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 556 毫秒
1.
The pharmacokinetics of ibuprofen enantiomers were investigated in a crossover study in which seven healthy male volunteers received single oral doses of 800 mg racemic ibuprofen as a soluble granular formulation (sachet) containing L-arginine (designated trade name: Spedifen®), 400 mg (-)R-ibuprofen arginine or 400 mg (+)S-ibuprofen arginine. Plasma levels of both enantiomers were monitored up to 480 minutes after drug intake using an enantioselective analytical method (HPLC with ultraviolet detection) with a quantitation limit of 0.25 mg/l. Substantial inter-subject variability in the evaluated pharmacokinetic parameters was observed in the present study. After (+)S-ibuprofen arginine, the following mean pharmacokinetic parameters ±SD were calculated for (+)S-ibuprofen: tmax 28.6 ± 28.4 min; Cmax 36.2 ± 7.7 mg/l; AUC 86.4 ± 14.9 mg · h/l; t½ 105.2 ± 20.4 min. After (-)R-ibuprofen arginine, the following mean pharmacokinetic parameters were calculated for (+)S-ibuprofen and (-)R-ibuprofen, respectively: tmax 90.0 ± 17.3 and 50.5 ± 20.5 min; Cmax 9.7 ± 3.0 and 35.3 ± 5.0 mg/l; AUC 47.0 ± 17.2 and 104.7 ± 27.7 mg · h/l; t½ 148.1 ± 63.6 and 97.7 ± 23.3 min. After racemic ibuprofen arginine, the following mean pharmacokinetic parameters were calculated for (+)S- and (-)R-ibuprofen, respectively: tmax 30.7 ± 29.1 and 22.9 ± 29.8 min.; Cmax 29.9 ± 5.6 and 25.6 ± 4.4 mg/l; AUC 105.1 ± 23.0 and 65.3 ± 15.0 mg · h/l; t½ 136.6 ± 20.7 and 128.6 ± 45.0 min. Tmax values of S(+)- and (-)R-ibuprofen after a single dose of 400 mg of each enantiomer did not differ significantly from the corresponding parameters obtained after a single dose of 800 mg of racemic ibuprofen arginine, indicating that the absorption rate of (-)R- and (+)S-ibuprofen is not different when the two enantiomers are administered alone or as a racemic compound. An average of 49.3 ± 9.0% of a dose of the (-)R-ibuprofen arginine was bioinverted into its antipode during the study period (480 minutes post-dosing). The percent bioinversion during the first 30 minutes after (-)R-ibuprofen arginine intake averaged 8.1 ± 3.9%. The mean AUC of (+)S-ibuprofen calculated after 800 mg racemic ibuprofen arginine (105.1 ± 23.0 mg · h/l) was lower than the mean AUC value obtained by summing the AUCs of (+)S-ibuprofen after administration of 400 mg (+)S-ibuprofen arginine and 400 mg (-)R-ibuprofen arginine (133.4 ± 26.6 mg · h/l). In conclusion, the administration of Spedifen® resulted in very rapid absorption of the (+)S-isomer (eutomer) with tmax values much lower than those observed for this isomer when conventional oral solid formulations such as capsules or tablets of racemic ibuprofen are administered. This characteristic is particularly favourable in those conditions in which a very rapid analgesic effect is required. Chirality 9:297–302, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

