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 tautomeric constants of a series of azo dyes were estimated in the gas phase by using electron ionization mass spectrometry. It was shown that the relative amount of the keto tautomer increases from 4‐phenylazo‐1‐phenol to 4‐phenylazo‐anthracen‐1‐ol, thus confirming the quantum‐chemical predictions. The existence of the enol tautomer of 4‐phenylazo‐anthracen‐1‐ol is shown for the first time by mass spectrometry in the gas phase. This finding is supported by flash photolysis measurements in solution. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

2.
The Schiff base enaminones (3Z)‐4‐(5‐ethylsulfonyl‐2‐hydroxyanilino)pent‐3‐en‐2‐one, C13H17NO4S, (I), and (3Z)‐4‐(5‐tert‐butyl‐2‐hydroxyanilino)pent‐3‐en‐2‐one, C15H21NO2, (II), were studied by X‐ray crystallography and density functional theory (DFT). Although the keto tautomer of these compounds is dominant, the O=C—C=C—N bond lengths are consistent with some electron delocalization and partial enol character. Both (I) and (II) are nonplanar, with the amino–phenol group canted relative to the rest of the molecule; the twist about the N(enamine)—C(aryl) bond leads to dihedral angles of 40.5 (2) and −116.7 (1)° for (I) and (II), respectively. Compound (I) has a bifurcated intramolecular hydrogen bond between the N—H group and the flanking carbonyl and hydroxy O atoms, as well as an intermolecular hydrogen bond, leading to an infinite one‐dimensional hydrogen‐bonded chain. Compound (II) has one intramolecular hydrogen bond and one intermolecular C=O...H—O hydrogen bond, and consequently also forms a one‐dimensional hydrogen‐bonded chain. The DFT‐calculated structures [in vacuo, B3LYP/6‐311G(d,p) level] for the keto tautomers compare favourably with the X‐ray crystal structures of (I) and (II), confirming the dominance of the keto tautomer. The simulations indicate that the keto tautomers are 20.55 and 18.86 kJ mol−1 lower in energy than the enol tautomers for (I) and (II), respectively.  相似文献   

3.
The self‐chemical ionization (SCI) in quadrupole mass spectrometry was developed to determine the structure of (E)‐2‐iod‐3‐(phenylsulfinyl)‐2‐propen‐1‐ol and its 6 analogues. Some techniques that increase the sample quantity and heating speed and shorten vaporization time to obtain high pressure in the ion source were applied to increase the chance of ionmolecule reactions. The structures of these compounds were identified by mass spectral data of MH+ and some characteristic fragment ions. Compared with the mass spectra for 2‐iodo‐3‐(phenylsulfinyl)‐2‐propen‐l‐ols obtained in electron impact ionization (EI), SCI showed more information, in particular, an improvement in amount of information at the high mass area. The absence of reagent gas makes the spectrum clean and simple.  相似文献   

4.
The relative stability of 1,4‐benzodiazepin‐2‐one tautomers in the gas phase and model solvents was calculated at the M06 and ωB97XD levels of theory. The two density functionals were benchmarked earlier and demonstrated as excellent models to study tautomerism in a vast array of chemical systems. A number of commercially available 1,4‐benzodiazepin‐2‐ones were investigated computationally for the first time. In addition, some biologically active and newly devised benzodiazepines were considered, which may be important in designing structures with desired (bio)chemical features. Special attention was paid to determine substituent effects on the Gibbs free energies of keto, enol, and iminol forms for each respective benzodiazepine. It was demonstrated that (i) the replacement of the benzene ring by the heterocyclic ring in the benzodiazepine system may stabilize the iminol tautomer, and (ii) the electron‐withdrawing substituent at the C3‐position of the respective benzodiazepine may stabilize the enol tautomer relative to the parent keto form. It is concluded that substituent effects may govern the chemical reactivity and biological properties of selected benzodiazepines.  相似文献   

