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.
N‐Boc/Fmoc/Z‐N′‐formyl‐gem‐diaminoalkyl derivatives, intermediates particularly useful in the synthesis of partially modified retro‐inverso peptides, have been characterized by both positive and negative ion electrospray ionization (ESI) ion‐trap multi‐stage mass spectrometry (MSn). The MS2 collision induced dissociation (CID) spectra of the sodium adduct of the formamides derived from the corresponding N‐Fmoc/Z‐amino acids, dipeptide and tripeptide acids show the [M + Na‐NH2CHO]+ ion, arising from the loss of formamide, as the base peak. Differently, the MS2 CID spectra of [M + Na]+ ion of all the N‐Boc derivatives yield the abundant [M + Na‐C4H8]+ and [M + Na‐Boc + H]+ ions because of the loss of isobutylene and CO2 from the Boc protecting function. Useful information on the type of amino acids and their sequence in the N‐protected dipeptidyl and tripeptidyl‐N′‐formamides is provided by MS2 and subsequent MSn experiments on the respective precursor ions. The negative ion ESI mass spectra of these oligomers show, in addition to [M‐H]?, [M + HCOO]? and [M + Cl]? ions, the presence of in‐source CID fragment ions deriving from the involvement of the N‐protecting group. Furthermore, MSn spectra of [M + Cl]? ion of N‐protected dipeptide and tripeptide derivatives show characteristic fragmentations that are useful for determining the nature of the C‐terminal gem‐diamino residue. The present paper represents an initial attempt to study the ESI‐MS behavior of these important intermediates and lays the groundwork for structural‐based studies on more complex partially modified retro‐inverso peptides. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

2.
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   

3.
The microalga Haematococcus pluvialis produces the pigment astaxanthin mainly in esterified form with a multitude of fatty acids, which results in a complex mixture of carotenol mono‐ and diesters. For rapid fingerprinting of these esters, matrix‐assisted laser desorption ionization time of flight mass spectrometry (MALDI‐TOF/TOF‐MS) might be an alternative to traditional chromatographic separation combined with MS. Investigation of ionization and fragmentation of astaxanthin mono‐ and diester palmitate standards in MALDI‐TOF/TOF‐MS showed that sodium adduct parent masses [M + Na]+ gave much simpler MS2 spectra than radical / protonated [M]+● / [M + H]+ parents. [M + Na]+ fragments yielded diagnostic polyene‐specific eliminations and fatty acid neutral losses, whereas [M]+● / [M + H]+ fragmentation resulted in a multitude of non‐diagnostic daughters. For diesters, a benzonium fragment, formed by polyene elimination, was required for identification of the second fatty acid attached to the astaxanthin backbone. Parents were forced into [M + Na]+ ionization by addition of sodium acetate, and best signal‐to‐noise ratios were obtained in the 0.1 to 1.0 mM range. This method was applied to fingerprinting astaxanthin esters in a crude H. pluvialis extract. Prior to MALDI‐TOF/TOF‐MS, the extract was fractionated by normal phase Flash chromatography to obtain fractions enriched in mono‐ and diesters and to remove pheophytin a, which compromised monoester signals. All 12 types of all‐trans esterified esters found in LC were identified with MALDI‐TOF/TOF‐MS, with the exception of two minor monoesters. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

4.
We report the first positive chemical ionization (PCI) fragmentation mechanisms of phthalates using triple‐quadrupole mass spectrometry and ab initio computational studies using density functional theories (DFT). Methane PCI spectra showed abundant [M + H]+, together with [M + C2H5]+ and [M + C3H5]+. Fragmentation of [M + H]+, [M + C2H5]+ and [M + C3H5]+ involved characteristic ions at m/z 149, 177 and 189, assigned as protonated phthalic anhydride and an adduct of phthalic anhydride with C2H5+ and C3H5+, respectively. Fragmentation of these ions provided more structural information from the PCI spectra. A multi‐pathway fragmentation was proposed for these ions leading to the protonated phthalic anhydride. DFT methods were used to calculate relative free energies and to determine structures of intermediate ions for these pathways. The first step of the fragmentation of [M + C2H5]+ and [M + C3H5]+ is the elimination of [R? H] from an ester group. The second ester group undergoes either a McLafferty rearrangement route or a neutral loss elimination of ROH. DFT calculations (B3LYP, B3PW91 and BPW91) using 6‐311G(d,p) basis sets showed that McLafferty rearrangement of dibutyl, di(‐n‐octyl) and di(2‐ethyl‐n‐hexyl) phthalates is an energetically more favorable pathway than loss of an alcohol moiety. Prominent ions in these pathways were confirmed with deuterium labeled phthalates. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

