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 15 毫秒
1.
More than 310 kinds of cluster ions of S(m) P(n) H(k) (k+) are observed in a single ESI mass spectrum of a mixed solution of serine and phosphoric acid. Some typical cluster ions are selected, activated by collision in a FT ICR cell, and the dissociation pathways were deduced in detail. For large singly protonated ions, the collisions cause the ejection of subunits of serine or phosphoric acid subsequently producing the ions of S(2) P(4) H(1) (1+) , which can be further dissociated by the loss of phosphoric acid molecules in turn and form the protonated serine dimer and monomer. However, for the doubly protonated ions, the dissociation pathways change from the loss of a protonated serine dimer for the ions of S(7) P(9) H(2) (2+) to the neutral loss of H(3) PO(4) for the ions of S(7) P(12) H(2) (2+) or the neutral loss of serine or H(3) PO(4) for the larger clusters, indicating the effect of cluster sizes on the process of dissociation. The structure of S(2) P(4) H(1) (1+) is suggested based on B3LYP/6-31G(d,p) calculations. The diversity and structural orderliness of the hetero-cluster ions are mainly attributed to the network of hydrogen bonds inside the cluster ions and the extraordinary amphotericity of the components.  相似文献   

2.
The gas-phase coordination behavior of the [Cr(III)(Salen)]PF(6) complex at the free axial positions has been studied in the presence of amines as ligands (propylamine and a series of diamines) under electrospray ionization conditions. The [Cr(III)(Salen)](+) complex formed stable five- and six-coordinated complex ions, [Cr(III)(Salen)(L)](+) and [Cr(III)(Salen)(L)(2)](+), respectively, where L = solvent molecule or amine. When diamines were used as ligands, abundant [Cr(III)(Salen)(L)](+) ions were observed in which two axial positions of the [Cr(III)(Salen)](+) species are occupied by the two amino groups of the diamine ligand. The relative abundances of ligated complex ions, fragment ions, and solvent adducts of fragment ions in the ESI mass spectra, were found to depend on the cone voltage used to record the spectrum. The ESI mass spectra of [Cr(III)(Salen)](+) in the presence of diamines as ligands, and experiments on ligand-pickup in the collision cell, clearly demonstrated that the [Cr(III)(Salen)(L)](+) complex ion is stable for 1,2-diaminoethane and 1,3-diaminopropane. The stability of [Cr(III)(Salen)(L)](+) ions gradually decreased from 1,4-diaminobutane to 1,6-diaminohexane, and then showed a slight increase for 1,7-diaminoheptane and 1,8-diaminooctane. The collision-induced dissociation spectra of [Cr(III)(Salen)(L)](+) ions support the above observations.  相似文献   

3.
We have examined interactions between mitoxantrone (MXT) and DNA duplexes or triplexes with different base compositions by using electrospray ionization mass spectrometry (ESI‐MS), respectively. MXT interacts preferentially with DNA duplexes compared to the triplexes. In the mass spectrum of the duplex–MXT mixture, the complex peaks dominated in the ratios of duplex/MXT of 1:1, 1:2 and 1:3, and the 1:2 duplex/MXT peak was the most abundant. In contrast, only 1:1 triplex–MXT complexes were observed in the mass spectrum of the triplex–MXT mixture, and the most intensive peak was a free triplex ion without MXT. Moreover, no sequence selectivity of MXT to different DNA duplexes was found while MXT showed greater affinity to the triplexes that have adjacent TAT or C+GC sequences. In the course of sustained off‐resonance irradiation collision‐induced dissociation (SORI‐CID), the MXT‐duplex complexes generated two separated strands, and the MXT remained on the purine strand side. UV/Vis spectra showed that MXT interacted with DNA by intercalation. Compared with emodin (a duplex intercalator) and napthylquinoline (a triplex binder), we found that the side chain of MXT might play a role in the binding of MXT to the duplexes and the triplexes. ESI‐MS shows an advantage in speed and straightforwardness for the study of drug interactions with nucleic acids. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

