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 gas‐phase interactions of cysteine with di‐organotin and tri‐organotin compounds have been studied by mass spectrometry experiments and quantum calculations. Positive‐ion electrospray spectra show that the interaction of di‐ and tri‐organotins with cysteine results in the formation of [(R)2Sn(Cys‐H)]+ and [(R)3Sn(Cys)]+ ions, respectively. MS/MS spectra of [(R)2Sn(Cys‐H)]+ complexes are characterized by numerous fragmentation processes, notably associated with elimination of NH3 and (C,H2,O2). Several dissociation routes are characteristic of each given organic species. Upon collision, both the [(R)3Sn(Gly)]+ and [(R)3Sn(Cys)]+ complexes are associated with elimination of the intact amino acid, leading to the formation of [(R)3Sn]+ cation. But for the latter complex, two additional fragmentation processes are observed, associated with the elimination of NH3 and C3H4O2S. Calculations indicate that the interaction between organotins and cysteine is predominantly electrostatic but also exhibits a considerable covalent character, which is slightly more pronounced in tri‐organotin complexes. A preferred bidentate interaction of the type ‐η2‐S‐NH2, with sulfur and the amino group, is observed. As for the [(R)3Sn(Cys)]+ complexes, their stability is due to the combination of the hydrogen bond taking place between the amino group and the sulfur lone pair and the interaction between the carboxylic oxygen atom and the metal. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

2.
Electrospray ionization mass spectrometry (ESI/MS) has allowed the discovery of novel dimer ions emerging from solutions of metalloporphyrin salts and their investigation by collision‐induced dissociation (CID) with N2 molecules. ESI mass spectra have been recorded for the formation of the oxygen or chloride‐bridged dimer ions [(FeTPP)2OH]+, [(MnTPP)2OH]+, [(FeTPP)2Cl]+ and [(MnTPP)2Cl]+ derived from various solutions of FeTPPCl and MnTPPCl salts. The CID of [(FeTPP)2OH]+ proceeds mainly by neutral loss of (FeTPP)OH to form [FeTPP]+ and, to a minor extent, to form the charge‐reversed products. The CID of [(MnTPP)2OH]+ exhibits exclusively the product ion [MnTPP]+ by loss of neutral (MnTPP)OH. [(FeTPP)2Cl]+ and [(MnTPP)2Cl]+ dissociate by loss of (Fe/MnTPP)Cl to give rise to [Fe/MnTPP]+. [(FeTPP)2O]+ and [(FeTPP)2OH]+ were generated from a solution of the dimer, (FeTPP)2O. Dissociation of [(FeTPP)2O]+ yields two product ions, [FeTPP]+ and [(FeTPP)O]+, with higher onsets compared to the equivalent fragments formed from [(FeTPP)2OH]+. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

3.
Four pairs of positional isomers of ureidopeptides, FmocNH‐CH(R1)‐φ(NH‐CO‐NH)‐CH(R2)‐OY and FmocNH‐CH(R2)‐φ(NH‐CO‐NH)‐CH(R1)‐OY (Fmoc = [(9‐fluorenyl methyl)oxy]carbonyl; R1 = H, alkyl; R2 = alkyl, H and Y = CH3/H), have been characterized and differentiated by both positive and negative ion electrospray ionization (ESI) ion‐trap tandem mass spectrometry (MS/MS). The major fragmentation noticed in MS/MS of all these compounds is due to ? N? CH(R)? N? bond cleavage to form the characteristic N‐ and C‐terminus fragment ions. The protonated ureidopeptide acids derived from glycine at the N‐terminus form protonated (9H‐fluoren‐9‐yl)methyl carbamate ion at m/z 240 which is absent for the corresponding esters. Another interesting fragmentation noticed in ureidopeptides derived from glycine at the N‐terminus is an unusual loss of 61 units from an intermediate fragment ion FmocNH = CH2+ (m/z 252). A mechanism involving an ion‐neutral complex and a direct loss of NH3 and CO2 is proposed for this process. Whereas ureidopeptides derived from alanine, leucine and phenylalanine at the N‐terminus eliminate CO2 followed by corresponding imine to form (9H‐fluoren‐9‐yl)methyl cation (C14H11+) from FmocNH = CHR+. In addition, characteristic immonium ions are also observed. The deprotonated ureidopeptide acids dissociate differently from the protonated ureidopeptides. The [M ? H]? ions of ureidopeptide acids undergo a McLafferty‐type rearrangement followed by the loss of CO2 to form an abundant [M ? H ? Fmoc + H]? which is absent for protonated ureidopeptides. Thus, the present study provides information on mass spectral characterization of ureidopeptides and distinguishes the positional isomers. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