5.
The purpose of this work is to explore the mass fragment characterization of commonly used drugs through a novel approach, which involves isotope‐selective tandem mass spectrometry (MS/MS). Collision‐induced dissociation (CID) was performed with a low‐resolution linear ion trap mass spectrometer in positive electrospray ionization. Three pharmacologically active ingredients, i.e. omeprazole, meloxicam and brinzolamide, selected as model compounds in their own formulation, were investigated as a sodiated adduct [C17H19N3O3S + Na]+ (omeprazole) and as protonated adducts, [C14H13N3O4S2 + H]+ and [C12H21N3O5S3 + H]+, meloxicam and brinzolamide, respectively. Selecting a narrow window of ±0.5 m/z units, precursor ion fragmentation by CID‐MS/MS of isotopologues A + 0, A + 1 and A + 2 was found very useful to confirm the chemical formula of product ions, thus aiding the establishment of characteristic fragmentation pathways of all three examined compounds. The correctness of putative molecular formula of product ions was easily demonstrated by exploiting the isotope peak abundance ratios (i.e. IF+0/IF+1 and IF+0/IF+2) as simple constraints in low‐resolution MS instrumentations. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

6.
Sesquiterpene lactones (SL) have been reported with various biological effects. Among the described SL skeletons, hirsutinolide and glaucolide have not been extensively studied by mass spectrometry (MS), especially how to distinguish them in organic matrices. Thus, this paper reports (1) a strategy of their differentiation based on MS behavior during the ionization and (2) a proposal of the fragmentation pattern for both SL‐subtypes. ESI(+)‐HRMS data of four isolated SL (hirsutinolides 1 and 3 ; glaucolides 2 and 4 ) were recorded by direct and UPLC water‐sample combined injections. These analyses revealed that hirsutinolides and glaucolides formed [M+Na]+ ion during the operation of the direct MS injection, and ([M+Na]+ and [M+H‐H2O]+) and [M+H]+ ions were respectively observed for hirsutinolides and glaucolides during the operation of combined UPLC water and sample MS injection. Computational simulations showed that the complex hirsutinolide ( 1 )‐Na+ formed with a lower preparation energy compared with the complex glaucolide ( 2 )‐Na+. However, despite their different behavior during the ionization process, ESI(+)‐HRMS/MS analyses of 1 ‐ 4 gave similar fragmentation patterns at m/z 277, 259, 241, and 231 that can be used as diagnostic ions for both skeletons. Moreover, the differentiation strategy based on the nature of the complex SL‐adducts and their MS/MS fragmentation pattern were successfully applied for the chemical characterization of the extract from Vernonanthura tweedieana using UPLC‐ESI‐HRMS/MS. Among the characterized metabolites, SL with hirsutinolide and glaucolide skeletons showed the aforementioned diagnostic fragments and an ionization behavior that was similar to those observed during the water‐sample combined injection.  相似文献   