2.
The dissolution characteristics of propranolol enantiomers from tablet formulations containing cellulose, or one of eight cellulose derivatives, were determined under a range of conditions. The derivatives examined were: cellulose tris(phenylcarbamate) (1), cellulose tris(2,3-dichlorophenylcarbamate) (2), cellulose tris 2,4-dichlorophenylcarbamate (3), cellulose tris(2,6-dichlorophenylcarbamate) (4), cellulose tris(2,3-dimethylphenylcarbamate) (5), cellulose tris(3,4-dichlorophenylcarbamate) (6), cellulose tris (3,5-dichlorophenylcarbamate) (7), cellulose tris(3,5-dimethylphenylcarbamate) (8). In water at 25°C, the release rates of (-)R-propranolol were generally greater than those of (-)-S-propranolol, although these differences were not always statistically significant; only compounds 5 and 8 demonstrated significant enantioselectivity. Using compound 8 in further experiments, statistically significant stereoselective dissolution of propranolol HCl was observed in buffer pH 7.4 at 25°C (intrinsic dissolution rates: 0.41 ± 0.01 mgcm2min−1 for R-propranolol and 0.30 ± 0.02 mgcm2min−1 for S-propranolol; P = 0.003). The cumulative amounts of enantiomers released at every time point were also found to be statistically significant (mean ratio R:S 1.25 ± 0.05). The observed low stereoselectivity of 8 with propranolol base was probably attributable to low solubility in pH 7.4 buffer, although stereoselective release did increase with time. This suggested that there is a relationship between stereoselectivity and contact time in an aqueous environment. Results also suggested that increased temperature may affect the release process as well as stereoselective interactions of 8 with individual enantiomers. To conclude, differential release of rac-propranolol from cellulose derivative matrices has been demonstrated, which supports the principle of stereoselective retardation as a potential means of stereoselective drug delivery for solid dosage forms. Chirality 9:307–312, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

3.
The stability (half-life, t½) of the large catalase (CAT) isolated from Aspergillus terreus was decreased under acidic conditions (maximum t½ ~8.5 months at pH ≤ 6) versus alkaline conditions (t½ ~15 months at pH 8–12). Acidic conditions induce the dissociation of haem from CAT, as revealed from a reduction in the Soret peak intensity at 405 nm and an increase in the peak current at Fe3+/Fe2+ redox potentials. This increase in current is attributed to the facile electron transfer from the free haem generated on the electrode surface as a result of its disintegration from the insulating protein matrix. The haem isolated from CAT at acidic condition was reconstituted with apo-CAT at alkaline denaturing conditions to regenerate the CAT activity.  相似文献   

4.
The nonsteroidal antiinflammatory drug oxindanac exists as two enantiomers, with most of its pharmacological activity residing in the (S)-isomer. The behavior of its enantiomers was investigated in dogs. Bidirectional inversion occurred in heparinised plasma and blood, with a ratio of enantiomers [S:R] of 7.3:1 being achieved at equilibrium after incubation for 24 h at 37°C. There was no detectable inversion of either isomer in plasma incubated at 4°C for up to 8 h or in aqueous solution at 37°C for up to 36 h. Bidirectional inversion also occurred in vivo, with a ratio of plasma AUC (0 ∞)s [S:R] of 8.1:1. The ratio of enantiomers reached equilibrium within 2 hr following (S)- or rac-oxindanac, and within 8 h following (R)-oxindanac. Elimination t½s of the isomers were the same (R, 12.1 h, S, 13.3 h). There were no differences in the ratio of enantiomers following oral or intravenous application, suggesting that a systemic site for inversion was predominant. Although concentrations of the respective isomers were similar at equilibrium following administration of either (R)-, (S)-, or rac-oxindanac, AUC (0 ∞)s differed due to the delay in reaching equilibrium. The extent of inversion to the (S)-isomer was 100, 73.2, and 60.7% after administration of (S)-, rac-, and (R)-oxindanac, respectively. Although pharmacological activity might be equivalent at equilibrium following administration of either (R)-, (S)-, or rac-oxindanac; efficacy at early time points should be superior in the order (S) > racemate > (R). In conclusion both enantiomers of oxindanac undergo conversion to their respective antipodes in dogs, although the inversion of R to S is more efficient than that of S to R. This bidirectional inversion occurred in vivo, and in vitro in plasma and blood. © 1994 Wiley-Liss, Inc.  相似文献   

