4]+ adducts. The only exceptions were the [M + Ag]+ adducts of the PPO/POP system, which produced calibration curves with negative slopes. Sodium adducts provided the most consistent level of isomeric discrimination for the TAGs studied and also offered the most convenience in that they required no additive to the mobile phase. Therefore, calibration curve data derived from [M + Na]+ adducts were applied to the quantification of TAG regioisomers in sunflower and olive oils. The regiospecific analysis showed that palmitic acid was typically located at positions sn‐1 or sn‐3, whereas unsaturated fatty acids, oleic and linoleic acids were mostly found at the sn‐2 position. Copyright © 2010 Crown in the right of Canada. Published by John Wiley & Sons, Ltd.  相似文献   

13.
The electron impact mass spectra of isomeric methyl ethyl and ethyl methyl halosuccinates (X = Cl and Br) are surprisingly different. Only the isomers with the ethyl group remote from the halogen give rise to [M - X]+ ions. A low-energy collision-induced dissociation study of deuterium-labelled analogues of the former isomers indicates that the [M - X]+ ions are mixtures of protonated methyl ethyl maleate (major component, > 85%) and fumarate, and the loss of the halogen atom is a multi-step process including at least two specific hydrogen transfers. Migration of a β-hydrogen atom to the carbonyl oxygen within the ethoxycarbouyl group produces a primary radical site in a distonic intermediate which, by subsequent abstraction of a hydrogen atom from C(3), triggers the ejection of X from C(2) with concomitant double bond formation. Whereas in the other isomer an [M - X]+ ion is absent or negligible, a characteristic double loss of C2H4 and CO2 is observed.  相似文献   

14.
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   

15.
The positive and negative FAB mass spectra of a series of alkoxy- and chloro-silanes Xm(CH3)3-mSi(CH2)nR [m = 1 or 3, n = 3, 10 or 17, X = Cl or OMe or OEt, R = Me, NH2, glycidoxy, COOMe, NHCO(CH2)7COOMe or NHCO(CH2)10CH2OAc] were recorded in NBA and NPOE matrices. The chlorosilanes underwent rapid hydrolysis into silanols which condense to form siloxanes, the process being complete in NBA and partial in NPOE, yielding siloxane-based fragment ions in the positive spectra and silyloxyanions in the negative spectra. The alkoxysilanes were more resistant to hydrolysis, affording abundant [MH – HX]+ ions (X = OMe or OEt) in their positive FAB spectra and moderate to high intensity [M – H]? ions in the negative mode, the latter undergoing characteristic sequential loss of C2H4, EtOH and C2H4. Significant variations were observed in the positive spectra of all the silanes with change of matrix.  相似文献   

16.
Secondary ion mass spectra of N-methylpyridinium halides (C+X?, where C+ is a pyridinium cation and X? is a halogen anion) exhibit the C+ ions, a series of cluster ions ((C+)n(X?)n–1) and, furthermore, remarkable [CX – R]+ ions (R = H or Me). The mechanism of the formation of [CX – R]+ ions was investigated by the use of deuterated compounds and B/E and B2/E constant linked-scan measurements. A possible explanation is proposed in which the ions are produced through substitution reactions between species constituting the C2X+ cluster ions in the gas phase.  相似文献   

17.
The main fragmentation of the compounds MX3-noxn (oxH=8-quinolinol. n = 3; M=AL, Ga, In, Sc, Cr or Fe. n = 2; M=In or Fe; X=Cl or Br. InIox2. n = 1; M=AL, In or Fe; X= Cl or Br) involves loss of X and intact ox. radicals. The comparative abundances of the fragments are primarily related to the common oxidation states of the metals. For example, all the Mox3 compounds show the ions [Mox3]+ and [Mox2]+. The ions [Mox]+ and [M]+ are present when M=Ga, In, Cr or Fe but for the elements with only one oxidation state (Al or Sc) [M]+ is absent and [Mox]+ has only very low abundance. When M= Cr or Fe metal-containing ions arising from loss of species such as CO, H2O, HX, C2H2, H and OH by fragmentation of the ox ligand are also present; this behaviour is rationalised in terms of the ability of these metals to undergo a unit change in oxidation state. When n=1 the ions [MXox2]+ and [Mox2]+ and when n= 2 the ions [MX2ox]+ and [Mox3]+ are present; these ions arise by ionization and fragmentation of species formed by redistribution reactions in the mass spectrometer.  相似文献   

18.
A study of the chemical ionization (CI) and collisional activation (CA) spectra of a number of α, β-unsaturated nitriles has revealed that the even-electron ions such as [MH]+ and [MNH4]+ produced under chemical ionization undergo decomposition by radical losses also. This results in the formation of M +˙ ions from both [MH]+ and [MNH4]+ ions. In the halogenated molecules losses of X˙ and HX compete with losses of H˙ and HCN. Elimination of X˙ from [MH]+ is highly favoured in the bromoderivative. The dinitriles undergo a substitution reaction in which one of the CN groups is replaced with a hydrogen radical and the resulting mononitrile is ionized leading to [M ? CN + 2H]+ under CI(CH4) or [M ? CN + H + NH4] and [M ? CN + H + N2H7]+ under CI(NH3) conditions.  相似文献   

19.
The electron-impact-induced mass spectra of 1,3-dioxolane (la), 1,3-dithiolane (2a) and 1,3-oxatbiolane (3a) and their 2-methyl (1b–3b) and 2,2-dimethyl [(CH3)2: 1c–3c or (CD3)2: 1d–3d] derivatives have been studied in detail to gain further insight into their ion structures and competing reaction pathways with low-resolution, high-resolution, metastable and collision-induced dissociation (CID) techniques. For compounds 1a–1d the most significant reaction is loss of H˙ and CH3˙ by α-cleavage and a subsequent formation of CHO+ and C2H3O+ ions. The [M ? H]+ ions from 1a and 1b give a C2H3O+ ion which does not have the acyl cation structure as shown by their CID spectra. In compounds 3a–3d the sulphur-containing ions predominate, the C2H3O+ now having the acyl cation structure. 1,3-Dithiolanes (2a–2d) exhibit the most complicated fragmentation patterns. Furthermore the [M ? H]+ ion from 2a and [M ? CH3]+ ion from 2b have different structures as well as the [M ? H]+ ion from 2b and [M ? CH3]+ ion from 2c, as shown by their CID spectra. This can be utilized to explain why 3a–3c and 2a give principally a thiiranyl cation, whereas 2b gives a mixture of this and the thioacyl cation and 2c practically only the open-chain thioacetyl cation.  相似文献   

20.
The main fragmentation pathways of the N-1, C-2 and C-4 stereoisomers of the 1,2-dimethyl-4-R-transdecahydroquinoline-4-ol N-oxides (R=C?CH, CH?CH2 and C2H5) under electron impact are discussed. The correlation between the mass spectrometric chromatographic behaviour and the configuration of polar groups in the N-oxides examined is discussed. The mass spectra of the N-1 stereoisomers may be subdivided into two groups, depending only on the orientation of N→O group and not of the 4-OH group. The spectra of N-oxides with the axial N-oxide group reveal less intense ions and much more intense [M? CH3]+, [M? O]+, [M? OH]+ and ions, whereas in the spectra of their equatorial epimers the abundance of the ions exceeds the intensities of the latter ions.  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