7.
A high‐performance liquid chromatography/electrospray ionization multi‐stage tandem mass spectrometry (HPLC/ESI‐MSn) method was developed to analyze two structurally related groups of Amaryllidaceae alkaloids (AmAs), crinane‐ and tazettine‐type alkaloids, in the species Crinum latifolium and C. asiaticum, as well as different organs of C. latifolium. In ESI‐MSn spectra of the two types of alkaloids, characteristic fragmentation reactions were observed that allowed us to determine and differentiate them. Based on the fragmentation rules of reference standards, crinane‐type alkaloids displayed concurrent neutral loss of C2H5N (43 u) and C2H6N (44 u) as well as characteristic ions of m/z 213 and 211, whereas tazettine‐type alkaloids exhibited neutral loss of C3H7N (57 u) [or C2H5N (43 u), C3H7NO (73 u)] from the [M+H]+ and [M+H–H2O]+ ions. These were supported by quadrupole time‐of‐flight (Q‐Tof)‐MS/MS analysis. The chemical complexity of the mixture was resolved by profiling. The compositions of the main crinane‐ and tazettine‐type alkaloids in the above‐mentioned species and organs were also compared. Overall, 28 AmAs comprising 14 crinane‐type and 14 tazettine‐type alkaloids were identified and studied by MS. Among them, 14 AmAs were tentatively characterized from the two species for the first time. This method allowed a rapid analysis of alkaloid distribution and composition of Crinum species, and may also be used for quality control and screening of extracts designated for pharmaceutical application. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
2,3‐Dimethyl‐2,3‐dinitrobutane (DMNB) is an explosive taggant added to plastic explosives during manufacture making them more susceptible to vapour‐phase detection systems. In this study, the formation and detection of gas‐phase [M+H]+, [M+Li]+, [M+NH4]+ and [M+Na]+ adducts of DMNB was achieved using electrospray ionisation on a triple quadrupole mass spectrometer. The [M+H]+ ion abundance was found to have a strong dependence on ion source temperature, decreasing markedly at source temperatures above 50°C. In contrast, the [M+Na]+ ion demonstrated increasing ion abundance at source temperatures up to 105°C. The relative susceptibility of DMNB adduct ions toward dissociation was investigated by collision‐induced dissociation. Probable structures of product ions and mechanisms for unimolecular dissociation have been inferred based on fragmentation patterns from tandem mass (MS/MS) spectra of source‐formed ions of normal and isotopically labelled DMNB, and quantum chemical calculations. Both thermal and collisional activation studies suggest that the [M+Na]+ adduct ions are significantly more stable toward dissociation than their protonated analogues and, as a consequence, the former provide attractive targets for detection by contemporary rapid screening methods such as desorption electrospray ionisation mass spectrometry. Copyright © 2009 Commonwealth of Australia. Published by John Wiley & Sons, Ltd.  相似文献   

9.
The fragmentation pathways of protonated imine resveratrol analogues in the gas‐phase were investigated by electrospray ionization–tandem mass spectrometry. Benzyl cations were formed in the imine resveratrol analogues that had an ortho‐hydroxyl group on the benzene ring A. The specific elimination of the quinomethane neutral, CH2 = C6H4 = O, from the two isomeric ions [M1 + H]+ and [M3 + H]+ via the corresponding ion–neutral complexes was observed. The fragmentation pathway for the related meta‐isomer, ion [M2 + H]+ and the other congeners was not observed. Accurate mass measurements and additional experiments carried out with a chlorinated analogue and the trideuterated isotopolog of M1 supported the overall interpretation of the fragmentation phenomena observed. It is very helpful for understanding the intriguing roles of ortho‐hydroxyl effect and ion–neutral complexes in fragmentation reactions and enriching the knowledge of the gas‐phase chemistry of the benzyl cation. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

