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.
Menadione (2-methyl-1,4-naphthoquine), also known as vitamin K3, has been widely used as a model compound in the field of oxidative stress-related research. The metabolism of menadione has been studied, and it is known that menadione undergoes a two-electron reduction by NAD(P)H:Quinone oxidoreductase 1 (NQO1) after which the reduced form of menadione (2-methyl-1,4-naphthalenediol, menadiol) is glucuronidated and excreted in urine. To investigate which human UDP-glucuronosyltransferase (UGT) isoforms participate in the glucuronidation of menadiol reduced by NQO1 from menadione, we first constructed heterologously expressed NQO1 in Sf9 cells and tested the menadiol glucuronidating activity of 16 human recombinant UGT isoforms. Of the 16 UGT isoforms, UGTs 1A6, 1A7, 1A8, 1A9, and 1A10 catalyzed menadiol glucuronidation, and, of these, UGTs 1A6 and 1A10 catalyzed menadiol glucuronidation at much higher rates than the other UGTs. Menadiol was regioselectively glucuronidated in the manner of 4-position > 1-position by UGTs 1A7, 1A8, 1A9, and 1A10. In contrast to these UGTs, only UGT1A6 exhibited 1-menadiol-preferential glucuronidating activity. The results suggest possible detoxification pathways for quinones via NQO1 reduction followed by UGT glucuronidation.  相似文献   

2.
Phylloquinone (2-methyl-3-phytyl-1,4-naphthoquinone; vitamin K1) is vital to plants. It is responsible for the one-electron transfer at the A1 site of photosystem I, a process that involves turnover between the quinone and semi-quinone forms of phylloquinone. Using HPLC coupled with fluorometric detection to analyze Arabidopsis leaf extracts, we detected a third redox form of phylloquinone corresponding to its fully reduced - quinol-naphthoquinone ring (PhQH2). A method was developed to quantify PhQH2 and its corresponding oxidized quinone (PhQ) counterpart in a single HPLC run. PhQH2 was found in leaves of all dicotyledonous and monocotyledonous species tested, but not in fruits or in tubers. Its level correlated with that of PhQ, and represented 5-10% of total leaf phylloquinone. Analysis of purified pea chloroplasts showed that these organelles accounted for the bulk of PhQH2. The respective pool sizes of PhQH2 and PhQ were remarkably stable throughout the development of Arabidopsis green leaves. On the other hand, in Arabidopsis and tomato senescing leaves, PhQH2 was found to increase at the expense of PhQ, and represented 25-35% of the total pool of phylloquinone. Arabidopsis leaves exposed to light contained lower level of PhQH2 than those kept in the dark. These data indicate that PhQH2 does not originate from the photochemical reduction of PhQ, and point to a hitherto unsuspected function of phylloquinone in plants. The putative origin of PhQH2 and its recycling into PhQ are discussed.  相似文献   

3.
Menaquinone-7, which is vitamin K2 (menatetrenone) with seven isoprene units, is highly contained in the fermented soybean. The effect of menaquinone-7 (MK-7) on osteoblastic bone formation was investigated. Femoral-diaphyseal and metaphyseal tissues of young male rats (4 weeks old) were cultured for 48 h in a medium containing either vehicle or MK-7 (10–7–10–5 M). Calcium content, alkaline phosphatase activity, and deoxyribonuclic acid (DNA) content in the diaphyseal and metaphyseal tissues was significantly increased in the presence of MK-7 (10–6 and 10–5 M). The effect of MK-7 in increasing the diaphyseal and metaphyseal calcium content and alkaline phosphatase activity was completely prevented in the presence of cycloheximide (10–6 M), an inhibitor of protein synthesis. Moreover, osteoblastic MC3T3-E1 cells after subculture were cultured for 24 h in a serum-free medium containing MK-7 (10–7–10–5 M). Protein content, alkaline phophatase activity, osteocalcin and DNA content in the cells was significantly increased in the presence of MK-7 (10–6 and 10–5 M). The effect of MK-7 in increasing protein content, alkaline phosphatase activity, and osteocalcin production in the cells was completely blocked by cycloheximide. This study demonstrates that MK-7 has an anabolic effect on bone tissue and osteoblastic MC3T3-E1 cells in vitro, suggesting that the compound can stimulate osteoblastic bone formation.  相似文献   

