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.
In studying the metabolic pathways underlying the mechanism of carcinogenesis of the heterocyclic amine of 2‐amino‐3‐methylimidazo[4,5‐f]quinoline (IQ), we recently found a new metabolite which gave an [M + H]+ ion of m/z 217 when subjected to electrospray ionization (ESI) in positive‐ion mode. Following ip injection of this metabolite of m/z 217 (designated as m/z 217) to beta‐naphthoflavone‐treated mice, 57% of the total radioactivity was recovered in a 24‐h mouse urine sample. HPLC separation followed by MS analysis indicates that the urine sample contained m/z 217 (36 ± 3% of total recovered radioactivity) and two other peaks that gave rise to the [M + H]+ ions of m/z 393 (31 ± 4%, designated as m/z 393) and m/z 233 (14 ± 1%, designated as m/z 233). Beta‐glucuronidase treatment of m/z 393 resulted in a radioactive peak corresponding to m/z 217. ESI in combination with various mass spectrometry techniques, including multiple‐stage mass spectrometry, exact mass measurements and H/D exchange followed by tandem mass spectrometry, was used for structural characterization. The urinary metabolites of m/z 217, 393 and 233 were identified as 1,2‐dihydro‐2‐amino‐5‐hydroxy‐3‐methylimidazo[4,5‐f]quinoline, 1,2‐dihydro‐2‐amino‐5‐O‐glucuronide‐3‐methylimidazo[4,5‐f]quinoline and 1,2‐dihydro‐2‐amino‐5,7‐dihydroxy‐3‐methylimidazo[4,5‐f]quinoline, respectively. Our results demonstrated that m/z 217 is biotransformed in vivo to m/z 393 by O‐glucuronidation and to m/z 233 by oxidation. The observation of these more polar metabolites relative to IQ suggests that they may arise from a previously undescribed detoxicification pathway. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

2.
In this study, clostebol metabolic profiles were investigated carefully. Clostebol was administered to one healthy male volunteer. Urinary extracts were analyzed by liquid chromatography quadrupole time‐of‐flight mass spectrometry (MS) using full scan and targeted MS/MS techniques with accurate mass measurement for the first time. Liquid–liquid extraction and direct injection were applied to processing urine samples. Chromatographic peaks for potential metabolites were found by using the theoretical [M–H]? as target ion in full scan experiment, and their actual deprotonated ions were analyzed in targeted MS/MS mode. Fourteen metabolites were found for clostebol, and nine unreported metabolites (two free ones and seven sulfate conjugates) were identified by MS, and their potential structures were proposed based on fragmentation and metabolism pathways. Four glucuronide conjugates were also first reported. All the metabolites were evaluated in terms of how long they could be detected and S1 (4ξ‐chloro‐5ξ‐androst‐3ξ‐ol‐17‐one‐3ξ‐sulfate) was considered to be the long‐term metabolite for clostebol misuse detected up to 25 days by liquid–liquid extraction and 14 days by direct injection analysis after oral administration. Five conjugated metabolites (M2, M5, S2, S6 and S7) could also be the alternative biomarkers for clostebol misuse. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

3.
This paper describes the application of gas chromatography–mass spectrometry (GC‐MS) for in vitro and in vivo studies of 6‐OXO in horses, with a special aim to identify the most appropriate target metabolite to be monitored for controlling the administration of 6‐OXO in racehorses. In vitro studies of 6‐OXO were performed using horse liver microsomes. The major biotransformation observed was reduction of one keto group at the C3 or C6 positions. Three in vitro metabolites, namely 6α‐hydroxyandrost‐4‐ene‐3,17‐dione (M1), 3α‐hydroxyandrost‐4‐ene‐6,17‐dione (M2a) and 3β‐hydroxyandrost‐4‐ene‐6,17‐dione (M2b) were identified. For the in vivo studies, two thoroughbred geldings were each administered orally with 500 mg of androst‐4‐ene‐3,6,17‐trione (5 capsules of 6‐OXO®) by stomach tubing. The results revealed that 6‐OXO was extensively metabolized. The three in vitro metabolites (M1, M2a and M2b) identified earlier were all detected in post‐administration urine samples. In addition, seven other urinary metabolites, derived from a further reduction of either one of the remaining keto groups or one of the remaining keto groups and the olefin group, were identified. These metabolites included 6α,17β‐dihydroxyandrost‐4‐en‐3‐one (M3a), 6,17‐dihydroxyandrost‐4‐en‐3‐one (M3b and M3c), 3β,6β‐dihydroxyandrost‐4‐en‐17‐one (M4a), 3,6‐dihydroxyandrost‐4‐en‐17‐one (M4b), 3,6‐dihydroxyandrostan‐17‐one (M5) and 3,17‐dihydroxyandrostan‐6‐one (M6). The longest detection time observed in urine was up to 46 h for the M6 metabolite. For blood samples, the peak 6‐OXO plasma concentration was observed 1 h post administration. Plasma 6‐OXO decreased rapidly and was not detectable 12 h post administration. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