10.
Matrix‐assisted laser desorption/ionization in‐source decay (MALDI‐ISD) induces N–Cα bond cleavage via hydrogen transfer from the matrix to the peptide backbone, which produces a c′/z? fragment pair. Subsequently, the z? generates z′ and [z + matrix] fragments via further radical reactions because of the low stability of the z?. In the present study, we investigated MALDI‐ISD of a cyclic peptide. The N–Cα bond cleavage in the cyclic peptide by MALDI‐ISD produced the hydrogen‐abundant peptide radical [M + 2H]+? with a radical site on the α‐carbon atom, which then reacted with the matrix to give [M + 3H]+ and [M + H + matrix]+. For 1,5‐diaminonaphthalene (1,5‐DAN) adducts with z fragments, post‐source decay of [M + H + 1,5‐DAN]+ generated from the cyclic peptide showed predominant loss of an amino acid with 1,5‐DAN. Additionally, MALDI‐ISD with Fourier transform‐ion cyclotron resonance mass spectrometry allowed for the detection of both [M + 3H]+ and [M + H]+ with two 13C atoms. These results strongly suggested that [M + 3H]+ and [M + H + 1,5‐DAN]+ were formed by N–Cα bond cleavage with further radical reactions. As a consequence, the cleavage efficiency of the N–Cα bond during MALDI‐ISD could be estimated by the ratio of the intensity of [M + H]+ and [M + 3H]+ in the Fourier transform‐ion cyclotron resonance spectrum. Because the reduction efficiency of a matrix for the cyclic peptide cyclo(Arg‐Gly‐Asp‐D‐Phe‐Val) was correlated to its tendency to cleave the N–Cα bond in linear peptides, the present method could allow the evaluation of the efficiency of N–Cα bond cleavage for MALDI matrix development. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

11.
In the search for novel natural products in plants, particularly those with potential bioactivity, it is important to efficiently distinguish novel compounds from previously isolated, known compounds, a process known as dereplication. In this study, electrospray ionization‐multiple stage tandem mass spectrometry (ESI‐MSn) was used to study the behaviour of 12 pregnane glycosides and genins previously isolated from Marsdenia tenacissima, a traditional Chinese medicinal plant, as a basis for dereplication of compounds in a plant extract. In addition to [M + Na]+ and [M + NH4]+ ions, a characteristic [M‐glycosyl + H]+ ion was observed in full‐scan mode with in‐source fragmentation. Sequential in‐trap collision‐induced dissociation of [M + Na]+ ions from 11,12‐diesters revealed consistent preferred losses of substituents first from C‐12, then from C‐11, followed by losses of monosaccharide fragments from the C‐3 tri‐ and tetrasaccharide substituents. A crude methanol extract of M. tenacissima stems was analysed using high‐performance liquid chromatography coupled to ESI‐MS. Several previously isolated pregnane glycosides were dereplicated, and the presence of an additional nine novel pregnane glycosides is predicted on the basis of the primary and fragment ions observed, including two with a previously unreported C4H7O C‐11/C‐12 substituent of pregnane glycosides. This study is the first report of prediction of the structures of novel pregnane glycosides in a crude plant extract by a combination of in‐source fragmentation and in‐trap collision‐induced dissociation and supports the usefulness of LC‐ESI‐MSn not only for dereplication of active compounds in extracts of medicinal plants but also for detecting the presence of novel related compounds. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

12.
Low molecular weight polyisobutylenes (PIB) with chlorine, olefin and succinic acid end‐groups were studied using direct analysis in real time mass spectrometry (DART‐MS). To facilitate the adduct ion formation under DART conditions, NH4Cl as an auxiliary reagent was deposited onto the PIB surface. It was found that chlorinated adduct ions of olefin and chlorine telechelic PIBs, i.e. [M + Cl]? up to m/z 1100, and the deprotonated polyisobutylene succinic acid [M? H]? were formed as observed in the negative ion mode. In the positive ion mode formation of [M + NH4]+, adduct ions were detected. In the tandem mass (MS/MS) spectra of [M + Cl]?, product ions were absent, suggesting a simple dissociation of the precursor [M + Cl]? into a Cl? ion and a neutral M without fragmentation of the PIB backbones. However, structurally important product ions were produced from the corresponding [M + NH4]+ ions, allowing us to obtain valuable information on the arm‐length distributions of the PIBs containing aromatic initiator moiety. In addition, a model was developed to interpret the oligomer distributions and the number average molecular weights observed in DART‐MS for PIBs and other polymers of low molecular weight. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

