首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Objective: A series of studies was undertaken to determine the cytochrome P450 isoform(s) involved in naproxen demethylation and whether this included the same isoforms reported to be involved in the metabolism of other NSAIDs. Methods: (S)-Naproxen was incubated with human liver microsomes in the presence of a NADPH-generating system and the formation of desmethylnaproxen was measured by high-performance liquid chromatography (HPLC). To further clarify the specific isoforms involved, experiments were conducted with preparations expressing only a single P450 isoform (vaccinia virus-expressed cells and microsomes derived from a lymphoblastoid cell line, each transfected with specific P450 cDNAs) as well as inhibition studies using human liver microsomes and putative specific P450 inhibitors. Results: In human liver microsomes (n=7), desmethylnaproxen formation was observed with a mean kM of 92 (21) μmol · l−1, Vmax of 538 pmol · min−1 · mg−1 protein and Cint2 (reflective of a second binding site) of 0.36 μl · min−1 · mg−1 protein. This Cint2 term was added since Eadie-Scatchard analysis suggested the involvement of more than one enzyme. Studies using putative specific P450 inhibitors demonstrated inhibition of this␣reaction by sulfaphenazole, (apparent Ki= 1.6 μmol · l−1), warfarin (apparent Ki=27 μmol · l−1), piroxicam (apparent Ki=23 μmol · l−1) and tolbutamide (apparent Ki=128 μmol · l−1). No effect was observed when α-naphthoflavone and troleandomycin were employed as inhibitors, but reaction with furafylline produced, on average, a maximum inhibition of 23%. At a naproxen concentration of 150 μmol · l−1, formation of desmethylnaproxen was observed in cells expressing P450 1A2, 2C8, 2C9 and its allelic variant 2C9R144C. To further characterize these reactions, saturation kinetics experiments were conducted for the P450s 1A2, 2C8 and 2C9. The kM and Vmax for P450 1A2 were 189.5 μmol · l−1 and 7.3 pmol · min−1 · pmol−1 P450, respectively. Likewise, estimates of kM and Vmax for P450 2C9 were 340.5 μmol · l−1 and 41.4 pmol · min−1 · pmol−1 P450, respectively. Reliable estimates of kM and Vmax could not be made for P450 2C8 due to the nonsaturable nature of the process over the concentration range studied. Conclusion: Multiple cytochrome P450 isoforms (P450 1A2, 2C8 and 2C9) appear to be involved in naproxen demethylation, although 2C9 appears to be the predominant form. Received: 16 September 1996 / Accepted in revised form: 20 December 1996  相似文献   

2.
Objective: The present study was conducted to identify in vitro the cytochrome P450(CYP) isoform involved in the metabolic conversion of reduced haloperidol to haloperidol using microsomes derived from human AHH-1 TK +/− cells expressing human cytochrome P450s. The inhibitory and/or stimulatory effects of reduced haloperidol or haloperidol on CYP2D6-catalyzed carteolol 8-hydroxylase activity were also investigated. Results: The CYP isoform involved in the oxidation of reduced haloperidol to haloperidol was CYP3A4. CYP1A1, 1A2, 2A6, 2B6, 2C8, 2C9, 2C19, 2D6, and 2E1 were not involved in the oxidation. The kM value for the CYP3A4 expressed in the cells was 69.7 μmol · l−1, and the Vmax was 4.87 pmol · min−1 · pmol−1 P450. Troleandomycin, a relatively selective probe for CYP3A enzymes, inhibited the CYP3A4-mediated oxidation of reduced haloperidol in a dose-dependent manner. Quinidine and sparteine competitively inhibited the oxidative reaction with a ki value of 24.9 and 1390 μmol · l−1, respectively. Carteolol 8-hydroxylase activity, which is a selective reaction probe for CYP2D6 activity, was inhibited by reduced haloperidol with a ki value of 4.3 μmol · l−1. Haloperidol stimulated the CYP2D6-mediated carteolol 8-hydroxylase activity with an optimum concentration of 1 μmol · l−1, whereas higher concentrations of the compound (>10 μmol · l−1) inhibited the hydroxylase activity. Conclusion: It was concluded that CYP3A4, not CYP2D6, is the principal isoform of cytochrome P450 involved in the metabolic conversion of reduced haloperidol to haloperidol. It was further found that reduced haloperidol is a substrate of CYP3A4 and an inhibitor of CYP2D6, and that haloperidol has both stimulatory and inhibitory effects on CYP2D6 activity. Received: 10 April 1997 / Accepted in revised form: 16 December 1997  相似文献   