4.
Sun X  Jin C  Mei Y  Yang G  Guo Z  Zhu L 《Inorganic chemistry》2004,43(1):290-296
Interactions of cis-[Pd(en)(H(2)O)(2)](2+) (en, ethylenediamine) and cis-[Pt(NH(3))(2)(H(2)O)(2)](2+) with microperoxidase-11 (MP-11) in a molar ratio of 1:1 or 2:1 at pH 1.4 were investigated via electrospray mass spectrometry and MS/MS analysis at room temperature and at 40 degrees C with an incubation time of 2 or 3 days. The composition of the Pd(II)- and Pt(II)-anchored MP-11 was confirmed on the basis of the precise molecular mass and the simulated isotope distribution pattern. MS/MS analysis revealed that the Pd(II) center anchored to the side chain of Cys7 as Pd(II) and MP-11 were mixed in an equimolar ratio and to side chains of Cys7 and Cys4 as Pd(II) and MP-11 mixed in a 2:1 molar ratio. When Pt(II) and MP-11 were mixed in a 2:1 molar ratio, Pt(II) first anchored to the side chain of Cys7, and then to the side chain of Cys4 with time. The initial coordination of Pd(II) and Pt(II) to the side chain of Cys7 is the essential step for the Pd(II)- and Pt(II)-promoted cleavage of the His8-Thr9 bond in MP-11. These results support the hypothesis that the Pd(II)-mediated cleavage of the His18-Thr19 bond in cytochorome c is due to the identical binding mode.  相似文献   

5.
Electrospray ionization tandem mass spectrometry (ESI-MS(n)) and the phase solubility method were used to characterize the gas-phase and solution-phase non-covalent complexes between rutin (R) and alpha-, beta- and gamma-cyclodextrins (CDs). The direct correlation between mass spectrometric results and solution-phase behavior is thus revealed. The order of the 1 : 1 association constants (K(c)) of the complexes between R and the three CDs in solution calculated from solubility diagrams is in good agreement with the order of their relative peak intensities and relative collision-induced dissociation (CID) energies of the complexes under the same ESI-MS(n) condition in both the positive and negative ion modes. Not only the binding stoichiometry but also the relative stabilities and even binding sites of the CD-R complexes can be elucidated by ESI-MS(n). The diagnostic fragmentation of CD-R complexes, with a significant contribution of covalent fragmentation of rutin leaving the quercetin (Q) moiety attached to the CDs, provides convincing evidence for the formation of inclusion complexes between R and CDs. The diagnostic fragment ions can be partly confirmed by the complexes between Q and CDs. The gas-phase stability order of the deprotonated CD-R complexes is beta-CD-R > alpha-CD-R > gamma-CD/R; beta-CD seems to bind R more strongly than the other CDs.  相似文献   

6.
Ageing products of a commercial jet engine oil based on pentaerythritol tetraesters which were formed upon operation in an aviation turbine were detected by electrospray ionization mass spectrometry (ESI-MS) and characterized by LC-ESI-MS. The fatty acid composition of these ageing products was investigated by ESI-MS-MS analysis. The ammonium adducts of the newly formed pentaerythritol tetraester degradation products were found to be suitable parent ions for further structure elucidation work. ESI-MS, LC-ESI-MS and ESI-MS-MS proved to be versatile tools to study the chemical composition (distribution of homologues) as well as the mechanism of ageing of ester based lubricants on a molecular level. Due to its high sensitivity, ESI-MS can also be used to characterize and identify trace levels of ester-based lubricants.  相似文献   