4.
In this study, tamoxifen metabolic profiles were investigated carefully. Tamoxifen was administered to two healthy male volunteers and one female patient suffering from breast cancer. Urinary extracts were analyzed by liquid chromatography quadruple time‐of‐flight mass spectrometry using full scan and targeted MS/MS techniques with accurate mass measurement. Chromatographic peaks for potential metabolites were selected by using the theoretical [M + H]+ as precursor ion in full‐scan experiment and m/z 72, 58 or 44 as characteristic product ions for N,N‐dimethyl, N‐desmethyl and N,N‐didesmethyl metabolites in targeted MS/MS experiment, respectively. Tamoxifen and 37 metabolites were detected in extraction study samples. Chemical structures of seven unreported metabolites were elucidated particularly on the basis of fragmentation patterns observed for these metabolites. Several metabolic pathways containing mono‐ and di‐hydroxylation, methoxylation, N‐desmethylation, N,N‐didesmethylation, oxidation and combinations were suggested. All the metabolites were detected in the urine samples up to 1 week. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

5.
The behavior of water entrapped in reverse micelles (RMs) formed by two catanionic ionic liquid‐like surfactants, benzyl‐n‐hexadecyldimethylammonium 1,4‐bis‐2‐ethylhexylsulfosuccinate (AOT‐BHD) and cetyltrimethylammonium 1,4‐bis‐2‐ethylhexylsulfosuccinate (AOT‐CTA), was investigated by using dynamic (DLS) and static (SLS) light scattering, FTIR, and 1H NMR spectroscopy techniques. To the best of our knowledge, this is the first report in which AOT‐CTA has been used to create RMs and encapsulate water. DLS and SLS results revealed the formation of RMs in benzene and the interaction of water with the RM interface. From FTIR and 1H NMR spectroscopy data, a difference in the magnitude of the water–catanionic surfactant interaction at the interface is observed. For the AOT‐BHD RMs, a strong water–surfactant interaction can be invoked whereas for AOT‐CTA this interaction seems to be weaker. Consequently, more water molecules interact with the interface in AOT‐BHD RMs with a completely disrupted hydrogen‐bond network, than in AOT‐CTA RMs in which the water structure is partially preserved. We suggest that the benzyl group present in the BHD+ moiety in AOT‐BHD is responsible for the behavior of the catanionic interface in comparison with the interface created in AOT‐CTA. These results show that a simple change in the cationic component in the catanionic surfactant promotes remarkable changes in the RMs interface with interesting consequences, in particular when using the systems as nanoreactors.  相似文献   

6.
Triptolide (TP), the primary active component of the herbal medicine Tripterygium wilfordii Hook F, has shown promising antileukemic and anti‐inflammatory activity. The pharmacokinetic profile of TP indicates an extensive metabolic elimination in vivo; however, its metabolic data is rarely available partly because of the difficulty in identifying it due to the absence of appropriate ultraviolet chromophores in the structure and the presence of endogenous interferences in biological samples. In the present study, the biotransformation of TP was investigated by improved data‐dependent accurate mass spectrometric analysis, using an LTQ/Orbitrap hybrid mass spectrometer in conjunction with the online hydrogen (H)/deuterium (D) exchange technique for rapid structural characterization. Accurate full‐scan MS and MS/MS data were processed with multiple post‐acquisition data‐mining techniques, which were complementary and effective in detecting both common and uncommon metabolites from biological matrices. As a result, 38 phase I, 9 phase II and 8 N‐acetylcysteine (NAC) metabolites of TP were found in rat urine. Accurate MS/MS data were used to support assignments of metabolite structures, and online H/D exchange experiments provided additional evidence for exchangeable hydrogen atoms in the structure. The results showed the main phase I metabolic pathways of TP are hydroxylation, hydrolysis and desaturation, and the resulting metabolites subsequently undergo phase II processes. The presence of NAC conjugates indicated the capability of TP to form reactive intermediate species. This study also demonstrated the effectiveness of LC/HR‐MSn in combination with multiple post‐acquisition data‐mining methods and the online H/D exchange technique for the rapid identification of drug metabolites. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