4.
We have investigated gas‐phase fragmentation reactions of protonated benzofuran neolignans (BNs) and dihydrobenzofuran neolignans (DBNs) by accurate‐mass electrospray ionization tandem and multiple‐stage (MSn) mass spectrometry combined with thermochemical data estimated by Computational Chemistry. Most of the protonated compounds fragment into product ions B ([M + H–MeOH]+), C ([ B –MeOH]+), D ([ C –CO]+), and E ([ D –CO]+) upon collision‐induced dissociation (CID). However, we identified a series of diagnostic ions and associated them with specific structural features. In the case of compounds displaying an acetoxy group at C‐4, product ion C produces diagnostic ions K ([ C –C2H2O]+), L ([ K –CO]+), and P ([ L –CO]+). Formation of product ions H ([ D –H2O]+) and M ([ H –CO]+) is associated with the hydroxyl group at C‐3 and C‐3′, whereas product ions N ([ D –MeOH]+) and O ([ N –MeOH]+) indicate a methoxyl group at the same positions. Finally, product ions F ([ A –C2H2O]+), Q ([ A –C3H6O2]+), I ([ A –C6H6O]+), and J ([ I –MeOH]+) for DBNs and product ion G ([ B –C2H2O]+) for BNs diagnose a saturated bond between C‐7′ and C‐8′. We used these structure‐fragmentation relationships in combination with deuterium exchange experiments, MSn data, and Computational Chemistry to elucidate the gas‐phase fragmentation pathways of these compounds. These results could help to elucidate DBN and BN metabolites in in vivo and in vitro studies on the basis of electrospray ionization ESI‐CID‐MS/MS data only.  相似文献   

5.
Electrospray ionization of dilute aqueous solutions of copper(II) chloride‐containing traces of pyridine (py) as well as ammonia permits the generation of the gaseous ions (py)2Cu+ and (py)2CuCl+, of which the latter is a formal copper(II) compound, whereas the former contains copper(I). Collision‐induced dissociation of the mass‐selected ions in an ion‐trap mass spectrometer (IT‐MS) leads to a loss of pyridine from (py)2Cu+, whereas an expulsion of atomic chlorine largely prevails for (py)2CuCl+. Theoretical studies using density functional theory predict a bond dissociation energy (BDE) of BDE[(py)2Cu+ ‐Cl] = 125 kJ mol?1, whereas the pyridine ligand is bound significantly stronger, i.e. BDE[(py)CuCl+ ‐py] = 194 kJ mol?1 and BDE[(py)Cu+ ‐py] = 242 kJ mol?1. The results are discussed with regard to the influence of the solvation on the stability of the CuI/CuII redox couple. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

6.
Absolute bond dissociation energies of water to sodium glycine cations and glycine to hydrated sodium cations are determined experimentally by competitive collision-induced dissociation (CID) of Na+Gly(H2O)x, x = 1–4, with xenon in a guided ion beam tandem mass spectrometer. The cross sections for CID are analyzed to account for unimolecular decay rates, internal energy of reactant ions, multiple ion–molecule collisions, and competition between reaction channels. Experimental results show that the binding energies of water and glycine to the complexes decrease monotonically with increasing number of water molecules. Ab initio calculations at four different levels show good agreement with the experimental bond energies of water to Na+Gly(H2O)x, x = 0–3, and glycine to Na+(H2O), whereas the bond energies of glycine to Na+(H2O)x, x = 2–4, are systematically higher than the experimental values. These discrepancies may provide some evidence that these Na+Gly(H2O)x complexes are trapped in excited state conformers. Both experimental and theoretical results indicate that the sodiated glycine complexes are in their nonzwitterionic forms when solvated by up to four water molecules. The primary binding site for Na+ changes from chelation at the amino nitrogen and carbonyl oxygen of glycine for x = 0 and 1 to binding at the C terminus of glycine for x = 2–4. The present characterization of the structures upon sequential hydration indicates that the stability of the zwitterionic form of amino acids in solution is a consequence of being able to solvate all charge centers.  相似文献   