3.
Objective: Pain intensity and the plasma concentrations of metoprolol and its major metabolite α-hydroxymetoprolol as well as noradrenaline (NA), adrenaline (A) and neuropeptide Y (NPY) were determined in patients with pain due to definite or suspected acute myocardial infarction (AMI) after graded metoprolol infusion. Pain intensity and metoprolol kinetics were assessed over 8 h. Methods: Twenty-seven patients of either sex, aged 48–84 years with ongoing chest pain upon arrival to the Coronary Care Unit (CCU) were subdivided into two groups: (1) patients with ECG signs of threatening transmural myocardial damage (n=15); and (2) patients without such ECG signs (n=12). Pain intensity was assessed by a numerical rating scale (NRS) and venous blood was obtained for determination of plasma catecholamine and NPY concentrations. A continuous infusion of metoprolol (3 mg · min−1 i.v) was started and serial blood samples for plasma catecholamines, NPY as well as metoprolol and its major metabolite α-hydroxymetoprolol were obtained from the contralateral arm. Results: Initial pain intensity was 5.9 (arbitrary units) and 5.4 in the groups with and without signs of transmural myocardial damage, respectively. One third of the patients with ST changes reported full pain relief (NRS=0) within 70 min after starting metoprolol infusion (accumulated dose, 15–180 mg). Among the patients without ST changes upon arrival, full pain relief was obtained in 70% (accumulated dose, 30–120 mg). There was a dose-dependent relation between accumulated metoprolol dose and pain relief. The diagnosis of acute myocardial infarction (AMI) was confirmed in all 15 patients with ECG signs on arrival of transmural myocardial damage. The mean metoprolol dose in this group was 91(12) mg. The mean metoprolol dose in the 12 patients without ST changes was 64(8) mg. In all, seven of these patients developed definite AMI. The terminal half-life of unchanged metoprolol ranged from 2.5 to 8.5 h in group 1 and from 2.2 to 5.2 h in group 2. In group 1, metoprolol half-life was 4.5 h and total plasma clearance (CL) 54.1 l · h−1. In group 2, the metoprolol half-life was 3.7 h and total plasma clearance 75.4 l · h−1. There was a significant difference in clearance between the groups. After the intravenous metoprolol infusion, α-hydroxymetoprolol concentrations increased gradually. In groups 1 and 2, maximal concentrations in plasma (Cmax) were 143 and 135 nmol · l−1 for α-hydroxymetoprolol and 2830 and 1653 nmol · l−1 for metoprolol, respectively. Plasma NA or NPY did not differ between the groups. In contrast, plasma A was significantly higher during the initial 90 min of observation in patients with ECG signs of transmural myocardial damage. Conclusion: High-dose intravenous metoprolol was well tolerated in patients with suspected AMI. There was a more rapid and almost complete pain relief in patients without signs of transmural ischaemia compared with the patients with ECG signs of transmural AMI at arrival. In the later group of patients, plasma clearance of metoprolol was significantly reduced. Received: 23 August 1996 / Accepted in revised form: 6 March 1997  相似文献   

4.
Objectives: Thiopurine S-methyltransferase (TPMT) activity, when measured in red blood cells (RBC) with a recently published TPMT activity assay using 6‐thioguanine (6-TG) as substrate, could not be reproduced in another laboratory. We investigated factors which could influence the results of the TPMT activity measurement. Methods: We tested twelve 6-TG and four 6-mercaptopurine (6-MP) compounds from different suppliers as substrates and determined the enzyme kinetic parameters Km and Vmax . Furthermore, we studied the influence of different 6-TG compounds on the affinity of the methyl donor S-adenosyl-l-methionine (SAM) to the TPMT enzyme. Results: All 6-TG products were of equal purity (declared >98% by the supplier); this was ascertained by HPLC. However, the rate of methylation obtained following incubation with 6-TG from different suppliers ranged from 10% to 100% when incubated with the same RBC lysate. The lowest apparent Km value for a 6-TG was 22.3 μmol · l−1, while the product with the highest methylation rate showed a Km of 156 μmol · l−1. From these results we assume that there is a contaminant in some 6-TG products, which acts as a strong inhibitor of TPMT activity. Compounds possibly used for the synthesis of 6-TG (guanine, pyridine, 6-chloroguanine) did not affect the methylation rate. Thioxanthine, which is known to be a strong inhibitor of TPMT when added to the assay system to give a 2% contamination, reduced TPMT activity from 100% to 72%. Using 6-MP from different suppliers as substrate resulted in Km values ranging from 110 to 162 μmol · l−1 and Vmax values ranging from 54 to 68 nmol 6‐MMP · g−1Hb · h−1. The Km value for the methyl donor SAM was similar to and independent from the thiopurine substrates tested (range 4.9–11 μmol · l−1 SAM). In contrast to other investigators, we found non-enzymatic S-methylation, which was negligible under our assay conditions (3% with 128 μmol · l−1 SAM), but could become relevant in experiments using higher SAM concentrations. Conclusions: TPMT enzyme activity determined with 6‐TG as substrate may be strongly inhibited by a contaminant in some of the 6-TG lots distributed. Received: 28 June 1998 / Accepted in revised form: 18 January 1999  相似文献   

