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 use of orthogonal acceleration quadrupole time-of-flight (Q-TOF) mass spectrometry to determine the collisionally activated dissociation (CAD) of a test compound 1-(3-[5-[1,2,4-triazol-4-yl]-1H-indol-3-yl]propyl)-4-(2-[3-fluorophenyl]ethyl)piperazine is described. At unit-mass resolution the identity of many ions is ambiguous because of the complexity of the resulting product ion spectrum. Using the high resolution capabilities of the Q-TOF instrument, exact masses for each fragment were determined. These data were used to infer molecular formulas for each fragment through software interpretation and, by further applying chemical intuition, the majority of ions were fully assigned. Additionally, by utilizing in-source fragmentation at high cone voltage, analyses of second-generation products allowed derivation of a consistent sequential fragmentation pathway. This study clearly demonstrates the power of Q-TOF mass spectrometry to elucidate complex product ion spectra.  相似文献   

2.
The electron-impact induced fragmentation of twenty 3-alkyl-1-phenyl-Δ2-1,2,4-triazolin-5-ones, three 1-phenyl-1,2,4-triazolin-3,5-diones and ten 2-alkyl-4-phenyloxadiazolin-5-ones has been studied by conventional mass spectrometry. The major cleavages take place in the 1-phenyl-Δ2-1,2,4-triazolin-5-one nucleus, producing three major fragment ions. 4-Phenyloxadiazolin-5-ones exhibit a similar fragmentation pattern to 1-phenyl-Δ2-1,2,4-triazolin-5-ones. Furthermore, several additional fragmentation processes are observed in the case of specific 1,2,4-triasolines.  相似文献   

3.
The 1,5-diphenyl-1,5-azapentanediyl biradical Ia was generated by photolysis of 1,2-diphenylazacyclopentane (pyrrolidine 1a). Among the reaction pathways followed by Ia, C-N bond reformation with ring closure was found to be the predominating process, as determined by separate irradiation of either of the pure enantiomers of 1a. Disproportionation was a minor process and took place only via H abstraction by the C5 benzylic radical. Another minor pathway was C5-aryl coupling, with formation of 5-phenyl-2,3,4,5-tetrahydro-1H-benzo[b]azepine (4a), which is equivalent to photo-Claisen rearrangement of 1a. Likewise, the 1,4-diphenyl-1,4-azabutanediyl biradical Ib was generated by photolysis of 1,2-diphenylazacyclobutane (azetidine 1b). This species underwent predominating C2-C3 cleavage, as indicated by the extensive styrene formation. Although N1-C4 bond reformation also took place, this is not the major pathway occurring from Ib. Besides, C4-aryl coupling to give 4-phenyl-1,2,3,4-tetrahydroquinoline (4b) was also observed. All the possible reaction pathways were theoretically studied at the UB3LYP/6-31G computational level; the results were found to be in good agreement with the experimental observations.  相似文献   

4.
Cycloaddition of sulfene to N,N-disubstituted 3-amino-2-methyl-1-phenyl-2-propen-1-ones (I) and 3-amino-1,2-diphenyl-2-propen-1-ones (II) occurred in good to moderate yield only in the case of aliphatic N-substitution to give 4-dialkylamino-3,4-dihydro-(5-methyl-6-phenyl)(5,6-diphenyl)-1,2-oxathiin 2,2-dioxides. Polar 1,4-cycloaddition of dichloroketene to I and II occurred only in the former case, giving in good to moderate yield N,N-disubstituted 4-amino-3,3-dichloro-3,4-dihydro-5-methyl-6-phenyl-2H-pyran-2-ones which were dehydrochlorinated with DBN to N,N-disubstituted 4-amino-3-chloro-5-methyl-6-phenyl-2H-pyran-2-ones. In the reaction of 2-methyl-1-phenyl-3-diphenylamino-2-propen-1-one with dichloroketene, a product was isolated which was proven by uv, ir, nmr and chemical evidence to be the dipolar ion VI, the supposed intermediate of the polar 1,4-cycloaddition of dichloroketene to N,N-disubstituted enaminones.  相似文献   