7.
Three new organotin(IV) carboxylates of composition [(R)2Sn(O2CC6H4C‐OC6H4CH3)2] 1 , [(R)2Sn(O2CC6H4COC6H4‐C2H5)2]2 2 , and R3SnO2 CC6H4COC6H4CH3 3 , were obtained by reactions of (R)2SnO [R = o‐tolyl] with 2‐(4‐methyl benzoyl) benzoic acid and 2‐(4‐ethyl benzoyl) benzoic acid, and reaction of Cy3SnOH [Cy = cyclohexyl] with 2‐(4‐methyl benzoyl) benzoic acid, respectively. The complexes 1 , 2 and 3 have been characterized by elemental analysis, infrared (IR), 1HNMR, and X‐ray crystallography diffraction analyses. The complex 1 has two folded symmetrical structure; the tin atom in 1 is found to adopt a distorted toward skew‐trapezoidal bipyramidal geometry. Molecular structure of the complex 2 is centrosymmetric, and the internal tin atom is five‐coordinated and is in distorted trigonal bipyramidal geometry. The crystal structure of the complex 3 is found to exhibit distorted tetrahedral geometry. Pilot studies have indicated that the complexes 1 and 2 have shown good antitumor activities. © 2010 Wiley Periodicals, Inc. Heteroatom Chem 21:304–313, 2010; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20614  相似文献   

8.
1H and 119Sn NMR results indicate that, when Ph3SnOH is dissolved in CD2Cl2, it dehydrates to (Ph3Sn)2O, only a small amount of Ph3SnOH remaining in equilibrium at room temperature. As a result, the reaction of TiCl4 with Ph3SnOH in CH2Cl2 proceeds via hydrolysis of the halide to precipitate amorphous TiO2 that contains adsorbed organotin species. Calcination of the amorphous precursor to 723 K yields nanoparticles of tin‐doped TiO2 photocatalysts, that contain anatase and rutile phases, and may also contain a segregated SnO2 phase. The reaction conditions that lead to the formation of a SnO2 phase have been studied and we have found that it is formed when the amorphous precipitate is not thoroughly washed with CH2Cl2 or when non‐recrystallized commercial Ph3SnOH is used as a starting material. The catalysts obtained have a high activity for the photooxidation of toluene in the gas phase. In particular, a material obtained from non‐recrystallized Ph3SnOH is particularly promising because the toluene photooxidation rate is more than twice as high as when using Degussa P25. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

9.
The structure of the title compound, 2C4H9N2O3+·C2O42−, which has been determined by X‐ray diffraction, contains discrete glycyl­glycine (HGly–Gly)+ cations in general positions and oxalate anions which lie across centres of inversion. Although the geometry of the (HGly–Gly)+ cation is not significantly different compared with other structures containing this residue, a few changes in conformation are observed which indicate the presence of molecular interactions. The molecular network in the crystal consists of one nearly linear O—H⋯O, five N—H⋯O and two weak C—H⋯O hydrogen bonds.  相似文献   

10.
The effect of triorganotin compounds, R3SnX, on the growth of three wild strains of Ceratocystis ulmi (C. ulmi) fungus, two aggressive and one non-aggressive strains, was evaluated in shake culture. In all cases, the triphenyltins were the more effective organotins for the inhibition of C. ulmi in vitro. The anionic group, X, did not have a significant role in the inhibition, suggesting that the species involved in the inhibition is the triphenyltin moiety (Ph3Sn+) or the hydrated triphenyltin moiety (Ph3Sn(H2O)+2). It is further suggested that the triphenyltin species Ph3SnOH and Ph3SnOAc are the preferred compounds for the control of Dutch elm disease. The tolerance of aggressive isolates to fungitoxins appears to depend more on the nature of the fungicide than on the type of fungus.  相似文献   

11.
Ion/molecule reactions of saturated hydrocarbons (n‐hexane, cyclohexane, n‐heptane, n‐octane and isooctane) in 28‐Torr N2 plasma generated by a hollow cathode discharge ion source were investigated using an Orbitrap mass spectrometer. It was found that the ions with [M+14]+ were observed as the major ions (M: sample molecule). The exact mass analysis revealed that the ions are nitrogenated molecules, [M+N]+ formed by the reactions of N3+ with M. The reaction, N3+ + M → [M+N]+ + N2, were examined by the density functional theory calculations. It was found that N3+ abstracts the H atom from hydrocarbon molecules leading to the formation of protonated imines in the forms of R′R″C?NH2+ (i.e. C–H bond nitrogenation). This result is in accord with the fact that elimination of NH3 is the major channel for MS/MS of [M+N]+. That is, nitrogen is incorporated in the C–H bonds of saturated hydrocarbons. No nitrogenation was observed for benzene and acetone, which was ascribed to the formation of stable charge‐transfer complexes benzene????N3+ and acetone????N3+ revealed by density functional theory calculations. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