5.
Objective: The pharmacokinetics of cloxacillin was investigated in 14 men and 24 women undergoing cemented hip (n = 19; age range 56–90) or knee replacement surgery (n = 19; age range 51–84) for osteoarthritis. Cloxacillin 1 g was given intravenously as a bolus dose at the induction of anesthesia, and plasma samples and urine were collected for 6 h. Drug levels were determined using HPLC. Results: Preoperative serum creatinine levels were 84 μmol · l−1 in hip patients and 72 μmol · l−1 in knee patients. The calculated values for creatinine clearance were 63 and 85 ml · min−1 · 1.73 m−2, respectively. Total clearance of cloxacillin was 134 ml · min−1 · 1.73 m−2 in eighteen evaluated patients undergoing hip replacement, and 162 ml · min−1 · 1.73 m−2 in eighteen patients undergoing knee surgery. Renal clearance was 72 and 79 ml · min−1 · 1.73 m−2, respectively. Non-renal clearance was 57 ml · min−1 · 1.73 m−2 in hip patients and 77 ml · min−1 · 1.73 m−2 in knee patients. Renal clearance of cloxacillin correlated with the estimated creatinine clearance (r = 0.652). Although women received higher doses than men (median 2.02 vs 2.32 mmol · 1.73 m−2), there were no sex differences in clearance corrected for body surface area. Conclusion: Total clearance of cloxacillin was lower in patients undergoing hip replacement than in patients undergoing replacement of the knee, but there was no difference between men and women. Received: 7 May 1996 / Accepted in revised form: 15 October 1996  相似文献   

6.
Objective: To investigate the in vitro potential of selective serotonin reuptake inhibitors (SSRIs) to inhibit two CYP2C9-catalysed reactions, tolbutamide 4-methylhydroxylation and (S)-warfarin 7-hydroxylation. Methods: The formation of 4-hydroxytolbutamide from tolbutamide and that of 7-hydroxywarfarin from (S)-warfarin as a function of different concentrations of SSRIs and some of their metabolites was studied in microsomes from three human livers. Results: Both tolbutamide 4-methylhydroxylation and (S)-warfarin 7-hydroxylation followed one enzyme Michaelis-Menten kinetics. Kinetic analysis of 4-hydroxytolbutamide formation yielded a mean apparent Michaelis-Menten constant (Km) of 133 μM and a mean apparent maximal velocity (Vmax) of 248 pmol · min−1 · mg−1; formation of 7-hydroxywarfarin yielded a mean Km of 3.7 μM and a mean Vmax of 10.5 pmol · min−1 ·  mg−1. Amongst the SSRIs and some of their metabolites tested, only fluvoxamine markedly inhibited both reactions. The average computed inhibition constant (Ki) values and ranges of fluvoxamine when tolbutamide and (S)-warfarin were used as substrate, were 13.3 (6.4–17.3) μM and 13.0 (8.4–18.7) μM, respectively. The average Ki value of fluoxetine for (S)-warfarin 7-hydroxylation was 87.0 (57.0–125) μM. Conclusion: Amongst the SSRIs tested, fluvoxamine was shown to be the most potent inhibitor of both tolbutamide 4-methylhydroxylation and (S)-warfarin 7-hydroxylation. Fluoxetine, norfluoxetine, paroxetine, sertraline, desmethylsertraline, citalopram, desmethylcitalopram had little or no effect on CYP2C9 activity in vitro. This is consistent with in vivo data indicating that amongst the SSRIs, fluvoxamine has the greatest potential for inhibiting CYP2C9-mediated drug metabolism. Received: 20 July 1998 / Accepted in revised form: 6 October 1998  相似文献   