7.
We have designed and synthesised a [Ru(CO)3Cl2(NAC)] pro‐drug that features an N‐acetyl cysteine (NAC) ligand. This NAC carbon monoxide releasing molecule (CORM) conjugate is able to simultaneously release biologically active CO and to ablate the concurrent formation of reactive oxygen species (ROS). Complexes of the general formulae [Ru(CO)3(L)3]2+, including [Ru(CO)3Cl(glycinate)] (CORM‐3), have been shown to produce ROS through a water–gas shift reaction, which contributes significantly, for example, to their antibacterial activity. In contrast, NAC‐CORM conjugates do not produce ROS or possess antibacterial activity. In addition, we demonstrate the synergistic effect of CO and NAC both for the inhibition of nitric oxide (formation) and in the expression of tumour‐necrosis factor (TNF)‐α. This work highlights the advantages of combining a CO‐releasing scaffold with the anti‐oxidant and anti‐inflammatory drug NAC in a unique pro‐drug.  相似文献   

8.
Flunitrazepam (FNZ) is a potent hypnotic, sedative, and amnestic drug used to treat severe insomnia. In our recent study, FNZ metabolic profiles were investigated carefully. Six authentic human urine samples were purified using solid phase extraction (SPE) without enzymatic hydrolysis, and urine extracts were then analyzed by liquid chromatography‐Q exactive‐HF hybrid quadrupole‐Orbitrap‐mass spectrometry (LC‐QE‐HF‐MS), using the full scan positive ion mode and targeted MS/MS (ddms2) technique to make accurate mass measurements. There were 25 metabolites, including 13 phase I and 12 phase II metabolites, which were detected and tentatively identified by LC‐QE‐HF‐MS. In addition, nine previously unreported phase II glucuronide conjugates and four phase I metabolites are reported here for the first time. Eight metabolic pathways, including N‐reduction and O‐reduction, N‐glucuronidation, O‐glucuronidation, mono‐hydroxylation and di‐hydroxylation, demethylation, acetylation, and combinations, were implicated in this work, and 2‐O‐reduction together with dihydroxylation were two novel metabolic pathways for FNZ that were identified tentatively. Although 7‐amino FNZ is widely considered to be the primary metabolite, a previously unreported metabolites (M12) can also serve as a potential biomarker for FNZ misuse.  相似文献   

9.
Trends in lanthanide(III) (LnIII) coordination were investigated within nanoconfined solvation environments. LnIII ions were incorporated into the cores of reverse micelles (RMs) formed with malonamide amphiphiles in n‐heptane by contact with aqueous phases containing nitrate and LnIII; both insert into pre‐organized RM units built up of DMDOHEMA (N,N′‐dimethyl‐N,N′‐dioctylhexylethoxymalonamide) that are either relatively large and hydrated or small and dry, depending on whether the organic phase is acidic or neutral, respectively. Structural aspects of the LnIII complex formation and the RM morphology were obtained by use of XAS (X‐ray absorption spectroscopy) and SAXS (small‐angle X‐ray scattering). The LnIII coordination environments were determined through use of L3‐edge XANES (X‐ray absorption near edge structure) and EXAFS (extended X‐ray absorption fine structure), which provide metrical insights into the chemistry across the period. Hydration numbers for the Eu species were measured using TRLIFS (time‐resolved laser‐induced fluorescence spectroscopy). The picture that emerges from a system‐wide perspective of the Ln? O interatomic distances and number of coordinating oxygen atoms for the extracted complexes of LnIII in the first half of the series (i.e., Nd, Eu) is that they are different from those in the second half of the series (i.e., Tb, Yb): the number of coordinating oxygen atoms decrease from 9 O for early lanthanides to 8 O for the late ones—a trend that is consistent with the effect of the lanthanide contraction. The environment within the RM, altered by either the presence or absence of acid, also had a pronounced influence on the nitrate coordination mode; for example, the larger, more hydrated, acidic RM core favors monodentate coordination, whereas the small, dry, neutral core favors bidentate coordination to LnIII. These findings show that the coordination chemistry of lanthanides within nanoconfined environments is neither equivalent to the solid nor bulk solution behaviors. Herein we address atomic‐ and mesoscale phenomena in the under‐explored field of lanthanide coordination and periodic behavior within RMs, providing a consilience of fundamental insights into the chemistry of growing importance in technologies as diverse as nanosynthesis and separations science.  相似文献   