5.
Acyl glucuronides are known to be labile conjugates, which undergo hydrolysis and bind irreversibly to proteins. The lipid-regulating agent (±)-beclobrate is immediately converted to the free acid after oral administration. Further metabolism leads to formation of the corresponding diastereomeric acyl glucuronides. Beclobric acid glucuronides were quantified by indirect measurement with an HPLC method based on chiral fluorescent derivatization of the carboxylic acid and subsequent normal-phase chromatography. The renal clearance of unchanged drug is low, with almost all drug excreted into urine as glucuronic acid conjugates. Beclobric acid glucuronide is also detectable in plasma. In vitro degradation studies with beclobric acid glucuronide (at a concentration of 5 μM in 150 mM phosphate buffer pH 7.4) exhibited a minor tendency for acyl migration and hydrolysis, i.e., a higher stability than has been observed for the acyl glucuronides of most other drugs. The in vitro degradation half-lives of the two beclobric acid β-1-O-acyl glucuronides were 22.7 and 25.7 h. After incubation with pooled plasma and human serum albumin in buffer pH 7.4 irreversible binding was measured in vitro. No significant difference between the two enantiomers was detected with respect to the magnitude of in vitro irreversible binding. In 3 healthy male volunteers the extent of irreversible binding of both beclobric acid enantiomers to plasma proteins was investigated after single and multiple oral doses of racemic beclobrate (100 mg once daily). Irreversible binding of both enantiomers was observed in all volunteers. The adduct densities for (?)- and (+)-beclobric acid after single 100 mg beclobrate doses were 0.147 × 10?4 and 0.177 × 10?4 mol/mol protein. Multipie dosing increased irreversible binding 3- to 4-fold. © 1993 Wiley-Liss, Inc.  相似文献   

6.
A stereoselective and sensitive achiral/chiral method for the determination of terfenadine acid metabolite in human plasma was developed. The metabolite was separated and quantitated using an achiral chromatographic procedure with a cyano column. The mobile phase was 1 mM sodium acetate buffer (pH 4.0) and acetonitrile (25:75% v/v) at a flow rate of 2 ml/min, at ambient temperature. The stereospecific resolution was accomplished using a chiral-AGP column and a mobile phase consisting of sodium acetate (0.01 M): methanol (98.7:1.3% v/v), and 20 mM di-n-butylamine at a flow rate of 1.2 ml/min. The column temperature was maintained at 32°C. The eluent was monitored at 230 nm (excitation) and 300 nm (emission) with a cut-off filter at 270 nm. This assay was used for a pharmacokinetic study in five subjects after administration of a single dose of 60 mg of terfenadine. The t½ values of the two enantiomers were similar, but the AUC values of the (+)-enantiomer were 2.05–2.35 times higher than those of (?)-enantiomer. © 1994 Wiley-Liss, Inc.  相似文献   

7.
Abstract

A group of 5′-O-myristoyl analogue derivatives of FLT (2) were evaluated as potential anti-HIV agents that were designed to serve as prodrugs to FLT. 3′-Fluoro-2′,3′-dideoxy-5′-O-(12-methoxydodecanoyl)thymidine (4) (EC50 = 3.8 nM) and 3′-fluoro-2′,3′-dideoxy-5′-O-(12-azidododecanoyl)thymidine (8) (EC50 = 2.8 nM) were the most effective anti-HIV-1 agents. There was a linear correlation between Log P and HPLC Log retention time for the 5 ′-O-FLT esters. The in vitro enzymatic hydrolysis half-life (t½), among the group of esters (3–8) in porcine liver esterase, rat plasma and rat brain homogenate was longer for 3′-fluoro-2′,3′-dideoxy-5 ′-O-(myristoyl)thymidine (7), with t½ values of 20.3, 4.6 and 17.5 min, respectively.  相似文献   