5.
The electron ionization (EI) mass spectra and electrospray ionization (ESI) mass spectra of a series of isomeric N‐chlorobenzyl‐ and N‐bromobenzyl‐substituted (E)‐2′(3′ or 4′)‐hydroxy‐4‐stilbazole chlorides and bromides (1–12) were recorded. The fragmentation pathways of all of the compounds and the characteristic fragment ions formed by EI‐MS were studied by means of B/E and B2/E constant linked‐scanning techniques. The formation of ions originating from preionization reactions, characteristic of quaternary halides under EI‐MS conditions, such as the elimination of chloro‐ or bromobenzyl halides, dehydrohalogenation or substitution reactions, is explained. As soft ionization methods cause no such degradation reactions, the ESI‐MS spectra of the studied compounds were also obtained for comparison. We thus demonstrated the applicability of EI‐MS even in cases when preionization takes place, as long as such secondary processes are properly accounted for. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

6.
蔺楠  王剑波等 《中国化学》2002,20(8):789-794
Electron impact-induced fragmentation mechanism of Trans-α-Aryl-β-enamino esters were investigated using mass-analyzed ion kinetic energy (MIKE) spectrometry and high resolution accurate mass data It was found that the main characteristic fragmentations of compounds studied were:an odd electron ion M^ -EtOH was formed by losing a neutral molecule of ethanol;and the skeletal rearrangements took place;and the ring opening reaction happened after losing a carbon monoxide;and the typical McLafferty rearrangement underwent in ester group.The cycliztion reation caused by losing neutral molecule of TsNH2 due to the ortho-effects of substituted group of gromatic ring was also observed.  相似文献   

7.
Chemical Weapons Convention (CWC)‐related compounds where the phosphorus atom is part of a ring have very limited representation in mass spectral libraries and the open literature. Here we report electron ionization (EI), chemical ionization (CI) and electrospray ionization tandem mass spectrometry (ESI‐MSn) spectra and retention indices for 2‐alkyl‐1,3,6,2‐dioxathiaphosphocane‐2‐oxides (alkyl C1 to C3) which are new cyclic chemicals covered under the CWC. The EI mass spectra show a pattern of ion fragmentation that is similar to that of other cyclic phosphonates in that loss of the alkylphosphonic acid as a neutral loss is more important than the presence of the protonated alkylphosphonic acid. In contrast to other cyclic phosphonates, the 2‐alkyl‐1,3,6,2‐dioxathiaphosphocane‐2‐oxides show almost no protonated alkylphosphonic acid and as a result the spectra do not carry the same distinctive signature of the phosphorus–carbon bond that is required for the chemical to be covered under the CWC. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

8.
The fragmentation patterns of a series of six novel synthesized benzopyranopyrimidine derivatives 16, possessing the same 2‐oxo‐2H‐benzopyrano[2,3‐d]pyrimidine backbone structure, were investigated by electrospray ionization mass spectrometry (ESI‐MS) and tandem mass spectrometry (MS/MS) techniques using a quadrupole orthogonal time‐of‐flight (QqToF)‐hybrid instrument. The series of six pure benzopyranopyrimidine compounds contained three constitutional isobaric isomers (compounds 46). A simple methodology, based on the use of ESI (positive ion mode) and increasing the declustering potential in the atmospheric pressure/vacuum interface resulting in collision‐induced dissociation (CID), was used to enhance the formation of the product ions. In general, the novel synthetic benzopyranopyrimidine derivatives 16 afforded exact accurate masses for the protonated molecules. This led to the confirmation of both molecular masses and chemical structures of the studied compounds. The breakdown routes of the protonated molecules were rationalized by conducting low‐energy CID‐MS/MS analyses. It was shown that the MS/MS fragmentation routes for the protonated molecules 1 and 2 were similar, and that the MS/MS fragmentations of the constitutional isobaric protonated molecules 5 and 6 were identical. It was also shown that the gas‐phase CID fragmentations of 5 and 6 were different from that of their constitutional isomer 4. Finally, the ESI‐MS and CID‐MS/MS analyses of the protonated molecules that were obtained from the monodeuterated benzopyranopyrimidine derivatives 16 confirmed the values obtained for the exact masses, the precise structural assignments of all product ions and all the pathways described in the proposed CID fragmentations. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

