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 62 毫秒
1.
Two monometayl- and four dimethyl-triazolocoumarin isomers were characterized by their electron impact mass spectra and by low-energy collision experiments performed on molecular ions M+˙ and other fragment ions with an ion-trap mass spectrometer. High-energy collision-activated dissociation measurements were performed on the protonated [M + H]+ and deprotonated [M ? H]? molecular ion obtained by fast atom bombardment and M+˙ species produced by electron impact ionization on a double-focusing, reverse-geometry instrument. The data obtained allowed unequivocal structural identification of all the compounds investigated.  相似文献   

2.
Studies of the genesis of the [M ? H]+ ion in flavanone and 2′-hydroxychalcone, performed with the aid of metastable decompositions and deuterium labelling, allow new structural notations to be postulated for the [M ? H]+ ions, which in turn provide evidence for the pathways in the [M ? H – ketene]+ fragmentation routes for these compounds.  相似文献   

3.
The unimolecular fragmentations of [M + H]+ and [M – H]? ions from four 2-aryl-2-methyl-1,3-dithianes are described and clarified with the aid of deuterated derivatives. Comparison of the MIKE spectra of [M + H]+ species obtained under chemical ionization and fast atom bombardment (FAB) conditions reveals differences which are attributed to the different energetics involved in the two ionization processes. It is suggested that FAB is a ‘softer’ ionization technique but, at the same time, it provides, for the possibility of solvation, reaction sites not available in gas-phase protonation. [M – H]? species and anionic fragments thereof were generally not obtained under FAB(?) conditions. [M – H]? ions are readily produced in gas-phase reactions with OH? via proton abstraction from C(4) or C(5), and from the 2-methyl substituent; and they fragment according to several reaction pathways.  相似文献   

4.
The collision-induced decompositions of the [M – H]? and [M + Li]+ ions of a few dinucleoside phenylphosphonates were studied using fast atom bombardment and linked scanning at constant B/E. Deprotonation takes place on the base or sugar moieties. The [M – H]? ion decomposes mainly by cleavage on either side of the phosphonate linkage, leading to the formation of mononucleotide fragment ions and also by cleavage of the basesugar bond. Rupture of the 3′-phosphonate bond is preferred. Unlike the normal charged nucleotides, these neutral nucleotides do not eliminate a neutral base from the [M – H]? ion. However, the mononucleotide fragment ions which can have the charge on the phosphorus oxygen eliminate neutral bases by charge-remote fragmentation. The 4,4′-dimethoxytrityl (DMT)-protected nucleotides show the additional fragmentation of loss of DMT. Li+ attachment can occur at several sites in the molecule. As observed for the [M – H]? ion, the major cleavage occurs on either side of the phosphonate bond in the fully deprotected nucleotides, cleavage of the ester bond on C(3′) being preferred. Cleavage of the 5′-phosphonate bond is not observed in the DMT-protected nucleotides. Many of the fragmentations observed can be explained as arising from charge-remote reactions.  相似文献   

5.
The dish-topped metastable peak for the fragmentation [C3H7]+ → [allyl]+ + H2 is generated by the threshold fragmentation. The fraction of the reverse activation energy which is partitioned as translational energy of the products is 0.9 ± 0.1. It is proposed that a similar partitioning coefficient applies to the excess internal energy above threshold.  相似文献   

6.
The closo‐dodecaborate [B12H12]2? is degraded at room temperature by oxygen in an acidic aqueous solution in the course of several weeks to give B(OH)3. The degradation is induced by Ag2+ ions, generated from Ag+ by the action of H2S2O8. Oxa‐nido‐dodecaborate(1?) is an intermediate anion, that can be separated from the reaction mixture as [NBzlEt3][OB11H12] after five days in a yield of 18 %. The action of FeCl3 on the closo‐undecaborate [B11H11]2? in an aqueous solution gives either [B22H22]2? (by fusion) or nido‐B11H13(OH)? (by protonation and hydration), depending on the concentration of FeCl3. In acetonitrile, however, [B11H11]2? is transformed into [OB11H12]? by Fe3+ and oxygen. The radical anions [B12H12] ˙ ? and [B11H11] ˙ ? are assumed to be the primary products of the oxidation with the one‐electron oxidants Ag2+ and Fe3+, respectively. These radical anions are subsequently transformed into [OB11H12]? by oxygen. The crystal structure analysis shows that the structure of [OB11H12]? is derived from the hypothetical closo‐oxaborane OB12H12 by removal of the B3 vertex, leaving a non‐planar pentagonal aperture with a three‐coordinate O vertex, as predicted by NMR spectra and theory.  相似文献   