8.
The steady-state pharmacokinetics in serum and urine of the enantiomers of citalopram and its metabolites, demethylcitalopram (DCT) and didemethylcitalopram (DDCT), were investigated after multiple doses of rac-citalopram for 21 consecutive days (40 mg per day) to healthy human subjects who were extensive metabolisers of sparteine and mephenytoin. Comparable pharmacokinetic variability was noted for (+)-(S)-, (−)-(R)- and rac-citalopram. Enantiomeric (S/R) serum concentration ratios for citalopram were always less than unity and were constant during the steady-state dosing interval. A modest, but statistically significant, stereoselectivity in the disposition of citalopram and its two main metabolites was observed. Serum levels of the (+)-(S)-enantiomers of citalopram, DCT, and DDCT throughout the steady-state dosing interval investigated were 37 ± 6%, 42 ± 3% and 32 ± 3%, respectively, of their total racemic serum concentrations. The (+)-(S)-enantiomers of citalopram, DCT, and DDCT were eliminated faster than their antipodes. For (−)-(R)- and (+)-(S)-citalopram, respectively, the serum t½ averaged 47 ± 11 and 35 ± 4 h and AUCss averaged 4,193 ± 1,118 h · nmol/l and 2,562 ± 1,190 h · nmol/l. The observed enantiospecificities were apparently more related to clearance, rather than to distributional mechanisms. Chirality 9:686–692, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

9.
Reboxetine, (RS)-2-[(RS)-α-(2-ethoxyphenoxy)benzyl]morpholine methanesulphonate, is a racemic compound and consists of a mixture of the (R,R)- and (S,S)-enantiomers. The pharmacokinetics of reboxetine enantiomers were determined in a crossover study in three male beagle dogs. Each animal received the following oral treatments, separated by 1-week washout period: 10 mg/kg reboxetine, 5 mg/kg (R,R)- and 5 mg/kg (S,S)-. Plasma and urinary levels of the reboxetine enantiomers were monitored up to 48 h post-dosing using an enantiospecific HPLC method with fluorimetric detection (LOQ: 1.1 ng/ml in plasma and 5 ng/ml in urine for each enantiomer). After reboxetine administration mean tmax was about 1 h for both enantiomers. Cmax and AUC were about 1.5 times higher for the (R,R)- than for the (S,S)-enantiomer, mean values ± SD being 704 ± 330 and 427 ± 175 ng/ml for Cmax and 2,876 ± 1,354 and 1,998 ± 848 ng.h/ml for AUC, respectively. No differences between the (R,R)- and (S,S)-enantiomers were observed in t½ (3.9 h). Total recovery of the two enantiomers in urine was similar, the Ae (0–48 h) being 1.3 ± 0.7 and 1.1 ± 0.7% of the enantiomer dose for the (R,R)- and the (S,S)-enantiomers, respectively. No marked differences in the main plasma pharmacokinetic parameters were found for either enantiomer on administration of the single enantiomers or reboxetine. No chiral inversion was observed after administration of the separate enantiomers, as already observed in humans. Chirality 9:303–306, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

10.
Abstract

Environmental and biological monitoring was carried out in the winter season of 2004 for 30 gasoline station workers (study subjects) and 30 office workers (controls) of Bangalore city, India. Personal air sampling was carried out in the breathing zone of workers using an Anasorb CSC sorbent tube (SKC 226-01) fitted to the low-flow personal samplers (PCXR4 and pocket pump Model No. 210-1002) at a flow rate of 200 ml min?1 during the shift work of 8 h. The benzene content adsorbed in the sorbent tube (SKC 226-01) was desorbed with 1 ml of benzene-free carbon disulfide on a developing vibrator and later analysed by Trace GC fitted with MXT-624 column and flame ionization detector. The mean time weighted average benzene concentration found among study and controls was 1.10±1.08 and 0.070±0.035 mg m?3, respectively. Biological monitoring for benzene exposure was performed by measuring trans,trans muconic acid (t,t-MA) in the end shift urine samples using HPLC-UV technique. End-shift urine samples (1 ml) were adjusted to pH 7–9 with phosphate buffer pH 7.4 passed through the preconditioned Q-SAX anion-exchange cartridge and the (t,t-MA) is extracted with 10% acetic acid and later analysed by HPLC-UV detection The mean t,t-MA found among study and controls were 563.16±281.81 and 266.88±110.65 µg g?1 creatinine. About 50% of the study subjects (15) have higher t,t-MA values than the biological exposure index of the American Conference of Government Industrial Hygienist (ACGIH). Correlation is significant at 5% level (p<0.05) between personal air benzene concentration and urinary t,t-MA in the study group. Based on these findings, the t,t-MA can be used as a biomarker for benzene exposure.  相似文献   