5.
Fast atom bombardment mass spectrometry in the positive mode was used for the characterization of sodiated glycerol phosphatidylcholines. The relative abundance (RA) of the protonated species is similar to the RA of the sodiated molecular species. The sodiated fragment ion, [M + Na - 59](+), corresponding to the loss of trimethylamine, and other sodiated fragment ions, were also observed. The decomposition of the sodiated molecule is very similar for all the studied glycerol phosphatidylcholines, in which the most abundant ion corresponds to a neutral loss of 59 Da. Upon collision-induced dissociation (CID) of the [M + Na](+) ion informative ions are formed by the losses of the fatty acids in the sn-1 and sn-2 positions. Other major fragment ions of the sodiated molecule result from loss of non-sodiated and sodiated choline phosphate, [M + Na - 183](+), [M + Na - 184](+.) and [M + Na - 205](+), respectively. The main CID fragmentation pathway of the [M + Na - 59](+) ion yields the [M + Na - 183](+) ion, also observed in the CID spectra of the [M + Na](+) molecular ion. Other major fragment ions are [M + Na - 205](+) and the fragment ion at m/z 147. Collisional activation of [M + Na - 205](+) results in charge site remote fragmentation of both fatty acid alkyl chains. The terminal ions of these series of charge remote fragmentations result from loss of part of the R(1) or R(2) alkyl chain. Other major informative ions correspond to acylium ions.  相似文献   

6.
The fragmentation of 1-phenyl-, l-(2′-pyridyl)- and 1-(4′-methyl-2′-quinolyl)-4-acetoacetyI-3-methyIpyrazol-5-ols (compounds 1, 2 and 3, respectively) on electron impact has been studied and the major processes interpreted. The common feature in the mass spectra of these compounds is the loss of ketene, acetonyl radical, acetone and two molecules of ketene from the molecular ion. Whereas the ion generated after the last process, which corresponds to 1-substituted-3-methyIpyrazol-5-ols, loses methyl cyanide in the case of 1, similar ions in the case of 2 and 3 lose ?2HO moiety, necessitating an intramolecular hydrogen transfer followed by ring fission and subsequent loss of methyl cyanide. All these and other related processes have been substantiated with the help of accurate mass measurements of the fragment ions and B/E linked-scan spectra.  相似文献   

7.
We have investigated secondary ion yield enhancement using Bin2+ (n=1, 3, 5) primary ions impacting phenylalanine, 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC), 1,2-dipalmitoyl-sn-glycero-3-phosphoethanolamine (DPPE), cholesterol, Irganox 1010, and polymer films adsorbed on silicon and aluminum. Secondary ion yields are increased using Bi2+and Bi3(2+) primary ions for the molecular layers and polymers that can undergo allyl cation rearrangements. For Irganox 1010, the deprotonated molecular ion yields (m/z 1175; [M-H]-) are one to two times larger for Bi2+ and Bi(3)2+ primary ions than for Bi+ and Bi3+ at the same primary ion velocities. In the positive ion mode, the largest fragment ion yield (m/z 899) is 1.5 times larger for Bi2+ ions than for Bi+. For Bi3(2+) the largest fragment ion yield is only 70% of the ion yield using Bi3+, but the secondary ion yields of the fragment ions at m/z 57 and 219 are enhanced. For polymers that can undergo allyl cation rearrangement reactions the secondary ion yield enhancements of the monomer ions range from 1.3 to 4.3. For Bi(5)2+ primary ions, secondary ion yields were the same or slightly larger than for Bi5+ in the negative ion mass spectra for Irganox 1010, but lower in the positive ion mode. No secondary ion yield enhancements were measured on polymer samples for Bi5(2+). For all polymer films studied, secondary ion intensities from the oligomer regions are substantially decreased using Bin2+ (n=1, 3, 5). We discuss differences in the ionization mechanisms for doubly and singly-charged Bi primary ion bombardment.  相似文献   

8.
We report the synthesis of 5-[5′-(1′,2′:3′,4′-di-O-isopropylidene-β-L-arabinopyranosyl)]tetrazole, from 1,2:3,4-di-O-isopropylidene-α-D-galacto-1,6-hexodialdo-1,5-pyranose oxime via 1,2:3,4-di-O-isopropylidene-α-D-galcturononitrile as intermediate by 1,3-dipolar cycloaddition. We also report the synthesis of 5-methyl- and 5-phenyl-2-[5′-(1′,2′:3′,4′-di-O-isopropylidene-β-L-arabinopyranosyl)]-1,3,4-oxadiazole from the tetrazole derivative. The physical and spectroscopic characterizations of the heterocyclic derivatives as well as the intermedi ate nitrile and the principal by product are described and we discuss its possible formation pathway. We present the preferential conformation in solution using computational calculation and spectroscopic data.  相似文献   