13.
Fragmentation reactions of β‐hydroxymethyl‐, β‐acetoxymethyl‐ and β‐benzyloxymethyl‐butenolides and the corresponding γ‐butyrolactones were investigated by electrospray ionization tandem mass spectrometry (ESI‐MS/MS) using collision‐induced dissociation (CID). This study revealed that loss of H2O [M + H ?18]+ is the main fragmentation process for β‐hydroxymethylbutenolide (1) and β‐hydroxymethyl‐γ‐butyrolactone (2). Loss of ketene ([M + H ?42]+) is the major fragmentation process for protonated β‐acetoxymethyl‐γ‐butyrolactone (4), but not for β‐acetoxymethylbutenolide (3). The benzyl cation (m/z 91) is the major ion in the ESI‐MS/MS spectra of β‐benzyloxymethylbutenolide (5) and β‐benzyloxymethyl‐γ‐butyrolactone (6). The different side chain at the β‐position and the double bond presence afforded some product ions that can be important for the structural identification of each compound. The energetic aspects involved in the protonation and gas‐phase fragmentation processes were interpreted on the basis of thermochemical data obtained by computational quantum chemistry. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

14.
Glycopeptidolipids (GPLs) are abundant in the cell walls of different species of mycobacteria and consist of tripeptide‐amino‐alcohol core of D‐Phe‐D‐allo‐Thr‐D‐Ala‐L‐alaninol linked to 3‐hydroxy or 3‐methoxy C26–34 fatty acyl chain at the N‐terminal of D‐Phe via amide linkage, and a 6‐deoxytalose (6‐dTal) and an O‐methyl rhamnose residues, respectively, attach to D‐allo‐Thr and the terminal L‐alaninol. They are important cell‐surface antigens that are implicated in the pathogenesis of opportunistic mycobacteria belonging to the Mycobacterium avium complex. In this contribution, we described multiple‐stage linear ion trap in conjunction with high‐resolution mass spectrometry towards structural characterization of complex GPLs as [M + Na]+ ions isolated from Mycobacterium smegmatis, a fast‐growing and non‐pathogenic mycobacterial species. Following resonance excitation in an ion trap, MSn spectra of the [M + Na]+ ions of GPLs contained mainly b and y series ions that readily determine the peptide sequence. Fragment ions from MSn also afford locating the 6‐dTal and O‐methyl rhamnose residues linked to the D‐allo‐Thr and terminal L‐alaninol of the peptide core, respectively, as well as recognizing the modifications of the glycosides, including their acetylation and methylation states and the presence of succinyl group. The GPL families consisting of 3‐hydroxy fatty acyl and of 3‐methoxy fatty acyl substituents are readily distinguishable. The MS profiles of the GPLs from cells are dependant on the conditions they were grown, and several isobaric isomers were identified for many of the molecular species. These multiple‐stage mass spectrometric approaches give detailed structures of GPL in complex mixtures of which the isomeric structures are difficult to define using other analytical methods. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

15.
Long‐chain ferulic acid esters, such as eicosyl ferulate ( 1 ), show a complex and analytically valuable fragmentation behavior under negative ion electrospay collision‐induced dissociation ((?)‐ESI‐CID) mass spectrometry, as studied by use of a high‐resolution (Orbitrap) mass spectrometer. In a strong contrast to the very simple fragmentation of the [M + H]+ ion, which is discussed briefly, the deprotonated molecule, [M – H]?, exhibits a rich secondary fragmentation chemistry. It first loses a methyl radical (MS2) and the ortho‐quinoid [M – H – Me]‐? radical anion thus formed then dissociates by loss of an extended series of neutral radicals, CnH2n + 1? (n = 0–16) from the long alkyl chain, in competition with the expulsion of CO and CO2 (MS3). The further fragmentation (MS4) of the [M – H – Me – C3H7]? ion, discussed as an example, and the highly specific losses of alkyl radicals from the [M – H – Me – CO]‐? and [M – H – Me – CO2]‐? ions provide some mechanistic and structural insights.  相似文献   