7.
The reverse activation energy, Erev, for the dissociation [C3H7]+ → [C3H5]+ + H2 has been determined as 0.24 ± 0.06 eV from measurements of the AP of [C3H5]+ produced by electron-impact from thermally generated sec-C3H7 radicals. The energy release observed in the corresponding metastable dissociation is 0.21 ± 0.01 eV, indicating that virtually all of Erev is partitioned as translational Kinetic energy of the fragmentation products. The metastable ion peak shape is also discussed with respect to the evaluation of the energy release.  相似文献   

8.
The deuterium kinetic isotope effect and the deuterium isotope effect upon kinetic energy release have been calculated for the loss of H2 from the [C3H7]+ ion. The calculations are based on the transition state structure suggested recently from ab initio calculations on the reaction pathway. The results obtained are in good agreement with experimental data.  相似文献   

9.
We studied the attraction between [C2Hn] and Tl(I) in the hypothetical [C2Hn–Tl]+ complexes (n = 2,4) using ab initio methodology. We found that the changes around the equilibrium distance C–Tl and in the interaction energies are sensitive to the electron correlation potential. We evaluated these effects using several levels of theory, including Hartree–Fock (HF), second‐order Møller–Plesset (MP2), MP4, coupled cluster singles and doubles CCSD(T), and local density approximation augmented by nonlocal corrections for exchange and correlation due to Becke and Perdew (LDA/BP). The obtained interaction energies differences at the equilibrium distance Re (C–Tl) range from 33 and 46 kJ/mol at the different levels used. These results indicate that the interaction between olefinic systems and Tl(I) are a real minimum on the potential energy surfaces (PES). We can predict that these new complexes are viable for synthesizing. At long distances, the behavior of the [C2Hn]–Tl+ interaction may be related mainly to charge‐induced dipole and dispersion terms, both involving the individual properties of the olefinic π‐system and thallium ion. However, the charge‐induced dipole term (R?4) is found as the principal contribution in the stability at long and short distances. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2007  相似文献   

10.
A good understanding of gas‐phase fragmentation chemistry of peptides is important for accurate protein identification. Additional product ions obtained by sodiated peptides can provide useful sequence information supplementary to protonated peptides and improve protein identification. In this work, we first demonstrate that the sodiated a3 ions are abundant in the tandem mass spectra of sodium‐cationized peptides although observations of a3 ions have rarely been reported in protonated peptides. Quantum chemical calculations combined with tandem mass spectrometry are used to investigate this phenomenon by using a model tetrapeptide GGAG. Our results reveal that the most stable [a3 + Na ? H]+ ion is present as a bidentate linear structure in which the sodium cation coordinates to the two backbone carbonyl oxygen atoms. Due to structural inflexibility, further fragmentation of the [a3 + Na ? H]+ ion needs to overcome several relatively high energetic barriers to form [b2 + Na ? H]+ ion with a diketopiperazine structure. As a result, low abundance of [b2 + Na ? H]+ ion is detected at relatively high collision energy. In addition, our computational data also indicate that the common oxazolone pathway to generate [b2 + Na ? H]+ from the [a3 + Na ? H]+ ion is unlikely. The present work provides a mechanistic insight into how a sodium ion affects the fragmentation behaviors of peptides. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

11.
Preparation of Trimercaptosulfonium Salts [S(SH)3]+AsF6? and [S(SH)3]+SbCl6? The preparation of the trimercaptosulfonium salts [S(SH)3]+AsF6? and [S(SH)3]+SbCl6? from SCl3+ salts with excessive H2S at 193 K is reported. The [S(SH)3]+SbCl6? is transferred into [S(SCl)3]+SbCl6? by reaction with Cl2 at low temperatures. The new [S(SH)3]+ cation is isoelectronic to P(PH2)3. In addition, its existence is supported by an ab-initio calculation. The results show a potential well for C3v configuration with SH bonds bended towards the top of the pyramid for the isolated ion. Also the results of a force-field calculation are reported.  相似文献   