9.
Two different tautomeric forms of a new Schiff base, C17H19N3O2·C17H19N3O2, are present in the crystal in a 1:1 ratio, namely the enol–imine form 4‐(1‐{[4‐(dimethylamino)benzylidene]hydrazono}ethyl)benzene‐1,3‐diol and the keto–amine form 6‐[(E)‐1‐{[4‐(dimethylamino)benzylidene]hydrazino}ethylidene]‐3‐hydroxycyclohexa‐2,4‐dien‐1‐one. The tautomers are formed by proton transfer between the hydroxy O atom and the imine N atom and are hydrogen bonded to each other to form a one‐dimensional zigzag chain along the crystallographic b axis via intermolecular hydrogen bonds.  相似文献   

10.
The fragmentation pathways in electron ionization (EI) mass spectra of a series of new N(5)‐oxides of alloxazines and iso‐alloxazine are presented, and compared with those of substituted alloxazines and iso‐alloxazine. The EI mass spectra of these compounds showed characteristic fragmentation pathways A, B and C, started by the ejection of atomic oxygen, a HNCO molecule and an OH . radical, respectively. On the basis of B/E and B2/E spectra, the mechanism of elimination of the OH . radical is discussed. The influence of the methyl substituent in the benzene ring of alloxazine on the mass fragmentation pathways is described. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

11.
Derivatives of 4‐hydroxypyrimidine are an important class of biomolecules. These compounds can undergo keto–enol tautomerization in solution, though a search of the Cambridge Structural Database shows a strong bias toward the 3H‐keto tautomer in the solid state. Recrystallization of 2‐amino‐5,6‐dimethyl‐4‐hydroxypyrimidine, C6H9N3O, from aqueous solution yielded triclinic crystals of the 1H‐keto tautomer, denoted form (I). Though not apparent in the X‐ray data, the IR spectrum suggests that small amounts of the 4‐hydroxy tautomer are also present in the crystal. Monoclinic crystals of form (II), comprised of a 1:1 ratio of both the 1H‐keto and the 3H‐keto tautomers, were obtained from aqueous solutions containing uric acid. Forms (I) and (II) exhibit one‐dimensional and three‐dimensional hydrogen‐bonding motifs, respectively.  相似文献   

12.
Emergence of new psychoactive substances, hallucinogenic phenethylamines in particular, in illicit market is a serious threat to human health in global scale. We have detected and identified N‐(2‐methoxybenzyl)‐2‐(2,4,6‐trimethoxyphenyl)ethanamine (2,4,6‐TMPEA‐NBOMe), a new compound in NBOMe series. Identification was achieved by means of gas chromatography/mass spectrometry (GC/MS), including high‐resolution mass spectrometry with tandem experiments (GC/HRMS and GC/HRMS2), ultra‐high performance liquid chromatography/high‐resolution mass spectrometry with tandem experiments (UHPLC/HRMS and UHPLC/HRMS2), and 1H and 13C nuclear magnetic resonance spectroscopy. The peculiarities of fragmentation of the compound under electron ionization (EI) and collision‐induced dissociation were studied. Despite of the empirical rule denying migration of the hydrogen atom in McLafferty rearrangement to the benzene ring with substituents in the both ortho‐positions, it easily occurs for 2,4,6‐TMPEA‐NBOMe in EI conditions. We have noticed that electron‐donating substituents, e.g. methoxy groups in the both ortho‐positions and para‐positions favor the rearrangement. For specially synthesized N‐methyl and N‐acyl derivatives McLafferty rearrangement is not observed. N‐Acyl derivatives demonstrate McLafferty rearrangement, but the charge retains at the alternative fragment involving N‐acyl carbonyl group. We have also showed that the hydrogen atoms in 2,4,6‐trimethoxybenzene ring may be easily substituted for deuterium or for strong electrophiles like trifluoroacetyl. Analytical characteristics of 2,4,6‐TMPEA‐NBOMe and of some derivatives thereof which enable their determination in various criminal seizures are given. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