9.
The splitting of the dihydropyridine ring of N-methyl-substituted 4-phenyl-5-oxo-4,5-dihydroindeno[1,2-b]pyridine in an acid medium takes place at the C-N bond. During the splitting of 1,2-dimethyl-4-phenyl-4,5-dihydroindeno[1,2-b]-pyridine, 4-phenyl-4-(indane-1,3-dion-2-yl)butan-2-one is formed, while in the case of the 3-ethoxycarbonyl derivative of indenopyridine, together with the Michael retroreaction leading to 2-benzylideneindane-1,3-dione, a recyclization of the intermediate product into a derivative of dihydroindeno-2-pyridone takes place.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 10, 1363–1366, October, 1986.  相似文献   

10.
The product ions of the BH(2)(+) ions formed by the glycosidic cleavage of N-(deoxyguanosin-O(6)-yl)-2-methylaniline, 4-(deoxyguanosin-8-yl)-2-methylaniline, and N-(deoxyguanosin-1-yl)-2-methylaniline have been studied using matrix-assisted laser desorption/ionization (MALDI) and post-source decay (PSD) to identify fragment ions and pathways that may be used to differentiate their structures. All three isomers may be distinguished based on their PSD product ion spectra using only femtomole quantities of sample. N-(Deoxyguanosin-O(6)-yl)-2-methylaniline produces product ions at m/z 107 and 134 that are diagnostic for 2-methylaniline attachment to the O(6) position of guanine. The BH(2)(+) ion from 4-(deoxyguanosin-8-yl)-2-methylaniline yields a product ion formed by the consecutive losses of 17 and 42 u neutral fragments that may be regarded as specific for guanine-arylamine adducts that possess two primary amine groups. The BH(2)(+) ion from 4-(deoxyguanosin-8-yl)-2-methylaniline yields no product ions that correlate with specificity for guanine N1 substitution. However, the product ion abundance ratio of the protonated arylamine to that of the ammonia loss ion may be used to differentiate an adduct formed by N1 substitution from other arylamine adducts of guanine studied thus far.  相似文献   

11.
Reaction of 1-naphthol with 2-chlorocyclohexanone in alkaline alcohol gave as the major product 5-(2′-oxocyclohexyl)-7,8,9,10-tetrahydronaphtho[1,2-b]benzofuran ( 1 ), which could be converted to the title compound 5 by reduction and dehydrogenation. This product arises from ambident alkylation of 1-naphthol at the 2-and 4-positions. Via the 2′-oxocyclohexyl ether, 5 was also synthesized from 4-phenyl-1-naphthol.  相似文献   

12.
This study has elucidated the fragmentation pathway for deprotonated isoflavones in electrospray ionization using MS(n) ion trap mass spectrometry and triple quadrupole mass spectrometry. Genistein-d(4) and daidzein-d(3) were used as references for the clarification of fragment structures. To confirm the relationship between precursor and product ions, some fragments were traced from MS(2) to MS(5). The previous literature for the structurally related flavones and flavanones located the loss of ketene (C(2)H(2)O) to ring C, whereas the present fragmentation study for isoflavones has shown that the loss of ketene occurs at ring A. In the further fragmentation of the [M-H-CH(3)](-*) radical anion of methoxylated isoflavones, loss of a hydrogen atom was commonly found. [M-H-CH(3)-CO-B-ring](-) is a characteristic fragment ion of glycitein and can be used to differentiate glycitein from its isomers. Neutral losses of CO and CO(2) were prominent in the fragmentation of deprotonated anions in ion trap mass spectrometry, whereas recyclization cleavage accounted for a very small proportion. In comparison with triple quadrupole mass spectrometry, ion trap MS(n) mass spectrometry has the advantage of better elucidation of the relationship between precursor and product ions.  相似文献   

13.
The addition of 7-chloro-2-hydrazono-5-phenyl-1,2-dihydro-3H-1,4-benzodiazepine 3 to a mixture of sodium acetate and 1,1′-carbonyldiimidazole 1 at room temperature gave, in moderate yields, carbonyl-1,1′-bis[7-chloro-5-phenyl-1,2-dihydro-3H-1,4-benzodiazepin-2-ylidene hydrazone] 7 instead of the expected 2-acetylhydrazono-7-chloro-5-phenyl-1,2-dihydro-3H-1,4-benzodiazepine 4 .  相似文献   