10.
Redox mediators (RMs) are considered an effective countermeasure to reduce the large polarization in lithium‐oxygen batteries. Nevertheless, achieving sufficient enhancement of the cyclability is limited by the trade‐offs of freely mobile RMs, which are beneficial for charge transport but also trigger the shuttling phenomenon. Here, we successfully decoupled the charge‐carrying redox property of RMs and shuttling phenomenon by anchoring the RMs in polymer form, where physical RM migration was replaced by charge transfer along polymer chains. Using PTMA (poly(2,2,6,6‐tetramethyl‐1‐piperidinyloxy‐4‐yl methacrylate)) as a polymer model system based on the well‐known RM tetramethylpiperidinyloxyl (TEMPO), it is demonstrated that PTMA can function as stationary RM, preserving the redox activity of TEMPO. The efficiency of RM‐mediated Li2O2 decomposition remains remarkably stable without the consumption of oxidized RMs or degradation of the lithium anode, resulting in an improved performance of the lithium‐oxygen cell.  相似文献   

11.
A modified electrode was fabricated by grafting of poly (2,6‐pyridinedicarboxylic acid) film (PDC) by electropolymerization of 2,6‐pyridinedicarboxylic acid on the glassy carbon electrode (GCE). Then, gold nanoparticles (NG) and 1,2‐naphthoquinone‐4‐sulfonic acid sodium (Nq) were immobilized on the PDC/GCE to prepare Nq/NG/PDC/GCE by immersing electrode into NG and Nq solution, respectively. The Nq species on NG/PDC/GCE could catalyze electrooxidation of N‐acetyl‐L ‐cysteine (NAC) with lowering the over potential by about 600 mV. This method used for detection of NAC in dynamic range from 4.0×10?6 M to 1.30×10?4 M with a detection of limit (2σ) 8.0×10?7 M.  相似文献   

12.
Corynoxeine (COR) is one of 4 bioactive oxindole alkaloids in Uncaria species. In this work two phase I metabolites, namely 11‐hydroxycorynoxeine (M1) and 10‐hydroxycorynoxeine (M2), and two phase II metabolites, namely 11‐hydroxycorynoxeine 11‐O‐β‐d ‐glucuronide (M3) and 10‐hydroxycorynoxeine 10‐O‐β‐d ‐glucuronide (M4), were detected in rat bile after oral dose of COR (0.105 mmol/kg), by optimized high‐performance liquid chromatography–tandem mass spectrometry (LC‐MSn) with electrospray ionization in positive ion mode. Structures of M1–4 were determined by LC‐MSn, nuclear magnetic resonance, circular dichroism and high‐resolution MS spectra. COR and its metabolites in rat bile were quantified by LC‐MSn. The LC‐MSn quantification methods for COR and its metabolites yielded a linearity with coefficient of determination ≥0.995 from 5.0 × 10?10 to 5.0 × 10?7 m . The recoveries of stability tests varied from 96.80 to 103.10%. Accuracy ranged from 91.00 to 105.20%. Relative standard deviation for intra‐day and inter‐day assay was <5.0%. After the oral dose 0.14% of COR was detected in rat bile from 0 to 8 h, in which in total 97.8% COR biotransformed into M1–4. M1 and M2 yielded 48.1 and 49.7%, which successively glucuronidated to M3 and M4 at 47.2 and 43.8%, respectively. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