12.
Spin‐labeled nitroxide derivatives of podophyllotoxin had better antitumor activity and less toxicity than that of the parent compounds. However, the 2‐H configurations of these spin‐labeled derivatives cannot be determined by nuclear magnetic resonance (NMR) methods. In the present paper, a high‐performance liquid chromatography‐diode array detection (HPLC‐DAD) and a high‐performance liquid chromatography‐electrospray ionization tandem mass spectrometry (HPLC‐ESI/MS/MS) method were developed and validated for the separation, identification of four pairs of diastereoisomers of spin‐labeled derivatives of podophyllotoxin at C‐2 position. In the HPLC‐ESI/MS spectra, each pair of diastereoisomers of the spin‐labeled derivatives in the mixture was directly confirmed and identified by [M+H]+ ions and ion ratios of relative abundance of [M‐ROH+H]+ (ion 397) to [M+H]+. When the [M‐ROH+H]+ ions (at m/z 397) were selected as the precursor ions to perform the MS/MS product ion scan. The product ions at m/z 313, 282, and 229 were the common diagnostic ions. The ion ratios of relative abundance of the [M‐ROH+H]+ (ion 397) to [M+H]+, [A+H]+ (ion 313) to [M‐ROH+H]+, [A+H‐OCH3]+ (ion 282) to [M‐ROH+H]+ and [M‐ROH‐ArH+H]+ (ion 229) to [M‐ROH+H]+ of each pair of diastereoisomers of the derivatives specifically exhibited a stereochemical effect. Thus, by using identical chromatographic conditions, the combination of DAD and MS/MS data permitted the separation and identification of the four pairs of diastereoisomers of spin‐labeled derivatives of podophyllotoxin at C‐2 in the mixture.  相似文献   

13.
In the solid state, the title compound, di‐μ‐hydroxo‐1:2κ2O;‐3:4κ2O‐dihydroxo‐1κO,4κO‐octakis(2‐methyl‐2‐phenyl­propyl)‐1κ2C,2κ2C,3κ2C,4κ2C‐di‐μ3‐oxo‐1:2:3κ3O;2:3:4κ3O‐tetratin(IV), [Sn4O2(OH)4(C10H13)8], forms centrosymmetric dimeric [(Neophyl2SnOH)(Neophyl2SnOH)O]2 mol­ecules (Neophyl = 2‐methyl‐2‐phenylpropyl), with an almost planar Sn–O framework that adopts a ladder‐type structure consisting of three four‐membered rings. The hydroxyl groups are shielded by the organic groups, which prevent them from further condensation and from the formation of hydrogen bonds.  相似文献   

14.
Methylation is one of the important posttranslational modifications of biological systems. At the metabolite level, the methylation process is expected to convert bioactive compounds such as amino acids, fatty acids, lipids, sugars, and other organic acids into their methylated forms. A few of the methylated amino acids are identified and have been proved as potential biomarkers for several metabolic disorders by using mass spectrometry–based metabolomics workstation. As it is possible to encounter all the N‐methyl forms of the proteinogenic amino acids in plant/biological systems, it is essential to have analytical data of all N‐methyl amino acids for their detection and identification. In earlier studies, we have reported the ESI‐MS/MS data of all methylated proteinogenic amino acids, except that of mono‐N‐methyl amino acids. In this study, the N‐methyl amino acids of all the amino acids ( 1 ‐ 21 ; including one isomeric pair) were synthesized and characterized by ESI‐MS/MS, LC/MS/MS, and HRMS. These data could be useful for detection and identification of N‐methyl amino acids in biological systems for future metabolomics studies. The MS/MS spectra of [M + H]+ ions of most N‐methyl amino acids showed respective immonium ions by the loss of (H2O, CO). The other most common product ions detected were [MH‐(NH2CH3]+, [MH‐(RH)]+ (where R = side chain group) ions, and the selective structure indicative product ions due to side chain and N‐methyl group. The isomeric/isobaric N‐methyl amino acids could easily be differentiated by their distinct MS/MS spectra. Further, the MS/MS of immonium ions inferred side chain structure and methyl group on α‐nitrogen of the N‐methyl amino acids.  相似文献   