7.
A new class of diastereomeric pairs of non‐natural amino acid peptides derived from butyloxycarbonyl (Boc‐)protected cis‐(2S,3R)‐ and trans‐(2S,3S)‐β‐norbornene amino acids including a monomeric pair have been investigated by electrospray ionization (ESI) tandem mass spectrometry using quadrupole time‐of‐flight (Q‐TOF) and ion‐trap mass spectrometers. The protonated cis‐BocN‐β‐nbaa (2S,3R) (1) (βnbaa = β‐norbornene amino acid) eliminates the Boc group to form [M+H–Boc+H]+, whereas an additional ion [M+H–C4H8]+ is formed from trans‐BocN‐β‐nbaa (2S,3S) (2). Similarly, it is observed that the peptide diastereomers (di‐, tri‐ and tetra‐), with cis‐BocN‐β‐nbaa (2S,3R)‐ at the N‐terminus, initially eliminate the Boc group to form [M+H–Boc+H]+ which undergo further fragmentation to give a set of product ions that are different for the peptides with trans‐BocN‐β‐nbaa (2S,3S)‐ at the N‐terminus. Thus the Boc group fragments differently depending on the configuration of the amino acid present at the N‐terminus. It is also observed that the peptide bond cleavage in these peptides is less favoured and most of the product ions are formed due to retro‐Diels‐Alder fragmentation. Interestingly, sodium‐cationized peptide diastereomers mainly yield a series of retro‐Diels‐Alder fragment ions which are different for each diastereomer as they are formed starting from [M+Na–Boc+H]+ in peptides with cis‐BocN‐β‐nbaa (2S,3R)‐ at the N‐terminus, and [M+Na–C4H8]+ in peptides with trans‐BocN‐β‐nbaa (2S,3S)‐ at the N‐terminus. All these results clearly indicate that these diastereomeric pairs of peptides yield characteristic product ions which help distinguish the isomers. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
Oligoribonucleotides (RNA) and modified oligonucleotides were subjected to low-energy collision-induced dissociation in a hybrid quadrupole time-of-flight mass spectrometer to investigate their fragmentation pathways. Only very restricted data are available on gas-phase dissociation of oligoribonucleotides and their analogs and the fundamental mechanistic aspects still need to be defined to develop mass spectrometry-based protocols for sequence identification. Such methods are needed, because chemically modified oligonucleotides can not be submitted to standard sequencing protocols. In contrast to the dissociation of DNA, dissociation of RNA was found to be independent of nucleobase loss and it is characterized by cleavage of the 5'-P-O bond, resulting in the formation of c- and their complementary y-type ions. To evaluate the influence of different 2'-substituents, several modified tetraribonucleotides were analyzed. Oligoribonucleotides incorporating a 2'-methoxy-ribose or a 2'-fluoro-ribose show fragmentation that does not exhibit any preferred dissociation pathway because all different types of fragment ions are generated with comparable abundance. To analyze the role of the nucleobases in the fragmentation of the phosphodiester backbone, an oligonucleotide lacking the nucleobase at one position has been studied. Experiments indicated that the dissociation mechanism of RNA is not influenced by the nucleobase, thus, supporting a mechanism where dissociation is initiated by formation of an intramolecular cyclic transition state with the 2'-hydroxyl proton bridged to the 5'-phosphate oxygen.  相似文献   

9.
Charged derivatives of peptides are useful in obtaining simpler collision-activated dissociation (CAD) mass spectra. An N-terminal charge-derivatizing reagent capable of reacting with picomole levels of peptide has been recently reported (Huang et al. Anal. Chem. 1997, 69, 137-144) in the contexts of analyses by fast atom bombardment (FAB) and matrix-assisted laser desorption/ionization (MALDI) mass spectrometry. Electrospray ionization (ESI) mass spectrometric investigation of these tris(trimethoxyphenylphosphonium) acetyl derivatives are described in this article, including studies by in-source fragmentation (ISF) and tandem mass spectrometry (MS/MS). Results from ISF are compared with those from MS/MS. Similarities and differences between ESI-ISF, MALDI-post-source decay (PSD), and FAB-CAD data are presented. Differences in fragmentation of these charged derivatives in the triple quadrupole and ion trap mass spectrometers also are discussed. Application of this derivatizing procedure to tryptic digests and subsequent analysis by liquid chromatography-mass spectrometry is also shown.  相似文献   

10.
The solutions containing one of the copper salts (CuCl2, Cu(ClO4)2, Cu(NO3)2, and CuSO4) and one of the non-steroidal anti-inflammatory drugs (NSAIDs, ibuprofen, ketoprofen or naproxen) were analyzed by electrospray ionization mass spectrometry. Three of the salts, namely CuCl2, Cu(ClO4)2 and Cu(NO3)2, yielded binuclear complexes of drug:metal stoichiometry 1:2. Existence of the complexes of such stoichiometry has not been earlier observed. For copper(II) chloride the complexes (ions of the type [M-HCOOH+Cu2Cl]+ and [M+Cu2Cl]+, M stands for the drug molecule) were formed in the gas phase. When copper(II) perchlorate or copper(II) nitrate was used, the observed binuclear copper complexes (ions of the type [M-H+Cu2(ClO4)2+CH3OH]+, [M-H+Cu2(ClO4)2]+ and [M-H+Cu2(NO3)2+CH3OH]+, [M-H+Cu2(NO3)2]+) were observed at low cone voltage, thus these complexes must have already existed in the solution analysed. Therefore, such complexes may also exist under physiological conditions.   相似文献   