13.
A series of Zn (II), Pd (II) and Cd (II) complexes, [(L) n MX 2 ] m (L = L‐a–L‐c; M = Zn, Pd; X = Cl; M = Cd; X = Br; n, m = 1 or 2), containing 4‐methoxy‐N‐(pyridin‐2‐ylmethylene) aniline ( L‐a ), 4‐methoxy‐N‐(pyridin‐2‐ylmethyl) aniline ( L‐b ) and 4‐methoxy‐N‐methyl‐N‐(pyridin‐2‐ylmethyl) aniline ( L‐c ) have been synthesized and characterized. The X‐ray crystal structures of Pd (II) complexes [L 1 PdCl 2 ] (L = L‐b and L‐c) revealed distorted square planar geometries obtained via coordinative interaction of the nitrogen atoms of pyridine and amine moieties and two chloro ligands. The geometry around Zn (II) center in [(L‐a)ZnCl 2 ] and [(L‐c)ZnCl 2 ] can be best described as distorted tetrahedral, whereas [(L‐b) 2 ZnCl 2 ] and [(L‐b) 2 CdBr 2 ] achieved 6‐coordinated octahedral geometries around Zn and Cd centers through 2‐equivalent ligands, respectively. In addition, a dimeric [(L‐c)Cd(μ ‐ Br)Br] 2 complex exhibited typical 5‐coordinated trigonal bipyramidal geometry around Cd center. The polymerization of methyl methacrylate in the presence of modified methylaluminoxane was evaluated by all the synthesized complexes at 60°C. Among these complexes, [(L‐b)PdCl 2 ] showed the highest catalytic activity [3.80 × 104 g poly (methyl methacrylate) (PMMA)/mol Pd hr?1], yielding high molecular weight (9.12 × 105 g mol?1) PMMA. Syndio‐enriched PMMA (characterized using 1H‐NMR spectroscopy) of about 0.68 was obtained with Tg in the range 120–128°C. Unlike imine and amine moieties, the introduction of N‐methyl moiety has an adverse effect on the catalytic activity, but the syndiotacticity remained unaffected.  相似文献   

14.
A simple and sensitive liquid chromatography tandem mass spectrometry method was validated for simultaneous quantification of evodiamine and its metabolites 10‐hydroxyevodiamine (M1), 18‐hydroxyevodiamine (M2), 10‐hydroxyevodiamine‐glucuronide (M3) and 18‐hydroxy‐ evodiamine‐glucuronide (M4) in rat plasma for the first time. The analytes were extracted with acetonitrile and separated on a C18 column within 3 min. The detection was achieved in positive selected reaction monitoring mode with precursor‐to‐product transitions at m/z 304.1 → 161.1 for evodiamine, m/z 320.1 → 134.1 for M1, m/z 320.1 → 150.1 for M2, m/z 496.2 → 134.1 for M3, m/z 496.2 → 171.1 for M4 and m/z 349.2 → 305.1 for camptothecin (internal standard). The linearity was evident over the tested concentration ranges with correlation coefficients >0.9991. The lower limits of quantification for evodiamine, M1, M2, M3 and M4 were 0.1, 0.1, 0.1, 0.25 and 0.25 ng mL−1, respectively. Extraction recoveries and matrix effects of the analytes were within the ranges of 84.51–97.21 and 90.13–103.30%, respectively. The accuracy (relative error) ranged from −8.14 to 7.23% while the intra‐ and inter‐day precisions (relative standard deviation) were < 9.31%. The validated assay was successfully applied for the pharmacokinetic study of evodiamine, M1, M2, M3 and M4 in rat. The current study will be helpful in understanding the in vivo disposition of evodiamine.  相似文献   

15.
Naphthoquine (NQ) is one of important partner drugs of artemisinin‐based combination therapy (ACT), which is recommended for the treatment of uncomplicated Plasmodium falciparum. NQ shows a high cure rate after a single oral administration. It is absorbed quickly (time to peak concentration 2–4 h) and has a long elimination half‐life (255 h). However, the metabolism of NQ has not been clarified. In this work, the metabolite profiling of NQ was studied in six liver microsomal incubates (human, cynomolgus monkey, beagle dog, mini pig, rat and CD1 mouse), seven recombinant CYP enzymes (1A2, 2B6, 2C8, 2C9, 2C19, 2D6 and 3A4) and rat (plasma, urine, bile and feces) using liquid chromatography tandem high‐resolution LTQ‐Orbitrap mass spectrometry (HRMSn) in conjunction with online hydrogen/deuterium exchange. The biological samples were pretreated by protein precipitation and solid‐phase extraction. For data processing, multiple data‐mining tools were applied in tandem, i.e. background subtraction and followed by mass defect filter. NQ metabolites were characterized by accurate MS/MS fragmentation characteristics, the hydrogen/deuterium exchange data and cLogP simulation. As a result, five phase I metabolites (M1–M5) of NQ were characterized for the first time. Two metabolic pathways were involved: hydroxylation and N‐oxidation. This study demonstrates that LC‐HRMSn in combination with multiple data‐mining tools in tandem can be a valuable analytical strategy for rapid metabolite profiling of drugs.  相似文献   