15.
A simple method was developed for the generation of cesium iodide (CsI) cluster ions up to m/z over 20,000 in matrix-assisted laser desorption/ionization mass spectrometry (MALDI MS). Calibration ions in both positive and negative ion modes can readily be generated from a single MALDI spot of CsI3 with 2-[(2E)-3-(4-tert-butylphenyl)-2-methylprop-2-enylidene] malononitrile (DCTB) matrix. The major cluster ion series observed in the positive ion mode is [(CsI)nCs]+, and in the negative ion mode is [(CsI)nI]. In both cluster series, ions spread evenly every 259.81 units. The easy method described here for the production of CsI cluster ions should be useful for MALDI MS calibrations.  相似文献   

16.
2‐[(2‐Ammonioethyl)amino]acetate dihydrate, better known as N‐(2‐aminoethyl)glycine dihydrate, C4H10N2O2·2H2O, (I), crystallizes as a three‐dimensional hydrogen‐bonded network. Amino acid molecules form layers in the ac plane separated by layers of water molecules, which form a hydrogen‐bonded two‐dimensional net composed of fused six‐membered rings having boat conformations. The crystal structure of the corresponding hydroiodide salt, namely 2‐[(2‐ammonioethyl)ammonio]acetate iodide, C4H11N2O2+·I, (II), has also been determined. The structure of (II) does not accommodate any solvent water molecules, and displays stacks of amino acid molecules parallel to the a axis, with iodide ions located in channels, resulting in an overall three‐dimensional hydrogen‐bonded network structure. N‐(2‐Aminoethyl)glycine is a molecule of considerable biological interest, since its polyamide derivative forms the backbone in the DNA mimic peptide nucleic acid (PNA).  相似文献   

17.
The reaction of the 2,2‐bis(organodichlorostannyl)propane [(Me3Si)2CH(Cl2)Sn]2CMe2 (A) with the corresponding organotin oxide {[(Me3Si)2CH(O)Sn]2CMe2}2 (B) does not provide the corresponding normally expected tetraorganodistannoxane {[(Me3Si)2CH(Cl)SnCMe2Sn(Cl)CH(SiMe3)2]O}n but a complex reaction mixture. One major product, namely the 2,4,6,8‐tetraorgano‐2,6‐dichloro‐1,5,9‐trioxa‐2,4,6,8‐tetrastannabicyclo[3.3.1]nonane derivative [(Me3Si)2CHSnCMe2Sn(Cl)CH(SiMe3)2]2O3 (C) was identified in situ by 2D 1H? 119Sn and 1H? 13C heteronuclear multiple quantum coherence and heteronuclear multiple bond correlation NMR spectroscopy as well as electrospray mass spectrometry. Compound C is proposed to be in equilibrium with an ionic species C′, the cation of which has an adamantane‐type structure. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

18.
The nitrate anion coordinates to the Sn? CH2? Sn unit of the title phosphonium stannate, [Ph4P]+ [(Ph2ClSn)2CH2 ·NO3]?, to give a six‐membered ring having the penta‐coordinated tin atoms in a trigonal bipyramidal geometry. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

19.
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   

20.
The reduction of digallane [(dpp‐bian)Ga? Ga(dpp‐bian)] ( 1 ) (dpp‐bian=1,2‐bis[(2,6‐diisopropylphenyl)imino]acenaphthene) with lithium and sodium in diethyl ether, or with potassium in THF affords compounds featuring the direct alkali metal–gallium bonds, [(dpp‐bian)Ga? Li(Et2O)3] ( 2 ), [(dpp‐bian)Ga? Na(Et2O)3] ( 3 ), and [(dpp‐bian)Ga? K(thf)5] ( 7 ), respectively. Crystallization of 3 from DME produces compound [(dpp‐bian)Ga? Na(dme)2] ( 4 ). Dissolution of 3 in THF and subsequent crystallization from diethyl ether gives [(dpp‐bian)Ga? Na(thf)3(Et2O)] ( 5 ). Ionic [(dpp‐bian)Ga]?[Na([18]crown‐6)(thf)2]+ ( 6 a ) and [(dpp‐bian)Ga]?[Na(Ph3PO)3(thf)]+ ( 6 b ) were obtained from THF after treatment of 3 with [18]crown‐6 and Ph3PO, respectively. The reduction of 1 with Group 2 metals in THF affords [(dpp‐bian)Ga]2M(thf)n (M=Mg ( 8 ), n=3; M=Ca ( 9 ), Sr ( 10 ), n=4; M=Ba ( 11 ), n=5). The molecular structures of 4 – 7 and 11 have been determined by X‐ray crystallography. The Ga? Na bond lengths in 3 – 5 vary notably depending on the coordination environment of the sodium atom.  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