13.
The electron ionization (EI) mass spectra of some new bridgehead 7,7‐dimethylnorbornane‐based β‐amino alcohols have been studied and their fragmentation patterns compared with those of isomeric 3,3‐dimethyl derivatives described by us previously. The dimethyl substitution at C7 results in a significant complication of the spectra, although all compounds show a C1–C2 bond cleavage with charge location at the nitrogen atom that leads to the base peak. Thus, two main fragmentation patterns dominated by cyclopentenylimmonium or methyleneimmonium ions are described depending on the position of the amino group in the norbornane framework (C1 or C2, respectively), as well as other secondary routes that can explain the appearance of the less noticeable fragments. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

14.
Tautomerism has been studied conventionally in solutions or in the solid state. However, the importance of mass spectrometry in the gas phase was realized relatively late. 2‐Phenacylquinolines are known to undergo ketimine‐enaminone tautomerism. The ratio of tautomers is dependent on the nature of the phenyl ring substituent and the Hammett substituent constants σ. Theoretical calculations indicate the presence of ketimine and enaminone tautomers in the gas phase. The electron ionization mass spectra of eight 2‐phenacylquinolines (ketimine form) were recorded at 70 eV in order to determine the fragmentation routes and to screen for the presence of their enaminone tautomers, (Z)‐2‐benzoylmethylene‐1,2‐dihydroquinolines, in the gas phase. The relative abundances or total ion currents of some ions correlated with the Hammett substituent constants and Hammett‐Brown constants. The product ions [M–CO]+ . and [M–HCO]+ were observed. A reaction mechanism is suggested for the formation of these ions, requiring skeletal rearrangements. The results furnish information relating to tautomerism in the gas phase. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

15.
The non‐degenerate tautomers of 3‐nitro‐1,2,4‐triazole‐5‐one (NTO) radical anions were investigated for the first time by an ESR method during electrochemical reduction of NTO in an aprotic medium. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

16.
The electron ionisation (EI) mass spectra of a series of bridgehead‐fused Δ2‐norbornanethiazolines, a new class of bridgehead‐norbornane derivatives, have been studied and their cleavage mechanisms rationalised on the basis of the substituent shifts as well as on the identification of relevant peaks through accurate mass measurements and collision‐induced dissociation tandem mass spectrometric experiments. The fragmentation patterns of isomeric pairs of 6,6‐ and 10,10‐dimethylnorbornanethiazolines are almost identical, probably due to an initial isomerisation of molecular ion previous to the fragmentation. In general, the dominant peaks in the spectra of all the studied compounds originate from initial α‐cleavages of C(5)–C(6) or C(1)–C(10) bonds, followed by concomitant homolytic cleavage of C(1)–C(9) and C(7)–C(10) bonds. The driving force for this fragmentation pathway, directed by the gem‐dimethyl group, is the formation of a highly stabilised thiazolilmethyl cation which constitutes the base peak in all the spectra and allows the identification of these interesting ligands. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