11.
Presented is a method for analyzing sulfated peptides, and differentiating the post-translational modification (PTM) from its isobaric counterpart phosphorylation, using quadrupole time-of-flight (Qq/TOF) mass spectrometry (MS) and positive ion nanoelectrospray MS/MS. A set of commercially available sulfo- and phosphopeptide standards was analyzed via in-source dissociation and MS/MS to generate fragmentation signatures that were used to characterize and differentiate the two modifications. All of the phosphorylated peptides retained their +80 Da modifications under collision-induced decomposition (CID) conditions and peptide backbone fragmentation allowed for the site-specific identification of the modification. In sharp contrast, sulfated peptides lost SO3 from the precursor as the collision energy (CE) was increased until only the non-sulfated form of the peptide was observed. The number of 80 Da losses indicated the number of sulfated sites. By continuing to ramp the CE further, it was possible to fragment the non-sulfated peptides and obtain detailed sequence information. It was not possible to obtain site-specific information on the location of the sulfate moieties using positive ion MS/MS as none of the original precursor ions were present at the time of peptide backbone fragmentation. This method was applied to the analysis of recombinant human B-domain deleted factor VIII (BDDrFVIII), which has six well-documented sulfation sites and several potential phosphorylation sites located in two of the sulfated regions of the protein. Seven peptides with single and multiple +80 Da modifications were isolated and analyzed for their respective PTMs. The fragmentation patterns obtained from the BDDrFVIII peptides were compared with those obtained for the standard peptides; and in all cases the peptides were sulfated. None of the potential phosphorylation sites were found to be occupied, and these results are consistent with the literature.  相似文献   

12.
Electrospray ionization mass (ESI-MS) spectrometry was used to investigate the nature of metal complexes of alachlor and their dissociations on activation. Ions of the first row transition metal series were employed to react with alachlor and the products were subjected to collision-induced dissociation (CID) for further structural characterization. The formation of diverse complex ions including doubly charged metal/alachlor complexes; [3L + M]2+ and [4L + M]2+ (L: alachlor and M: transition metal ions) were observed depending on the experimental conditions including the tube lens offset voltage (TLOV) and relative concentrations of alachlor and transition metal ions. It is clear that complexation with transition metal ions alters the reactive site of alachlor, promoting the loss of chlorine over the loss of CH3OH that is the major reaction pathway in uncomplexed system. Direct elimination of chlorine from alachlor molecule was confirmed by the use of MnBr2 instead of MnCl2. These evidences clearly illustrate the catalytic activities of the metal ions through insertion mechanism. The function of transition metal ions in complexation was emphasized comparing the fragmentation patterns with those of protonated molecule. A change in the oxidation state of copper from + 2 to + 1 during the dissociation of metal complex was observed in company with elimination of radicals which is specific for the copper complex ions.  相似文献   

13.
14.
15.
Non-covalent interactions between met- and leu-enkephalins and their antisense peptides were studied by electrospray ionization mass spectrometry. Mixtures of sense and antisense peptides gave both the corresponding homodimers and heterodimers. The relative abundance ratios of the heterodimer to that of the homodimer of the sense peptide and the relative stability constants of the heterodimers were compared with the corresponding values from mixtures of the sense peptides and a control peptide. The results show that there is a preferential interaction between the sense and antisense peptides compared with that between the sense and control peptides.  相似文献   