12.
13.
14.
15.
The geometries, successive binding energies, vibrational frequencies, and infrared intensities are calculated for the [Li(H2O)n]+ and [K(H2O)n]+ (n = 1?4) complexes. The basis sets used are 6-31G* and LANL 1DZ (Los Alamos ECP +DZ ) at the SCF and MP 2 levels. There is an agreement for calculated structures and frequencies between the MP 2/6-31G* and MP 2/LANL 1DZ basis sets, which indicates that the latter can be used for calculations of water complexes with heavier ions. Our results are in a reasonable agreement with available experimental data and facilitate experimental study of these complexes. © 1995 John Wiley & Sons, Inc.  相似文献   

16.
DFT investigations are carried out to explore the effective catalyst forms of DBU and H2O and the mechanism for the formation of 2,3‐dihydropyrido[2,3‐d]‐pyrimidin‐4(1H)‐ones. Three main pathways are disclosed under unassisted, water‐catalyzed, DBU and water cocatalyzed conditions, which involves concerted nucleophilic addition and H‐transfer, concerted intramolecular cyclization and H‐transfer, and Dimroth rearrangement to form the product. The results indicated that the DBU and water cocatalyzed pathway is the most favored one as compared to the rest two pathways. The water donates one H to DBU and accepts H from 2‐amino‐nicotinonitrile ( 1 ), forming [DBU‐H]+‐H2O as effective catalyst form in the proton migration transition state rather than [DBU‐H]+‐OH?. The hydrogen bond between [DBU‐H]+···H2O··· 1 ? decreases the activation barrier of the rate‐determining step. Our calculated results open a new insight for the green catalyst model of DBU‐H2O. © 2015 Wiley Periodicals, Inc.  相似文献   

17.
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   

18.
The unimolecular chemistry and structures of self‐assembled complexes containing multiple alkaline‐earth‐metal dications and deprotonated GlyGly ligands are investigated. Singly and doubly charged ions [Mn(GlyGly?H)n‐1]+ (n=2–4), [Mn+1(GlyGly?H)2n]2+ (n=2,4,6), and [M(GlyGly?H)GlyGly]+ were observed. The losses of 132 Da (GlyGly) and 57 Da (determined to be aminoketene) were the major dissociation pathways for singly charged ions. Doubly charged Mg2+ clusters mainly lost GlyGly, whereas those containing Ca2+ or Sr2+ also underwent charge separation. Except for charge separation, no loss of metal cations was observed. Infrared multiple photon dissociation spectra were the most consistent with the computed IR spectra for the lowest energy structures, in which deprotonation occurs at the carboxyl acid groups and all amide and carboxylate oxygen atoms are complexed to the metal cations. The N?H stretch band, observed at 3350 cm?1, is indicative of hydrogen bonding between the amine nitrogen atoms and the amide hydrogen atom. This study represents the first into large self‐assembled multimetallic complexes bound by peptide ligands.  相似文献   

19.
Unusual ionization behavior was observed with novel antineoplastic curcumin analogues during the positive ion mode of matrix‐assisted laser desorption ionization (MALDI) and dopant‐free atmospheric pressure photoionization (APPI). The tested compounds produced an unusual significant peak designated as [M ? H]+ ion along with the expected [M + H]+ species. In contrast, electrospray ionization, atmospheric pressure chemical ionization and the dopant‐mediated APPI (dopant‐APPI) showed only the expected [M + H]+ peak. The [M ? H]+ ion was detected with all evaluated curcumin analogues including phosphoramidates, secondary amines, amides and mixed amines/amides. Our experiments revealed that photon energy triggers the ionization of the curcumin analogues even in the absence of any ionization enhancer such as matrix, solvent or dopant. The possible mechanisms for the formation of both [M ? H]+ and [M + H]+ ions are discussed in this paper. In particular, three proposed mechanisms for the formation of [M ? H]+ were evaluated. The first mechanism involves the loss of H2 from the protonated [M + H]+ species. The other two mechanisms include hydrogen transfer from the analyte radical cation or hydride abstraction from the neutral analyte molecule. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

20.
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