14.
The one-pot reaction of 2-tert-butylthio-3-phenylcyclopropenethione (1a) and its 3-(2-thienyl) derivative (1b) with lithium pyrrolidinide at -70 degrees C, followed by methylation with methyl iodide, gives 6-methylthio-5-phenyl-2,3-dihydro-1H-pyrrolizine (2a) and its 5-(2-thienyl) derivative (2b), respectively. The reaction of 2-tert-butylthio-3-(pyrrolidin-1-yl)cyclopropenethione (1c) with phenyllithium gives also 2a in a high yield under similar conditions, and the reactions of 1a with N-lithium salts of 3-pyrroline, hexamethyleneimine, indoline, and carbazole, piperidine-potassium tert-butoxide mixture, and phenyllithium give 6-methylthio-5-phenyl-3H-pyrrolizine (3), 2-methylthio-3-phenyl-6,7, 8,9-tetrahydro-5H-pyrrolo[1,2-a]azepine (5), 6-tert-butylthio-5-methylthio-4-phenyl-1,2-dihydro-6H-pyrrolo[3,2, 1-ij]quinoline (6), 4-tert-butylthio-5-methylthio-6-phenyl-4H-pyrido[3,2,1-jk]carbazole (7), 2-methylthio-3-phenyl-5,6,7,8-tetrahydroindolizine (4), and 1-tert-butylthio-2-methylthio-3-phenylindene (9), respectively. The structures of 2a and 3 were determined by X-ray analyses of their tricarbonylchromium complexes.  相似文献   

15.
Low-energy collision-induced dissociation (CID) of acetylcholine (ACh) yields only two fragment ions: the dominant C(4)H(7)O(2)(+) ion at m/z 87, arising from trimethylamine loss; and protonated trimethylamine at m/z 60. Since the literature is replete with conflicting mechanisms for the loss of trimethylamine from ACh, in this article density functional theory (DFT) calculations are used to assess four competing mechanisms: (1) Path A involves a neighboring group attack to form a five-membered ring product, 2-methyl-1,3-dioxolan-2-ylium cation; (2) Path B is a neighboring group attack to form a three-membered ring product, 1-methyl-oxiranium ion; (3) Path C involves an intramolecular elimination reaction to form CO protonated vinylacetate; and (4) Path D is a 1,2-hydride migration reaction forming CH(2)-protonated vinylacetate. At the MP2/6-311++G(2d,p)//B3-LYP/6-31+G(d,p) level of theory path A is the kinetically favored pathway, with a transition-state energy barrier of 37.7 kcal mol(-1) relative to the most stable conformer of ACh. The lowest energy pathway for the formation of protonated trimethylamine was also calculated to proceed via path A, involving proton transfer within the ion-molecule complex intermediate, with the exocylic methyl group being the proton donor. To confirm the site of proton transfer, low-energy CID of acetyl-d(3)-choline (d(3)-ACh) was carried out, which revealed loss of trimethylamine and the formation of Me(3)ND(+).  相似文献   

16.
Electron impact mass spectra of alkyl 4-alkoxy-5-amino-3-methylthiophene-2-carboxylates were studied for the first time. These compounds, except for 4-(1-ethoxyethoxy) and 4-(ferrocenylmethoxy) derivatives, give rise to a stable molecular ion whose decomposition follows three pathways. The main fragmentation pathway of the molecular ion is elimination of alkyl radical from the 4-alkoxy group, the second pathway involves expulsion of alkoxy group from the ester moiety, and the third pathway is decomposition of the thiophene ring. The molecular ions of 4-(1-ethoxyethoxy)thiophenes decompose mainly via elimination of ethyl vinyl ether molecule with formation of [M–VinOEt]+ · odd-electron ion, and fragmentation of the latter follows general pathways. In the mass spectra of 4-(ferrocenylmethoxy)thiophenes the most abundant are ferrocenylmethyl ion with m/z 199 (I rel 100%) and fragment ions derived therefrom.  相似文献   