11.
Abstract: Transport and permeability properties of the blood-brain and blood-CSF barriers were determined by kinetic analysis of radioisotope uptake from the plasma into the CNS of the adult rat. Cerebral cortex and cerebellum uptake curves for 36Cl and 22Na were resolved into two components. The fast component (t½ 0.02–0.05 h, fractional volume 0.04–0.08) is comprised of the vascular compartment and a small perivascular space whereas the slow component (t½ 1.06–1.69 h, fractional volume 0.92–0.96) represents isotope movement across the blood-brain barrier into the brain extracellular and cellular compartments. Uptake curves of both 36Cl and 22Na into the CSF were also resolved into two components, a fast component (t½ 0.18 h, fractional volume 0.24) and a slow component (t½ 1.2 h, fractional volume 0.76). Evidence suggests that the fast component represents isotope movement across the blood-CSF barrier, i.e., the choroid plexuses, whereas the CSF slow component probably reflects isotope penetration primarily from the brain extracellular fluid into the CSF. The extracellular fluid volume of the cerebral cortex and cerebellum was estimated as ?13% from the initial slope of the curve of brain space versus CSF space curve for both 36Cl and 22Na. Like the choroid plexuses, the glial cell compartment of the brain appears to accumulate Cl from 2 to 6 times that predicted for passive distribution. The relative permeability of the blood-CSF and blood-brain barriers to 36Cl, 22Na, and [3H]mannitol was determined by calculating permeability surface-area products (PA). Analysis of the PA values for all three isotopes indicates that the effective permeability of the choroidal epithelium (blood/CSF barrier) is significantly greater than that of the capillary endothelium in the cerebral cortex and cerebellum (blood-brain barrier).  相似文献   

12.
Context: Ropivacaine (RVC) is an aminoamide local anesthetic widely used in surgical procedures. Studies with RVC encapsulated in liposomes and complexed in cyclodextrins have shown good results, but in order to use RVC for lengthy procedures and during the postoperative period, a still more prolonged anesthetic effect is required.

Objective: This study therefore aimed to provide extended RVC release and increased upload using modified liposomes.

Materials and methods: Three types of vesicles were studied: (i) large multilamellar vesicle (LMV), (ii) large multivesicular vesicle (LMVV) and (iii) large unilamellar vesicle (LUV), prepared with egg phosphatidylcholine/cholesterol/α-tocopherol (4:3:0.07?mol%) at pH 7.4. Ionic gradient liposomes (inside: pH 5.5, pH 5.5?+?(NH4)2SO4 and pH 7.4?+?(NH4)2SO4) were prepared and showed improved RVC loading, compared to conventional liposomes (inside: pH 7.4).

Results and discussion: An high-performance liquid chromatography analytical method was validated for RVC quantification. The liposomes were characterized in terms of their size, zeta potential, polydispersion, morphology, RVC encapsulation efficiency (EE(%)) and in vitro RVC release. LMVV liposomes provided better performance than LMV or LUV. The best formulations were prepared using pH 5.5 (LMVV 5.5in) or pH 7.4 with 250?mM (NH4)2SO4 in the inner aqueous core (LMVV 7.4in?+?ammonium sulfate), enabling encapsulation of as much as 2% RVC, with high uptake (EE(%) ~70%) and sustained release (~25?h).