16.
Neratinib is a tyrosine kinase inhibitor that has been approved by the US Food and Drug Administration for the treatment of breast cancer. However, its metabolism remains unknown. This study was carried out to investigate the in vitro and in vivo metabolism of neratinib using an UHPLC‐DAD‐Q Exactive Orbitrap‐MS instrument with dd‐MS2 on‐line data acquisition mode. The post‐acquisition data was processed using MetWorks software. Under the current conditions, a total of 12 metabolites were detected and structurally identified based on their accurate masses, fragment ions and chromatographic retention times. Among these metabolites, M3, M10 and M12 were unambiguously identified using chemically synthesized reference standards. M6 and M7 (GSH conjugates) were the major metabolites. The metabolic pathways of neratinib were proposed accordingly. Our findings suggested that neratinib was mainly metabolized via O‐dealkylation (M3), oxygenation (M8), N‐demethylation (M10), N‐oxygenation (M12), GSH conjugation (M1, M2, M4, M5, M6 and M7) and N‐acetylcysteine conjugation (M9 and M11). The α,β‐unsaturated ketone was the major metabolic site and GSH conjugation was the predominant metabolic pathway. In conclusion, this study provided valuable metabolic data and would benefit the assessment of the contributions to the overall activity or toxicity from the key metabolites.  相似文献   

17.
Ilaprazole is a new proton pump inhibitor designed for the treatment of gastric ulcers, and limited data is available on the metabolism of the drug. In this article, the structural elucidation of urinary metabolites of ilaprazole in human was described by HPLC‐ESI‐MS/MS and stopped‐flow HPLC‐NMR experiments. Urinary samples were precipitated by sodium carbonate solution, and then extracted by liquid–liquid extraction after adding ammonium acetate buffer solution. The enriched sample was separated using a C18 reversed‐phase column with the mobile phase composed of acetonitrile and 0.05 mol/L ammonium acetate buffer solution in a gradient solution, and then directly coupled to ESI‐MS/MS detection in an on‐line mode or 1H‐NMR (500 MHz) spectroscopic detection in a stopped‐flow mode. As a result, four sulfide metabolites, ilaprazole sulfide (M1), 12‐hydroxy‐ilaprazole sulfide (M2), 11,12‐dihydroxy‐ilaprazole sulfide (M3) and ilaprazole sulfide A (M4), were identified by comparing their MS/MS and NMR data with those of the parent drug and available standard compounds. The main biotransformation reactions of ilaprazole were reduction and the aromatic hydroxylation of the parent drug and its relative metabolites. The result testified that HPLC‐ESI‐MS/MS and HPLC‐NMR could be widely applied in detection and identification of novel metabolites. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

18.
Reactive metabolites are thought to play a pivotal role in the pathogenesis of some drug‐induced liver injury (DILI) and idiosyncratic adverse drug reactions (IADRs), which is of concern to patient safety and has been a cause of drugs being withdrawn from the market place. To identify drugs with a lower propensity for causing DILI and/or IADRs, high‐throughput assays to capture reactive metabolites are required in pharmaceutical industry for early drug discovery risk assessment. We describe the development of an assay to detect glutathione adducts with combined high sensitivity, enhanced specificity, and rapid data analysis. In this assay, compounds were incubated with human liver microsomes and a mixture of 1:1 of GSH (γ‐GluCysGly): GSX(γ‐GluCysGly‐13C215N) in a 96‐well plate format. UPLC‐UV and LTQ Orbitrap XL were employed to detect GSH‐adducts using the following mass spectrometry setups: (a) selected ion monitoring (SIM) at m/z of 274 ± 3 Da in negative mode with in‐source fragmentation (SCID), which enables simultaneously monitoring two characteristic product ions of m/z 272.0888 (γ‐glutamyl‐dehydroalanyl‐glycine) and 275.0926 (γ‐glutamyl‐dehydroalanyl‐glycine‐13C215N); (b) full scan mode for acquisition of exact mass of glutathione adducts; (c) data‐dependent MS2 scan through isotopic matching (M:M + 3.00375 = 1:1) for monitoring neutral loss fragments (144 Da from dehydroalanyl‐glycine) and for structural information of glutathione adducts. This approach was qualified using eight compounds known to form GSH conjugates as reported in the literature. The high sensitivity and specificity were demonstrated in identifying unique CysGly adducts in the case of clozapine, diclofenac, and raloxifene and in identifying GSH‐adducts of fragmented parent molecules in the case of amodiaquine and troglitazone. In addition, LC‐UV chromatograms in the presence or absence of GSH/GSX allowed for identification of the rearranged glutathione adducts without aforementioned characteristic fragment ions. Implement of this assay in drug discovery small molecule programs has successfully guided drug design.  相似文献   