7.
Objective: Recently the role of peripheral vasoconstriction in the aetiology of insulin resistance has been proposed. Celiprolol is a β1-selective adrenoceptor antagonist with partial agonist activity at the β2-receptor as well as vasodilator properties. The acute effects of celiprolol on skeletal muscle blood flow and insulin sensitivity were measured in this study. Methods: Celiprolol (2 times 0.5 mg · kg−1) or saline was given intravenously to five healthy males in random order. Muscle blood flow was measured in femoral regions using [15O]-labelled water and positron emission tomography (PET) during euglycaemic hyperinsulinaemia (serum insulin ˜65 mU · l−1) after an overnight fast. Thereafter, skeletal and heart muscle glucose uptake were determined using [18F]-2-deoxy-d-glucose. Results: Celiprolol increased muscle blood flow by 74%, from 3.4 to 5.9 ml · min−1 · 100 g−1 muscle in the basal state. It decreased peripheral resistance by 40%, from␣32.0 to 19.2 mmHg · ml−1 · min−1 · 100 g−1. Celiprolol significantly decreased diastolic blood pressure from 82 to 73 mmHg and increased heart rate from 61 to 68 beats · min−1, which suggests sympathetic activation. Insulin-stimulated glucose uptake was reduced by 46% in the whole body, from 39 to 21 μmol · kg−1 · min−1 and by 59% in the femoral muscles, from 99 to 41 μmol · kg−1 · min−1, with celiprolol as compared to saline. The effect on heart glucose uptake did not statistically differ between the treatments. Conclusion: Celiprolol given intravenously increased muscle blood flow and decreased peripheral resistance at rest. It also acutely increased heart rate probably via sympathetic activation, and decreased insulin sensitivity in the muscles of healthy male volunteers. The enhanced muscle perfusion when celiprolol is given intravenously does not explain the improved insulin sensitivity seen in the long-term oral use in dyslipidaemic hypertensive patients. Received: 19 September 1996 / Accepted in revised form: 13 November 1996  相似文献   

8.
Objective: To compare the venodilator potencies of the phosphodiesterase (PDE) III inhibitors amrinone and enoximone with the unspecific PDE inhibitors theophylline and pentoxifylline in human hand veins in vivo. Methods: Eighteen healthy nonsmokers (16 men and two women) were studied using the dorsal hand vein technique. After preconstriction with the selective α1-adrenergic-receptor agonist phenylephrine dose–response curves were constructed for amrinone (1–270 μg · min−1), enoximone (1–270 μg · min−1), theophylline (5–1500 μg · min−1) and pentoxifylline (2–877 μg · min−1) in a random order on separate occasions. Due to limitation in the maximum dose infused in order to avoid systemic effects, full dose–response curves could not be constructed for pentoxifylline. In this case, the individual dose of pentoxifylline and theophylline producing 50% venodilation were compared. Results: All PDE inhibitors induced dose-dependent venodilation. The value of maximum venodilation was the same for amrinone, enoximone and theophylline. The infusion rate needed to induce 50% of maximum venodilation (ED50) was not significantly different for amrinone (geometric mean, 8.8 μg · min−1) and enoximone (14.2 μg · min−1), whereas the ED50 of theophylline (84.0 μg · min−1) was significantly higher than either amrinone or enoximone. The dose necessary to dilate the vein to 50% the maximum dilation (as determined during sodium chloride infusion) was significantly higher for pentoxifylline than for theophylline (409 vs 71 μg · min−1). Conclusions: These findings demonstrate that enoximone and amrinone have similar venodilatory potency which is six times higher than that of theophylline. The least potent vasodilator in this study was pentoxifylline. Received: 16 September 1997 / Accepted in revised form: 4 December 1997  相似文献   