Conclusion: The encapsulation of RVC in ionic gradient liposomes significantly extended the duration of release of the anesthetic, showing that this strategy could be a viable means of promoting longer-term anesthesia during surgical procedures and during the postoperative period.  相似文献   

13.
Co‐occurring ocean warming, acidification and reduced carbonate mineral saturation have significant impacts on marine biota, especially calcifying organisms. The effects of these stressors on development and calcification in newly metamorphosed juveniles (ca. 0.5 mm test diameter) of the intertidal sea urchin Heliocidaris erythrogramma, an ecologically important species in temperate Australia, were investigated in context with present and projected future conditions. Habitat temperature and pH/pCO2 were documented to place experiments in a biologically and ecologically relevant context. These parameters fluctuated diurnally up to 10 °C and 0.45 pH units. The juveniles were exposed to three temperature (21, 23 and 25 °C) and four pH (8.1, 7.8, 7.6 and 7.4) treatments in all combinations, representing ambient sea surface conditions (21 °C, pH 8.1; pCO2 397; ΩCa 4.7; ΩAr 3.1), near‐future projected change (+2–4 °C, ?0.3–0.5 pH units; pCO2 400–1820; ΩCa 5.0–1.6; ΩAr 3.3–1.1), and extreme conditions experienced at low tide (+4 °C, ?0.3–0.7 pH units; pCO2 2850–2967; ΩCa 1.1–1.0; ΩAr 0.7–0.6). The lowest pH treatment (pH 7.4) was used to assess tolerance levels. Juvenile survival and test growth were resilient to current and near‐future warming and acidification. Spine development, however, was negatively affected by near‐future increased temperature (+2–4 °C) and extreme acidification (pH 7.4), with a complex interaction between stressors. Near‐future warming was the more significant stressor. Spine tips were dissolved in the pH 7.4 treatments. Adaptation to fluctuating temperature‐pH conditions in the intertidal may convey resilience to juvenile H. erythrogramma to changing ocean conditions, however, ocean warming and acidification may shift baseline intertidal temperature and pH/pCO2 to levels that exceed tolerance limits.  相似文献   

14.
Context: The acylglycerides of lipoic and dihydrolipoic acids may serve as slow-release sources for cutaneous delivery of these antioxidants when formulated in a liposomal vehicle.

Objective: Testing was conducted to determine the storage stability of lipoyl glycerides in phospholipid-based liposomes.

Materials and methods: Lipoyl glycerides prepared by transesterification of lipoic acid with high oleic sunflower oil were incorporated into unilamellar liposomes comprised of soy phosphatidylcholine (soyPC) or dioleoylphosphatidylcholine (DOPC).

Results: Lipoyl glycerides were stable in soyPC at 4?°C (90% remaining after five weeks) and decayed with a half-life (t½) of 14?d at 40?°C. In contrast, lipoyl glycerides embedded in DOPC were completely stable for four weeks at 40?°C. Dihydrolipoyl glycerides in soyPC converted to lipoyl glycerides at 4?°C (t½?=?14?d) over four weeks, and much more rapidly so at 40?°C (t½?=?1?d). A hydroperoxide accumulation analysis indicated that lipoyl glycerides and dihydrolipoyl glycerides were modified or degraded while suppressing autoxidation of the polyunsaturated fatty acids present in soyPC. Dynamic light scattering measurements found that liposomes containing lipoyl glycerides or dihydrolipoyl glycerides did not undergo significant size changes for at least 48?d, indicating that inclusion of the lipoic acid derivatives did not induce vesicle aggregation.

Discussion/Conclusion: Substitution of the soyPC with DOPC, which is not readily subject to autoxidation, provided a much more stable storage environment for lipoyl glycerides. These findings confirm the expectation that phospholipid liposomes need to be oxidatively stable vehicles for dermal delivery of lipoic acid derivatives.  相似文献   

15.
Neuronal nitric oxide synthase (nNOS) was purified on DEAE-Sepharose anion-exchange in a 38% yield, with 3-fold recovery and specific activity of 5 µmol.min?1.mg?1. The enzyme was a heterogeneous dimer of molecular mass 225?kDa having a temperature and pH optima of 40°C and 6.5, Km and Vmax of 2.6 μM and 996 nmol.min?1.ml?1, respectively and was relatively stable at the optimum conditions (t½?=?3?h). β-Amyloid peptide fragments Aβ17–28 was the better inhibitor for nNOS (Ki?=?0.81 µM). After extended incubation of nNOS (96?h) with each of the peptide fragments, Congo Red, turbidity and thioflavin-T assays detected the presence of soluble and insoluble fibrils that had formed at a rate of 5?nM.min?1. A hydrophobic fragment Aβ17–21 [Leu17 – Val18 – Phe19 – Phe20 – Ala21] and glycine zipper motifs within the peptide fragment Aβ17–35 were critical in binding and in fibrillogenesis confirming that nNOS was amyloidogenic catalyst.  相似文献   