4.
A differential pulse polarographic assay for plasma vitamin K3 (menadione) has been developed. Details of the assay are (i) lipid-soluble material is extracted from plasma into ether by the method of Bjornsson et al. [(1978) Thromb. Haemostas.2, 466–473]; (ii) ether is evaporated under nitrogen and the residue is dissolved in the supporting electrolyte, methanol: 0.2 m borate buffer (9:1), pH 6.8; (iii) current height is measured at ?0.32 V vs SCE on the differential pulse polarogram. The lower sensitivity limit of this technique after addition of standard vitamin K3 to plasma is 0.3 μm; the calibration curve is linear from 0.6 through 10 μm. Two patients treated with a single dose of menadiol sodium diphosphate, 20 mg/M2 i.m., achieved measurable plasma vitamin K3 levels at 0.5 to 1.0 h ranging between 0.5 (0.08 μg/ml) and 2 μm (0.3 μg/ml).  相似文献   

5.
In vitro studies show that docosahexaenoic acid (DHA) can be released from membrane phospholipid by Ca2+-independent phospholipase A2 (iPLA2), Ca2+-independent plasmalogen PLA2 or secretory PLA2 (sPLA2), but not by Ca2+-dependent cytosolic PLA2 (cPLA2), which selectively releases arachidonic acid (AA). Since glutamatergic NMDA (N-methyl-D-aspartate) receptor activation allows extracellular Ca2+ into cells, we hypothesized that brain DHA signaling would not be altered in rats given NMDA, to the extent that in vivo signaling was mediated by Ca2+-independent mechanisms. Isotonic saline, a subconvulsive dose of NMDA (25 mg/kg), MK-801, or MK-801 followed by NMDA was administered i.p. to unanesthetized rats. Radiolabeled DHA or AA was infused intravenously and their brain incorporation coefficients k*, measures of signaling, were imaged with quantitative autoradiography. NMDA or MK-801 compared with saline did not alter k* for DHA in any of 81 brain regions examined, whereas NMDA produced widespread and significant increments in k* for AA. In conclusion, in vivo brain DHA but not AA signaling via NMDA receptors is independent of extracellular Ca2+ and of cPLA2. DHA signaling may be mediated by iPLA2, plasmalogen PLA2, or other enzymes insensitive to low concentrations of Ca2+. Greater AA than DHA release during glutamate-induced excitotoxicity could cause brain cell damage.  相似文献   

6.
The effect of mutations in the genes encoding dehydrogenases and oxidases on the resistance of the Synechocystis sp. PCC 6803 cyanobacterium to menadione, an oxidative stress inducer, was studied. An enhanced sensitivity to menadione was observed in the mutants carrying inserts in the drgA gene encoding the NAD(P)H:quinone oxidoreductase (NQR) and in the ndhB gene encoding the subunit of NDH-1 complex. The menadione resistance in the mutants lacking oxidases (Ox), succinate dehydrogenase (SDH), and NDH-2 dehydrogenase do not differ from those in wild-type cells. An additional mutation in the drgA gene increased the sensitivity to menadione in the NDH-2 and Ox mutants. The double mutant that lacks both SDH and NQR was not viable. The expression of the drgA gene decreased during cell incubation in the dark but increased in the presence of glucose both in the dark and in light. Under photoautotrophic growth conditions, the dehydrogenase activity of the cells mainly depends on the NQR and NDH-1 functions. The re-reduction rate of the photosystem I reaction center (P700+) increased in wild-type and NDH-1 mutants after its oxidation with white light in the presence of DCMU after addition of menadione, and it decreased in the NQR mutant. The reduction of P700+ was accelerated in the presence of menadiol in all the strains studied. These results suggest that NQR provides defense of cyanobacterium cells from the toxic effect of menadione via its two-electron reduction to menadiol. An increased sensitivity of the NDH-1 mutant to menadione may result from the inhibition of respiration and the cyclic electron transport in photosystem I.  相似文献   