16.
A sensitive and reliable HPLC coupled with diode array detection and MS method was developed and validated for the first time to simultaneously identify and quantify eight characteristic 5,6,7,8‐tetrahydro‐2‐(2‐phenylethyl)chromones (THPECs) in Chinese eaglewood. Chromatographic separation was performed on a Zorbax SB C18 column with a gradient of acetonitrile/0.1% formic acid/water as the mobile phase. The MS fragmentation behavior of THPECs was characterized as the successive neutral loss of two molecules of H2O ([M+H–18–18]+) and then two molecules of CO ([M+H–18–18–28–28]+), which could be used to differentiate Chinese eaglewood from counterfeits. Validation of the developed analytical method showed good linearity, satisfactory precision, and good recovery. The established method was successfully applied to the simultaneous determination of eight THPECs in ten batches of Chinese eaglewood, which could be used as a tool for the quality control of Chinese eaglewood.  相似文献   

17.
The adduct ions of two tetramolecular G‐quadruplexes formed from the d(TGGGGT) and d(TTGGGGGT) single strands with a group of cationic porphyrins, with different charges and substituents, and one neutral porphyrin, were investigated by ESI‐MS and ESI‐MS/MS in the negative ion mode. Formation of [Q + nNH4++Pp+‐(z + n + p)H+]z‐ adduct ions (where Q = quadruplex, n = number of quartets minus 1, P = porphyrin and p+ =0,1,2,3,4) indicates that the porphyrins are bound outside the quadruplexes providing an additional stabilization to those structures. The fragmentation pathways of the [Q + nNH4++Pp+‐(z + n + p)H+]z‐ adduct ions depend on the number of positive charges (p+) of the porphyrins and on the overall complex charge (z), but do not show a significant dependence on the type of the substituent groups in the porphyrins. Formation of the ‘unfilled’ ions [Q + Pp+‐(z + p)H+]z‐ predominates for porphyrins with a higher number of positive charges. Strand separation with the formation of [T + Pp+‐(z‐2 + p)H+](z‐2)‐ and (SS‐2H+)2‐ ions, where T = [d(TG4T)]3 and [d(T2G5T)]3 and SS = d(TG4T) and d(T2G5T) is only observed for the complexes with a higher overall negative charge. Porphyrin loss with the formation of [Q + nNH4+‐(z + n)H+]z‐ ions occurs predominantly for the neutral and monocharged porphyrins. The predominant formation of the ‘unfilled’ ions, [Q + Pp+‐(z + n)H+]z‐, for porphyrins with a higher number of charges shows that these porphyrins can prevent strand separation and preserve, at least partially, the quadruplex structure. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

18.
In this work a systematic strategy integrating liquid chromatography/tandem mass spectrometry (LC/MS/MS) and online databases was developed to identify phosphocholines (PC) and lysophosphatidylcholines (LPC) in human red blood cells (RBCs). First of all, the neutral loss scan of 59 and the precursor ion scan of m/z 184 were performed to find out the possible lipids with phosphocholine head‐group structure in RBCs. The acquired [M+H]+ and [M+Na]+ adduct ions were then identified online using the Human Metabolome Database (HMDB) and the LIPID MAPS, which were then further confirmed by their MS/MS fragmentation. Based on the comparison of chemical structures of the detected PC and LPC with their corresponding MS/MS fragmentation pathways, several new diagnostic fragments or fragmentation pathway were found. These include, (1) the neutral losses of 183 could be used as a diagnostic fragmentation to discriminate PC and LPC; (2) product ions at m/z 104 could be used to distinguish LPC and their sn‐2 isomers; (3) fragment ions at m/z 184 are characteristic fragmentation that could be used for discrimination of sodiated ions [M+Na]+ and protonated ions [M+H]+, especially for co‐eluted PC or LPC with a molecular weight difference of 22. The structures of the above‐mentioned fragment ions were confirmed by quadrupole time‐of‐flight (Q‐TOF) MS. Furthermore, a PC and LPC focused LC/MS semi‐quantification approach was also developed and validated. This approach could be useful for future lipidomic study. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

19.
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   

20.
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   

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