16.
Recombinant penicillin acylase from Streptomyces lavendulae was covalently bound to epoxy-activated Sepabeads EC-EP303®. Optimization of the immobilization process led to a homogeneous distribution of the enzyme on the support surface avoiding the attachment of enzyme aggregates, as shown by confocal electron microscopy. The optimal immobilized biocatalyst had a specific enzymatic activity of 26.2IUgwetcarrier?1 in the hydrolysis of penicillin V at pH 8.0 and 40°C. This biocatalyst showed the highest activity at pH 8.5 and 65°C, 1.5 pH units lower and 5°C higher than its soluble counterpart. Substrate specificity of the derivative also showed its ability to efficiently hydrolyze other natural aliphatic penicillins such as penicillins K, F and dihydroF. The immobilized enzyme was highly stable at 40°C and pH 8.0 (t1/2=625 h vs. t1/2=397 h for the soluble enzyme), and it could be recycled for at least 30 consecutive batch reactions without loss of catalytic activity.  相似文献   

17.
Summary Sulphate uptake by rabbit ileal brush border membrane vesicles was stimulated by a transmembrane sodium gradient ([Na+] o >[Na+] i ), but not by a similar potassium gradient.35SO 4 2– influx (J oi SO4 ) from outside (o) to inside (i) these vesicles was a hyperbolic function of [SO 4 2– ] o and the affinity constant for anion transport was strongly influenced by [Na+] o (100mm Na+,K t SO4 =0.52mm SO 4 2– ; 10mm Na+,K t SO4 =4.32mm SO 4 2– ).J t SO4 was a sigmoidal function of [Na+] o at pH 7.4 for both low (0.2m) and high (4.0mm) [SO 4 2– ] o . The Na+-dependency ofJ t SO4 was examined at pH 6.0, 7.4, and 8.0 (same pH inside and outside). At pH 6.0 and 7.4 sigmoidal Na+-dependentJ t SO4 exhibited nonlinear Eadie-Hofstee plots indicative of a transport mechanism capable of binding a variable number of sodium ions over the [Na+] o range used. Hill plots of anion transport under these conditions displayed slopes near unity at low [Na+] o and slopes approximating 2.0 at higher cation concentrations. At pH 8.0, Na+-dependentJ t SO4 was hyperbolic and showed linear Eadie-Hofstee and Hill plots, the latter with a single slope near 1.0. When a H+ gradient was imposed across the vesicle wall (pH i =8.0, pH o =6.0), Na+-dependentJ t SO4 was hyperbolic and significantly increased at each [Na+] o over values observed using bilateral pH 8.0. In contrast, a H+ gradient oriented in the opposite direction (pH i =6.0, pH o =8.0) led to Na+-dependentJ t SO4 that was sigmoidal and significantly lower at each [Na+] o than values found using bilateral pH 6.0. Electrogenicity ofJ t SO4 at pH 8.0 for both high and low [Na+] o was demonstrated by using a valinomycin-induced transmembrane electrical potential difference. At pH 6.0, electrogenicJ t SO4 occurred only at low [Na+] o (5mm); anion transfer was electroneutral at 50mm Na+. A model is proposed for proton regulation of sodium sulphate cotransport where flux stoichiometry is controlled by [H+] i and sodium binding affinity is modified by [H+] o . Preliminary experiments with rabbit proximal tubular brush border membrane vesicles disclosed similarJ t SO4 kinetic properties and a common transport mechanism may occur in both tissues.  相似文献   