7.
This investigation tested the importance of excitatory amino acids' effects on regional cerebral O2 consumption and the concomitant changes in cerebral blood flow (rCBF) in isoflurane anesthetized rats. In the glutamate or N-methyl-D-aspartate (NMDA) groups, 10–2 M glutamate or NMDA was topically applied to the right cortex and the left cortex was used as a control. One mg/kg dizocilpine maleate (MK-801), a non-competitive NMDA receptor antagonist, was administered (iv) to the MK-801 group and saline was given to the control group. Cortical rCBF was determined using 14C-iodoantipyrine and regional O2 extraction was measured microspectrophotometrically. Cerebral O2 consumption increased 77% after glutamate (contralateral cortex: 9.0 ± 1.1 ml O2/min/100 g, glutamate treated cortex: 15.9 ± 3.9), while a 46% increase was observed with the same concentration of NMDA (contralateral cortex: 9.8 ± 2.0, NMDA treated cortex: 14.3 ± 5.5). After MK-801, the O2 consumption decreased to 37% of the control value (control cortex: 7.0 ± 1.3, MK-801 treated cortex: 2.6 ± 3.9). MK-801 significantly decreased cerebral O2 extraction from 7.1 ± 1.3 ml O2/100 ml (control cortex) to 5.3 ± 0.6 (MK-801 treated cortex). However, there was no significant difference in cerebral O2 extraction between treated and contralateral cortex in either the glutamate or NMDA groups. The increase in O2 consumption caused by glutamate or NMDA was coupled with increased rCBF. Glutamate increased rCBF from 95 ± 5 ml/min/100 g (contralateral cortex) to 165 ± 31 (treated cortex), while NMDA increased rCBF from 114 ± 12 (contralateral cortex) to 178 ± 60 (treated cortex). MK-801 decreased O2 consumption with a lesser decrease of rCBF. The rCBF was 48 ± 9 in the MK-801 treated cortex and 99 ± 22 in the control cortex. Some substances produced by the activation of NMDA receptors may be related to the coupling of cerebral metabolism and blood flow, since after blockade of NMDA receptors with MK-801, this relationship is uncoupled. These findings suggest that glutamatergic processes have a major effect on cerebral O2 consumption and that this is at least partly due to NMDA receptors.  相似文献   

8.
Synthetic vitamin K3 (VK3, 2-methyl-1,4-naphthoquinone, or menadione) has been found to exhibit antitumor activity against various human cancer cells at relative high dose. Parallel to our study on the mechanism of VK3 action and for future clinical trials in Taiwan, we developed a simple, sensitive and accurate high-performance liquid chromatographic method for the determination of VK3 in biological fluids. VK3 was extracted from the plasma samples with n-hexane. The chromatographic separation employed an ODS analytical column (5 μm, 250 × 4.6 mm I.D.) with a mobile phase of methanol-water (70:30 v/v) and UV detection at 265 nm. On completely drying of the extraction solution, n-hexane, by a stream of nitrogen, menadione was lost to a great extent. Methanol (70%, 200 μl) was added to the extraction solvent after extraction and centrifugation to prevent the loss of menadione. The absolute recovery was 82.4±7.69% (n = 7). The within-day and between-day calibration curves of VK3 in plasma in the ranges of interest (0.01–10.00 μg/ml; 0.01–5.00 μg/ml) showed good linearity (r>0.999) and acceptable precision. The limit of quantitation of VK3 was 10 ng/ml) showed good method has been succesfully applied to a pilot pharmacokinetic study of VK3 in rabbits receiving an intravenous high-dose bolus injection of 75 mg menadiol sodium diphosphate (Synkayvite). The pharmacokinetic properties of menadione could be described adequately by an open two-compartment model. The mean half-life of menadiol (transformation to menadione) was 2.60±0.12 min. The elimination half-life, volume of distribution and plasma clearance of menadione were 26.3±2.97 min, 25.7±0.78 1, and 0.68±0.10 1/min, respectively.  相似文献   