19.
In Cannabis sativa, Δ9‐Tetrahydrocannabinolic acid‐A (Δ9‐THCA‐A) is the non‐psychoactive precursor of Δ9‐tetrahydrocannabinol (Δ9‐THC). In fresh plant material, about 90% of the total Δ9‐THC is available as Δ9‐THCA‐A. When heated (smoked or baked), Δ9‐THCA‐A is only partially converted to Δ9‐THC and therefore, Δ9‐THCA‐A can be detected in serum and urine of cannabis consumers. The aim of the presented study was to identify the metabolites of Δ9‐THCA‐A and to examine particularly whether oral intake of Δ9‐THCA‐A leads to in vivo formation of Δ9‐THC in a rat model. After oral application of pure Δ9‐THCA‐A to rats (15 mg/kg body mass), urine samples were collected and metabolites were isolated and identified by liquid chromatography‐mass spectrometry (LC‐MS), liquid chromatography‐tandem mass spectrometry (LC‐MS/MS) and high resolution LC‐MS using time of flight‐mass spectrometry (TOF‐MS) for accurate mass measurement. For detection of Δ9‐THC and its metabolites, urine extracts were analyzed by gas chromatography‐mass spectrometry (GC‐MS). The identified metabolites show that Δ9‐THCA‐A undergoes a hydroxylation in position 11 to 11‐hydroxy‐Δ9‐tetrahydrocannabinolic acid‐A (11‐OH‐Δ9‐THCA‐A), which is further oxidized via the intermediate aldehyde 11‐oxo‐Δ9‐THCA‐A to 11‐nor‐9‐carboxy‐Δ9‐tetrahydrocannabinolic acid‐A (Δ9‐THCA‐A‐COOH). Glucuronides of the parent compound and both main metabolites were identified in the rat urine as well. Furthermore, Δ9‐THCA‐A undergoes hydroxylation in position 8 to 8‐alpha‐ and 8‐beta‐hydroxy‐Δ9‐tetrahydrocannabinolic acid‐A, respectively, (8α‐Hydroxy‐Δ9‐THCA‐A and 8β‐Hydroxy‐Δ9‐THCA‐A, respectively) followed by dehydration. Both monohydroxylated metabolites were further oxidized to their bishydroxylated forms. Several glucuronidation conjugates of these metabolites were identified. In vivo conversion of Δ9‐THCA‐A to Δ9‐THC was not observed. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

20.
This study reports an energy‐resolved mass spectrometric (ERMS) strategy for the characterization of position isomers derived from the reaction of hydroxyl radicals (OH) with diphenhydramine (DPH) that are usually hard to differentiate by other methods. The isomer analogues formed by OH attack on the side chain of DPH are identified with the help of a specific fragment ion peak (m/z 88) in the collision‐induced dissociation (CID) spectrum of the protonated molecule. In the negative ion mode, the breakdown curves of the deprotonated molecules show an order of stability (supported by density functional theory (DFT) calculations) ortho > meta > para of the positional isomers formed by the hydroxylation of the aromatic ring. The gas phase stability of the deprotonated molecules [M ? H]? towards the benzylic cleavage depends mainly on the formation of intramolecular hydrogen bonds and of the mesomeric effect of the phenol hydroxyl. The [M ? H]? molecules of ortho and meta isomers result a peak at m/z 183 with notably different intensities because of the presence/absence of an intramolecular hydrogen bonding between the OH group and C9 protons. The ERMS approach discussed in this report might be an effective replacement for the conventional methods that requires very costly and time‐consuming separation/purification methods along with the use of multi‐spectroscopic methods. Copyright © 2015 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