18.
Abstract— Kinetic parameters of the sulphotransferase reaction in rat brain were investigated in vitro at pH 7.4. Evidence is presented that the enzyme phenol sulphotransferase (EC 2.8.2.1) can be assayed with 4-methylumbelliferone or 3-methoxy-4-hydroxyphenylethyleneglycol as the substrate. Both assays give identical Vmax values, whereas Km values are 0.026 mm and 0.039 mm , respectively. Normetanephrine, metanephrine and the catecholamines adrenaline and dopamine, having a positive charge on the side chain at pH 7.4, do not inhibit 4-methylumbelliferone and 3-methoxy-4-hydroxyphenylethy-leneglycol sulphotransferase at this pH. Their deaminated metabolites 3,4-dihydroxyphenylethyleneglycol, 3,4-dihydroxymandelic acid, 3,4-dihydroxyphenylacetic acid, 3-methoxy-4-hydroxyphenylethylene glycol, 3-methoxy-4-hydroxyphenethanol and 3-methoxy-4-hydroxyphenylacetic acid inhibit both the enzyme activities. The type of inhibition is noncompetitive with the exception of 3-methoxy-4-hydroxy-phenylethyleneglycol, which is a competitive inhibitor of 4-methylumbelliferone sulphation. 3-Methoxy-4-hydroxy-mandelic acid does not inhibit the enzyme activities. It is concluded that the catecholamines themselves are not sulphated by rat brain in vitro at pH 7.4.  相似文献   

19.
Iron uptake by rat reticulocytes is blocked by 20 mM NH4Cl, while 125I-diferric transferrin (Tf) uptake is relatively unaffected. At pH 5.0 both apo- and diferric Tf bind with high affinity; at pH 7.4 diferric Tf binds avidly, but apoTf binds very poorly. The dissociation rate (4°C) of diferric Tf is extraordinarily slow at pH 5.0 (extrapolated t12 = 32 hrs) and faster at pH 7.4 (t12 = 101 min). At pH 5.0 apoTf also dissociates slowly (t12 = 205 min), but at pH 7.4 apoTf exhibits a much faster dissociation rate (t12 = 62 min). 20 mM NH4Cl slows the release of Tf from cells at 37°C, but the rate of externalization of ligand is unaffected. Ligand dissociation at 37° involves both externalization of receptor-ligand complexes and receptor-ligand separation; the NH4Cl effect may result from an increased fraction of externalized Tf in the differric form which may dissociate more slowly. Receptor-mediated movement of Tf through acid intracellular compartments provides a mechanism to remove iron from Tf and for apoTf to remain receptor-bound for externalization to the cell surface and subsequent dissociation.  相似文献   

20.
The purpose of this study was to investigate whether CYP2C19 activity can be estimated from plasma concentrations of lansoprazole enantiomers 4 h (C4h) after single administration by oral and enteral routes. Sixty‐nine subjects, 22 homozygous extensive metabolizers (homEMs), 32 heterozygous EMs (hetEMs), and 15 poor metabolizers (PMs), participated in the study. After a single oral or enteral dose of racemic lansoprazole (30 mg), plasma concentrations of lansoprazole enantiomers were measured 4 h postdose. The R/S ratio of lansoprazole at 4 h differed significantly among the three groups (P < 0.0001) regardless of the administration route. The R/S ratio of lansoprazole in CYP2C19 PMs ranged from 3.0 to 13.7, whereas in homEMs and hetEMs the ratio ranged from 8.6 to 90 and 2.1 to 122, respectively. The relationship between (S)‐lansoprazole concentration and R/S ratio of lansoprazole at C4h is given by the following formula: log10 [R/S ratio] = 2.2 – 0.64 × log10 [C4h of (S)‐lansoprazole] (r = 0.867, P < 0.0001). Thus, phenotyping CYP2C19 using the R/S enantiomer ratio of lansoprazole seems unlikely. However, to obtain a pharmacological effect similar to that in CYP2C19 PMs, we can presume that lansoprazole has a sufficient effect in the patient with an R/S enantiomer ratio at 4 h ≤ 13.70 and (S)‐lansoprazole concentration at 4 h ≥ 50 ng/ml. Chirality 2010. © 2009 Wiley‐Liss, Inc.  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