9.
The reference standard IUR-1601 ((S)-N-(2-chloro-3-(trifluoromethyl)benzyl)-1-(2-fluoroethyl)-5-oxopyrrolidine-2-carboxamide) was synthesized from tert-butyl (S)-5-oxopyrrolidine-2-carboxylate, fluoroethylbromide, and 2-chloro-3-(trifluoromethyl)benzylamine with overall chemical yield 12% in three steps. The target tracer [18F]IUR-1601 ((S)-N-(2-chloro-3-(trifluoromethyl)benzyl)-1-(2-[18F]fluoroethyl)-5-oxopyrrolidine-2-carboxamide) was synthesized from desmethyl-GSK1482160 with 2-[18F]fluoroethyl tosylate, prepared from 1,2-ethylene glycol-bis-tosylate and K[18F]F/Kryptofix2.2.2, in two steps and isolated by HPLC combined with SPE in 1–3% decay corrected radiochemical yield. The radiochemical purity was >99%, and the molar activity at end of bombardment (EOB) was 74–370?GBq/μmol. The potency of IUR-1601 in comparison with GSK1482160 was determined by a radioligand competitive binding assay using [11C]GSK1482160, and the binding affinity Ki values for IUR-1601 and GSK1482160 are 4.31 and 5.14?nM, respectively.  相似文献   

10.
Analysis of prostaglandin F (PGF) in urine is a useful indicator of renal prostaglandin synthesis. A mass fragmentographic method for PGF analysis in human urine was developed using [3,3,4,4-2H4]PGF as an internal standard and carrier. PGF was extracted from urine (20 ml) with chloroform, purified by preparative thin-layer chromatography and converted to the methyl ester trimethylsilyl ether before analysis by gas chromatograph—mass spectrometry. The specificity of the urine analysis was demonstrated by retention time and the use of two pairs of fragments m/e 494/498 and 513/517 with the same results. The coefficient of variation for duplicate analysis averaged 12.6%, n = 17. Urine from recumbent women contained 4.9 ± 2.6 (S.D.) ng/ml or 4.1 ± 1.0 ng PGF per mg creatinine (n = 10) with little diurnal variation. Male urine contained 5.0 ± 2.7 (S.D.) ng/ml or 3.7 ± 2.1 ng/mg creatinine (n = 10). Similar concentrations were found in boys and in girls. These observations indicate that urinary PGF originates from the kidneys with little contribution from the male accessory sexual glands. This method can also be applied to analysis of PGF in rabbit urine.  相似文献   

11.
Summary Menadiol diphosphate was introduced as a new substrate for nonspecific alkaline phosphatase, following a search for new and less expensive substrates, which give a more sensitive response and are easily synthesized in the laboratory. Menadiol released by phosphatase action can be assayed by its reduction of tetrazolium salts, or it can be coupled with diazonium salts; alternatively, the phosphate can be trapped by metal ions. The synthesis and purification of menadiol diphosphate are described, and it was shown to be sufficiently stable for qualitative and semiquantitative histochemistry, as well as for the immunohistochemistry of enzymes and cytoskeletal proteins with nonspecific alkaline phosphatase as the enzyme label. For qualitative as well as semiquantitative histochemistry and immunohistochemistry, the best results were obtained by applying the method with nitro-blue tetrazolium (NBT) to acetone-chloroform pretreated cryostat sections. Tetranitro-blue tetrazolium (TNBT), benzothiazolylphthalhydrazidyl tetrazolium (BSPT) and various diazonium salts were less suitable. Fast Blue BB and VB produced satisfactory results. Ce3+ ions and the DAB−Ni−H2O2 procedure yielded better results than Ca2+ ions in the Co−(NH4)2S visualization method. The NBT method with menadiol diphosphate is superior to existing methods employing azo, azoindoxyl or tetrazolium salts and to metal precipitation methods. The Ce3+ technique and the NBT/menadiol diphosphate method give similar results, and appear to be of equal value. In qualitative histochemistry and immunohistochemistry the NBT/menadiol diphosphate method resulted in higher quantities of precisely localized stain. Semiquantitative histochemistry with minimal incubation revealed more favorable kinetics for the menadiol diphosphate method, especially when using NBT. Supported by the Alexander von Humboldt-Stiftung and the Deutsche Forschungsgemeinschaft (Sfb 174)  相似文献   

12.
HPLC–MS/MS methods for the determination of a Hepatitis C NS3/NS4 protease inhibitor (MK-7009) in human plasma and Tween-treated urine were developed and validated over the concentration range 1–1000 ng/mL and 0.2–100 μg/mL respectively. A stable isotope labeled internal standard (ISTD), D4-MK-7009, was employed. Analytes were chromatographed by reversed phase HPLC and quantified by an MS/MS system. Electrospray ionization in the positive mode was employed. Multiple reaction monitoring of the precursor to product ion pairs m/z 758.6  637.4 MK-7009 and m/z 762.5  637.4 ISTD was used for quantitation. Analyte and internal standard were extracted from 250 μL of plasma using an automated 96-well liquid–liquid extraction. Plasma pH adjustment prior to extraction minimized ionization suppression in plasma samples from patients with Hepatitis C. The urine method involved direct dilution in the 96-well format of 0.020 mL Tween-treated urine. These methods have supported several clinical studies. Incurred plasma sample reanalysis demonstrated adequate assay reproducibility and ruggedness.  相似文献   