17.
Di(hydroxybutyl) ether (DHBE), a liver protecting drug, is composed of a mixture of three regioisomers: 4‐(3‐hydroxybutoxy)‐2‐butanol (1), 3‐(4‐hydroxy‐2‐butoxy)‐1‐butanol (2), and 3‐(3‐hydroxybutoxy)‐1‐butanol (3). Unequivocal differentiation of each regioisomer of DHBE was rapidly obtained without isolation of the single components, using GC‐MS with electron ionization (EI). The mass spectrum of 1 showed a rearrangement ion at m/z 118, characteristic of the 3‐hydroxybutyl chain, deriving from loss of acetaldehyde from the molecular ion, whereas 2 and 3 were characterized by the ion at m/z 117, expected from α‐cleavage of the 4‐hydroxy‐2‐butyl chain. The species at m/z 118, in turn, loses a water molecule via a mechanism involving both alcohol hydrogens, as shown by deuterium exchange experiments. Both this finding and theoretical calculations support a mechanism in which the loss of acetaldehyde in 1 occurs via a cyclic intermediate, stabilized by a strong hydrogen bond between the alcohol oxygen bearing the charge and the other alcohol oxygen, and involves initial hydrogen transfer from the former to the latter. The EI spectrum of 2, having two 4‐hydroxy‐2‐butyl chains, showed the fragmentations expected from classical fragmentation rules of aliphatic ethers and alcohols, whereas the EI spectrum of 3, bearing one 4‐hydroxy‐2‐butyl and one 3‐hydroxybutyl chain, showed essentially the characteristic fragments of both chains. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

18.
Flavan‐3‐ol monoglycosides, having four aglycons (+)‐catechin, (?)‐epicatechin, (?)‐epigallocatechin and epicatechin gallate monomeric units, are detected for the first time in Vitis vinifera L. cv. Merlot grape seeds and wine. These compounds were analyzed in red wine, seed and skin extracts by electrospray ionization quadrupole time of flight mass spectrometry (MS) in negative mode. Fragment ions derived from retro‐Diels Alder, heterocyclic ring fragmentation, benzofuran forming fragmentation and glycoside fragmentations were detected in targeted MS/MS mode. These compounds were not detected in skins; the comparative study showed evidence that these glycosylated compounds originate only from grape seeds. Our method allows for the identification of these glycosylated compounds based on their exact mass and their specific fragmentation pattern. However, exact glucose position on the monomeric units can not be determined. This work allowed us to partially identify 14 new flavan‐3‐ol monoglycosides, based on the exact mass of the molecular ions and their specific retro‐Diels Alder, heterocyclic ring fragmentation, benzofuran forming fragmentation and glycoside fragmentations. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

19.
An efficient and simple route for the preparation of 2‐oxopyridine‐fused 1,3‐diazaheterocyclic compounds via a three component reaction is described. It involves the reaction between alkylenediamines 1 , 1,1‐bis(methylsulfanyl)‐2‐nitroethene, and alkyl prop‐2‐ynoates 2 in refluxing THF (Table). The structures were corroborated by spectroscopic (IR, 1H‐ and 13C‐NMR, and EI‐MS) and elemental analyses. A plausible mechanism for this type of cyclization is proposed (Scheme).  相似文献   

20.
Many metabolomic applications use gas chromatography/mass spectrometry (GC/MS) under standard 70 eV electron ionization (EI) parameters. However, the abundance of molecular ions is often extremely low, impeding the calculation of elemental compositions for the identification of unknown compounds. On changing the beam‐steering voltage of the ion source, the relative abundances of molecular ions at 70 eV EI were increased up to ten‐fold for alkanes, fatty acid methyl esters and trimethylsilylated metabolites, concomitant with 2‐fold absolute increases in ion intensities. We have compared the abundance, mass accuracy and isotope ratio accuracy of molecular species in EI with those in chemical ionization (CI) with methane as reagent gas under high‐mass tuning. Thirty‐three peaks of a diverse set of trimethylsilylated metabolites were analyzed in triplicate, resulting in 342 ion species ([M+H]+, [M–CH3]+ for CI and [M]+ . , [M–CH3]+ . for EI). On average, CI yielded 8‐fold more intense molecular species than EI. Using internal recalibration, average mass errors of 1.8 ± 1.6 mm/z units and isotope ratio errors of 2.3 ± 2.0% (A+1/A ratio) and 1.7 ± 1.8% (A+2/A ratio) were obtained. When constraining lists of calculated elemental compositions by chemical and heuristic rules using the Seven Golden Rules algorithm and PubChem queries, the correct formula was retrieved as top hit in 60% of the cases and within the top‐3 hits in 80% of the cases. Copyright © 2010 John Wiley & Sons, Ltd.  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