16.
The presence of arginine as the naturally occurring amino acid with the highest gas-phase basicity strongly influences the fragmentation behavior of peptides undergoing collision-induced dissociation. Using a derivatization procedure recently developed in our group, based on a reversible reaction of the guanidino group with 2,3-butanedione and an arylboronic acid, we examined how this label affects the fragmentation patterns of labeled versus unlabeled peptides in MS/MS experiments. As part of this fundamental study, two groups of model peptides (angiotensins and bradykinins) as well as tryptic peptides were labeled according to our protocol and subjected to collision-induced dissociation (CID) in both a triple quadrupole and a quadrupole ion trap instrument. It was found that for angiotensins containing an AspArg sequence, C-terminal cleavage at Asp that occurs for native peptides was completely inhibited in Arg-labeled peptides. For bradykinins and peptides obtained from tryptic digests of standard proteins, some sample peptides were little affected by the tagging of arginine residues. Others, in contrast, exhibited an almost total loss of nonspecific backbone cleavage and their fragment ion spectra were dominated by loss of the arginine tag. These and other experimental results are discussed in view of the nature of the arginine tag and the concept of proton mobility.  相似文献   

17.
Alkaline metal adduct ions of sphingomyelin were formed by electrospray ionization in positive ion mode. Under low energy collisionally activated dissociation (CAD), the product ion spectra yield abundant fragment ions representative of both long chain base and fatty acid which permit unequivocal determination of the structure. Tandem spectra obtained by constant neutral loss scanning permit identification of sphingomyelin class and specific long chain base subclass in the mixture. The fragmentation pathways under CAD were proposed, and were further confirmed by source CAD tandem mass spectrometry. The total analysis of sphingomyelin mixtures from bovine brain, bovine erythrocytes, and chicken egg yolk is also presented.  相似文献   

18.
The composition and binding sites of cis-[Ru(II)(bpy)2]2+-bound sulfur-containing peptides of Met-Arg-Phe-Ala, glutathione and oxidized glutathione, and also histidine-containing peptide of oxidized insulin B chain, were investigated by electrospray ionization mass spectrometry (ESI-MS) and tandem mass spectrometry (MS/MS). The composition of Ru(II)-containing peptides was precisely determined by ESI-MS, zoom scan and simulation of isotope distribution patterns. MS/MS analysis shows that, in sulfur-containing peptides, the Ru(II) complex prefers to anchor to a carboxyl group, although some other potential binding sites of thiol, thioether and N-terminal amino groups present in these peptides, and in oxidized insulin B chain, Ru(II) first anchors to His10, then either to the hydroxyl group of Thr27 or to the carboxyl group of Ala30. Its secondary structure and microenvironment surrounding the potential binding sites may affect the binding ability of cis-[Ru(II)(bpy)2]2+ to oxidized insulin B chain.  相似文献   

19.
The electrospray ionization mass spectrometry and tandem mass spectrometry investigation showed that the binding sites of Zn^2+ with oxidized insulin B chain are His 5, His 10, and Arg 22, which lead to the selective cleavages of the peptide bonds at Ash 3- Gin 4, His 5-Leu 6, Gly 8-Ser 9, and Glu 21-Arg 22 of oxidized insulin B chain.  相似文献   

20.
We used solution-phase hydrogen/deuterium (H/D) exchange and multistage tandem mass spectrometry (MS/MS) experiments in an electrospray ion-trap mass spectrometer operating in the negative-ion mode to investigate the consequences of the loss of a high proton-affinity (PA) base from T-rich tetra and hexadeoxynucleotides. The T-rich oligodeoxynucleotides containing one or two other nucleobases take advantage of the mass spectral inertness of T because fragmentation of a T-rich oligomer is simple, allowing a tight focus on those processes of interest. Furthermore, determination of T-rich oligodeoxynucleotides may be a starting point in the development of a mass spectrometric scheme to understand the mutagenicity of various types of DNA damage by UV radiation. For nine oligodeoxynucleotides, the nucleobases were charged by nearly exclusive D transfer and then expelled as neutral bases. Loss of the base located at the 3' end is preferred over that from the 5' terminus when the two bases are identical. The observation of partially exchanged fragments from a completely exchanged precursor ion proves intramolecular H/D exchange between hydrogen atoms that can exchange in water and those that cannot. The multiplicity of the product-ion peaks provides information on decomposition pathways and origins of the product ions and shows that the loss of base is the first step in all fragmentation of hexanucleotides, but is a competitive process for tetranucleotide fragmentation.  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