13.
A new sensitive high-performance liquid chromatographic (HPLC) method with fluorescence detection was developed for the determination of 2-phenylethylamine (PEA) in human urine. The analytical procedure involved a simple extraction of the analyte from urine, followed by precolumn derivatisation of the sample with o-phthalaldehyde. The HPLC separation was performed under isocratic conditions using an Erbasil S C18 (250 × 4.0 mm I.D., particle size 3 μm) reversed-phase column. The limit of quantification was 0.5 ng of PEA/ml of urine. The method showed good linearity, accuracy and precision data in the concentration range 0.5–200 ng/ml of urine. The method was successfully applied to the determination of PEA urinary excretion in Parkinsonian patients after oral administration of the monoamine oxidase B (MAO-B) inhibitor, selegiline.  相似文献   

14.
Using an established high-performance liquid chromatography (HPLC) method based on anion exchange chromatography, fraction collection, and electrochemical detection, the oxidative DNA damage marker 8-hydroxy-2′-deoxyguanosine (8-OH-dG) can be analyzed rapidly and precisely in human urine samples. In addition, by ultraviolet (UV) detection, it was shown recently that it is possible to simultaneously analyze creatinine and 7-methylguanine (m7Gua), an RNA degradation product, in urine. By adding a fluorescence detector to the HPLC system, we now report that it is also possible to detect pteridins such as neopterin and biopterin. The fluorescence detection was evaluated in detail for neopterin, an immune response and tumor marker. The urinary content of neopterin, assessed by using the HPLC method, was verified with a commercial neopterin enzyme-linked immunosorbent assay (ELISA) kit as indicated by the high correlation between the two methods (r = 0.98). In urinary samples from 58 young healthy individuals (male and female nonsmokers, ages 19-39 years), it was found that there was no significant correlation (r = −0.04) between the levels of 8-OH-dG and neopterin (as normalized to urinary creatinine levels). In contrast, in urinary samples from 60 old healthy individuals (male and female nonsmokers, ages 60-86 years), there was a significant correlation (r = 0.47) found between the levels of 8-OH-dG and neopterin (as normalized to urinary creatinine levels). These findings strongly indicate that the higher level of immune response that was correlating with old age contributes significantly to the higher level of oxidative damage as assessed in the form of 8-OH-dG. Using this type of HPLC system, it is possible to evaluate oxidative DNA damage and immune response simultaneously using the respective urinary markers. These data may contribute to understanding of the pathophysiology of diseases such as infections and tumor progression where both oxidative stress and immune response occur simultaneously.  相似文献   

15.
In consideration of its relatively constant urinary excretion rate, creatinine in urine is a useful biochemical parameter to correct the urinary excretion rate of endogenous and exogenous biomolecules. Assays based on the reaction of creatinine and picric acid first reported by Jaffé in 1886 still belong to the most frequently used laboratory approaches for creatinine measurement in urine. Further analytical methods for creatinine include HPLC–UV, GC–MS, and LC–MS and LC–MS/MS approaches. In the present article we report on the development, validation and biomedical application of a new GC–MS method for the reliable quantitative determination of creatinine in human urine, plasma and serum. This method is based on the derivatization of creatinine (d0-Crea) and the internal standard [methyl-trideutero]creatinine (d3-Crea) with pentafluorobenzyl (PFB) bromide in the biological sample directly or after dilution with phosphate buffered saline, extraction of the reaction products with toluene and quantification in 1-μl aliquots of the toluene extract by selected-ion monitoring of m/z 112 for d0-Crea-PFB and m/z 115 for d3-Crea-PFB in the electron-capture negative-ion chemical ionization mode. The limit of detection of the method is 100 amol of creatinine. In an inter-laboratory study on urine samples from 100 healthy subjects, the GC–MS method was used to test the reliability of currently used Jaffé, enzymatic and HPLC assays in clinical and occupational studies. The results of the inter-laboratory study indicate that all three tested methods allow for satisfactory quantification of creatinine in human urine. The GC–MS method is suitable for use as a reference method for urinary creatinine in humans. In serum, creatine was found to contribute to creatinine up to 20% when measured by the present GC–MS method. The application of the GC–MS method can be extended to other biological samples such as saliva.  相似文献   