17.
Perfluoro(1-ethyl-1,2-dihydrocyclobutabenzene) reacts with pentafluorobenzene in SbF5 to give perfluoro(1-ethyl-2-phenyl-1,2-dihydrocyclobutabenzene). Analogous reaction of a mixture of perfluoro(1,1-diethyl-1,2-dihydrocyclobutabenzene) and perfluoro(1,2-diethyl-1,2-dihydrocyclobutabenzene) leads to the formation (after hydrolysis of the reaction mixture) of perfluorinated 7-phenyl-8,8-diethylbicyclo[4.2.0]octa-1,4,6-trien-3-one, 1,1-diethyl-2-(4-oxocyclohexa-2,5-dienylidene)-1,2-dihydrocyclobutabenzene, and 2-(pent-2-en-3-yl)benzophenone (from the 1,1-isomer) and perfluorinated (E)-1,2-diethyl-1-phenyl-1,2-dihydrocyclobutabenzene, 7,8-diethyl-8-phenylbicyclo[4.2.0]octa-1,4,6-trien-3-one, and 1-[2-(1-phenylprop-1-en-1-yl)-phenyl]propan-1-one (from the 1,2-isomer).  相似文献   

18.
Certain 1,2,4-trisubstituted imidazoles underwent electrophilic attack of azodicarbonyl compounds on the 5-position to form 5-(1,2-dialkoxycarbonyl)hydrazino- and 5-(4-phenyl-3,5-dioxo-1,2,4-triazolidin-1-yl)imidazole derivatives in moderate to high yields. The reaction was highly susceptible to the nature and the substitution pattern of the substituents on the imidazole ring. Thus 1,4-di- or 1,2,5-trisubstitued imidazoles, and 2-methylsulfinylimidazoles gave no reaction. Reductive cleavage of the tetrasubstituted imidazoles with zinc dust-acid gave the 1- or 1,2-cleaved product depending upon the reaction temperature, but the hydrazino moiety remained intact.  相似文献   

19.
The title compounds serve as potential precursors to aryloxenium ions, often proposed, but primarily uncharacterized intermediates in phenol oxidations. The uncatalyzed and acid-catalyzed decomposition of 4-acetoxy-4-phenyl-2,5-cyclohexadienone, 2a, generates the quinol, 3a. (18)O-Labeling studies performed in (18)O-H(2)O, and monitored by LC/MS and (13)C NMR spectroscopy that can detect (18)O-induced chemical shifts on (13)C resonances, show that 3a was generated in both the uncatalyzed and acid-catalyzed reactions by C(alkyl)-O bond cleavage consistent with formation of an aryloxenium ion. Trapping with N(3)(-) and Br(-) confirms that both uncatalyzed and acid-catalyzed decompositions occur by rate-limiting ionization to form the 4-biphenylyloxenium ion, 1a. This ion has a shorter lifetime in H(2)O than the corresponding nitrenium ion, 7a (12 ns for 1a, 300 ns for 7a at 30 degrees C). Similar analyses of the product, 3b, of acid- and base-catalyzed decomposition of 4-acetoxy-4-methyl-2,5-cyclohexadienone, 2b, in (18)O-H(2)O show that these reactions are ester hydrolyses that proceed by C(acyl)-O bond cleavage processes not involving the p-tolyloxenium ion, 1b. Uncatalyzed decomposition of the more reactive 4-dichloroacetoxy-4-methyl-2,5-cyclohexadienone, 2b', is also an ester hydrolysis, but 2b' undergoes a kinetically second-order reaction with N(3)(-) that generates an oxenium ion-like substitution product by an apparent S(N)2'mechanism. Estimates based on the lifetimes of 1a, 7a, and the p-tolylnitrenium ion, 7b, and the calculated relative stabilities of these ions toward hydration indicate that the aqueous solution lifetime of 1b is ca. 3-5 ps. Simple 4-alkyl substituted aryloxenium ions are apparently not stable enough in aqueous solution to be competitively trapped by nonsolvent nucleophiles.  相似文献   

20.
在质谱实验中,我们发现了与一般单分子分解反应相悖的四种类型碎片产物的重组反应物即(Ⅰ)裂解碎片的自组合产物如S_8~+及不同聚合度的聚硫离子;(Ⅱ)裂解碎片之间的相互组合物如[Fe(S_2CNC_4H_8)_3]~+和[Fe(S_2CNEt_2)_3]~+;(Ⅲ)裂解碎片的阴离子和阳离子部分碎片间的组合物如[Fe_2S_2(NO)_4(CH_3)_2]~+;(Ⅳ)歧化反应产物Cp_3Yb~+和YbL_3~+(Cp=C_5H_5),L=β-二酮)。本文阐述了这种重组反应或离子—分子反应在质谱条件下产生的可能性及其理论根据。  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