16.
This study investigated a means of quantifying urinary myoglobin using a novel reverse-phase high-performance liquid chromatography (RP–HPLC) method that is an alternative measure of exercise-induced muscle damage. It also investigated the effect of storage and alkalization on urinary myoglobin stability issues. An RP–HPLC method was validated by precision and repeatability experiments. Myoglobin stability was determined through spiked urine samples stored at various temperatures over an 8-week period using alkalization and dilution in a pH 7.0 buffer. The method was validated with urine collected from mixed martial arts fighters during a competition and training session. The method produced linearity from 5 to 1000 μg/ml (R2 = 0.997), intra- and inter-assay coefficients of variation from 0.32 to 2.94%, and a lower detection limit of 0.2 μg/ml in the final dilution and 2 μg/ml in the original urine sample. Recovery ranged from 96.4 to 102.5%, myoglobin remained stable at 4 °C when diluted in a pH 7.0 buffer after 20 h, and a significant increase (P < 0.01) and an identifiable peak were observed following a mixed martial arts contest and training session. Storage length and conditions had significant effects (P < 0.05) on stability. The method's simplicity and noninvasive nature means it can be used as an alternative muscle damage assay following exercise and trauma.  相似文献   

17.
Exogenous and endogenous oxidants constantly cause oxidative damage to DNA. Since the reactive oxidants itself are not suitable for analysis, oxidized bases like 8-hydroxy-2′-deoxyguanosine (8OHdG) are used as biomarkers for oxidative stress, either in cellular DNA or as elimination product in urine. A simple, fast and robust analytical procedure is described for urinary 8OHdG as an indicator of oxidative damage in humans. The adduct was purified from human urine by applying a single solid-phase extraction step on LiChrolut EN®. After evaporation of the eluate, the residue was resolved and an aliquote was injected into a HPLC system with a triple quadrupole mass spectrometer. The limit of detection was 0.2 ng ml−1 (7 fmol absolute) when using one product ion as quantifier and two further product ions as qualifier. The coefficient of variation was 10.1% (n=5 at 2.8 ng ml−1 urine). The sample throughput was about 50 samples a day. Thus, this method is more sensitive and much faster than the common method using HPLC with electrochemical detection. The results of a study with nine volunteers investigated at six time-points each over 5 days are presented. The mean excretion of 8OHdG was 2.1 ng mg−1 creatinine (range 0.17–5.9 ng mg−1 creatinine; 4 of 53 samples were below the LOD). A relatively large intra- (relative SD 66%) and inter-individual (relative SD 71%) variation in urinary 8OHdG excretion rates was found.  相似文献   

18.
Column chromatography and spectroscopy have been employed in analyzing pyrimidine derivatives obtained from alkaline-treated 7-methylguanosine (7-meGuo). High performance liquid chromatography (HPLC) revealed that the alkaline generated products consist predominantly of two forms of ring opened 7-methylguanine (rom7Gua) in equal amounts. Material from both Dowex 50 and Sephadex LH-20 columns was readily resolvable into two HPLC peaks. The species in one peak appears to be composed of formylated and that in the other of deformylated rom7Gua. The presence of a deformylated species is supported by the absence of radioactivity in one of the two peaks obtained when ring opened [8-14C]guanosine was analyzed by HPLC. The formylated species was retained on the liquid chromatography column for 8 min with a 3% methanol, 0.01 M NH4H2PO4 (pH 5.1) solvent and for 6 min with a 6% methanol, 0.01 N NH4H2PO4 (pH 5.1) solvent system; the deformylated species was retained for 6.3 min with the first solvent and 4.5 min with the second solvent. Subsequent to Dowex 50 chromatography in an ammonium formate solvent, about 90% of the material was formylated. When stored at 24°C for 72 h in a solvent without formate ions, the material was shown by HPLC to consist of equal amounts of the formylated and deformylated species. These results indicate that the two species of rom7Gua are in equilibrium. The rom7Gua excised from DNA by formamidopyrimidine (FAPy)-DNA glycosylase was shown to coelute with the formylated species.  相似文献   

19.
Abstract: Conantokin-G (con-G) is a 17-amino-acid polypeptide that acts as an N-methyl-d -aspartate (NMDA) antagonist. This action has been attributed to a specific but noncompetitive inhibition of the positive modulatory effects of polyamines at NMDA receptors. Con-G possesses several unusual structural features, including five γ-carboxyglutamate (Gla) residues and a high degree of helicity in aqueous media. Previous structure-activity studies indicated that one or more Gla residues are necessary for NMDA antagonist activity. Con-G analogues were synthesized with alanine (Ala), serine (Ser), and phosphoserine substituted for Gla to assess the contribution of individual Gla residues to biological activity and secondary structure. Replacement of Gla in positions 3 and 4 resulted in polypeptides with markedly reduced and no NMDA antagonist actions, respectively. In contrast, Gla residues in positions 7, 10, and 14 are not required for NMDA antagonist actions because the potencies of con-G analogues containing Ser7, Ser10, Ala14, and Ser14 to inhibit spermine-stimulated [3H]MK-801 binding are similar to the parent peptide. Moreover, the Ala7 derivative of con-G was about fourfold more potent than the parent peptide both as an inhibitor of spermine-stimulated increases in [3H]MK-801 binding (IC50 of ~45 nM) and in reducing NMDA-stimulated increases in cyclic GMP levels (IC50 of ~77 nM) in cerebellar granule cell cultures. Although con-G and its analogues assumed mixtures of 310 and α-helices, no clear-cut relationship was evinced between the NMDA antagonist properties of these peptides and the degree of helicity they assumed in aqueous solutions. Together with the inability of con-G to affect 5,7-dichloro[3H]kynurenic acid, [3H]CGP-39653, and [3H]ifenprodil binding, these data are consistent with the hypothesis that this polypeptide acts at a unique, polyamine-associated site on NMDA receptors.  相似文献   

20.
Diphenylmethoxyacetic acid (DPMA) is a major metabolite of diphenhydramine in monkeys, dogs, and humans. The metabolic fate of diphenhydramine (DPHM) in sheep is not yet well understood; however, preliminary studies have demonstrated the presence of DPMA in the plasma and urine of sheep following an intravenous bolus of DPHM. Our current studies employ the simultaneous intravenous co-administration of DPHM and the stable isotope analog of DPHM to investigate the pharmacokinetics of DPHM in sheep. In these studies, in order to investigate the pharmacokinetics of the DPMA metabolite, measurement of both unlabeled and stable-isotope labeled DPMA is required. Thus, a stable isotope analog of DPMA ([2H10]DPMA) was synthesized, characterized, and purified for use as an analytical standard. The quantitative method for the gas chromatography—electron-impact mass spectrometry (GC—EI-MS) analysis of DPMA and [2H10]DPMA used a single step liquid-liquid extraction procedure using toluene for sample cleanup. The samples were derivatized with N-methyl-N-(tert.-butyldimethylsilyl) trifluoroacetamide. A 1.0-μl aliquot of the prepared sample was injected into the GC-MS system and quantitated using selected-ion monitoring (SIM). One ion was monitored for each compound, namely, m/z 165 for the internal standard diphenylacetic acid, m/z 183 for DPMA, and m/z 177 for [2H10]DPMA. The ion chromatograms were free from chromatographic peaks co-eluting with the compound of interest. The calibration curve was linear from 2.5 ng/ml (limit of quantitation) to 250.0 ng/ml in both urine and plasma. The intra-day and inter-day variabilities of this assay method were within acceptable limits (below 20% at the limit of quantitation and below 10% at all other concentrations). This method was used to measure the concentration of DPMA and [2H10]DPMA in plasma and urine samples from a ewe in which equimolar amounts of DPHM and [2H10]DPHM were administered by an intravenous bolus dose via the femoral vein. DPMA appeared to persist longer in the plasma and the urine as compared to DPHM. This method is robust and reliable for the quantitation of DPMA and [2H10]DPMA in biological samples obtained from sheep (e.g. plasma and urine).  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