9.
Objectives: To determine the effect of an exacerbation of respiratory symptoms in cystic fibrosis (CF) on the activities of plasma benzoylcholinesterase and butyrylcholinesterase. Methods: Twenty-nine patients with CF in a respiratory exacerbation and 27 healthy volunteers matched for age and sex were recruited. Blood was obtained from the patients when commencing antibiotic treatment and 14 days later on completion of treatment. One blood sample was taken from the healthy volunteers. The activities of benzoylcholinesterase and butyrylcholinesterase were determined by spectrophotometric assay. The circulating inflammatory markers, C-reactive protein and neutrophil elastase-α1antiproteinase complex were also measured. Results: Benzoylcholinesterase activity was significantly (P = 0.001) lower in patients at the start of a respiratory exacerbation, compared with healthy controls [mean (SD): 917 (274) versus 1191(298) nmol · ml−1 · min−1]. Benzoylcholinesterase activity increased significantly in patients to 1013 (237) nmol · ml−1 · min−1, following a course of antibiotic treatment (P = 0.006). Butyrylcholinesterase activity was also lower (P = 0.001) in patients at the start of a respiratory exacerbation, compared with healthy controls [5.54 (1.64) versus 7.01 (1.79) μmol · ml−1 · min−1], and increased significantly in the patients to 6.31 (1.58) μmol · ml−1 · min−1 following treatment (P = 0.006). Conclusion: We demonstrated significant suppression of plasma esterase activities during an exacerbation of respiratory symptoms in CF, which was only partially reversed after antibiotic treatment. Further studies are needed to examine other pathways of drug metabolism in this group of chronically infected patients. Received: 8 June 1998 / Accepted in revised form: 18 September 1998  相似文献   

10.
Objective: Mibefradil is a novel calcium channel antagonist that selectively blocks T-channels. It acts to reduce hypertension, is cardioprotective and reduces ischemic episodes. Oxidative modification of low-density lipoproteins (LDL) is well known to contribute to coronary atherosclerosis and we therefore investigated to see whether mibefradil had antioxidative action on LDL. Methods: Human LDL were isolated by ultracentrifugation. In vitro oxidation of LDL (0.1 μmol · l−1 protein) in the presence of various concentrations of mibefradil was initiated by 3.2 μmol · l−1 copper ions. The kinetics of formation of conjugated dienes was followed photometrically. Malondialdehyde and lipoperoxides were determined at maximum oxidation. LDL (0.3 μmol · l−1) were also pre-incubated with mibefradil (120 μmol · l−1). Excessive mibefradil was separated by column technique. The resultant LDL were oxidized using copper ions or (AAPH) 2,2′-azobis(2-amidinopropane) hydrochloride. Results: The presence of mibefradil in the concentration range from 10 to 200 μmol · l−1 had dose-dependent effects. These were protection of LDL against oxidation measured as prolongation of the lagtime up to 250%, and reduction in the formation of malondialdehyde down to 65% and of lipoperoxides to 20%. Pre-incubation of LDL with mibefradil prolonged the lagtime of Cu-mediated oxidation up to 132% and of AAPH-mediated oxidation up to 138%. Conclusion: In addition to the T-channel blocking and antiproliferative effects, our results provide arguments for a protective role of mibefradil (10–200 μmol · l−1) on LDL against in vitro oxidation. This was shown with three independent parameters (lagtime, malondialdehyde and lipoperoxides) and in different oxidation models. Received: 11 March 1998 / Accepted in revised form: 1 July 1998  相似文献   

11.
Objective: To assess the cardiovascular and metabolic responses to increasing doses of formoterol administered from a dry powder inhaler. Methods: Twenty patients with mild to moderate asthma were given 12, 24, 48 and 96 μg of formoterol or a matched placebo on separate days. The doses were administered using a randomised, cross-over, double-blind design. The effects on heart rate, blood pressure, electromechanical systole (QS2I), the electrocardiographic QTc interval, plasma potassium (K); blood glucose and FEV1 were assessed prior to, and for 9 h following each dose. Results: There was no difference between the maximum effects of formoterol 12 μg and placebo; the 24 μg dose significantly decreased plasma K (−0.2 mmol · l−1) and increased blood glucose (1.8 mmol · l−1) compared to placebo. The two highest doses affected most of the variables with the 96 μg dose being significantly different from placebo for all indices, heart rate (9 beats · min−1), systol BP (4 mmHg), diastolic BP (−3 mmHg), QS2I (−11 ms), QTc (17 ms), plasma K (−0.5 mmol · l−1) and blood glucose (2.6 mmol · l−1). All doses of formoterol increased FEV1. Conclusion: Although there were dose-dependent effects on the extrapulmonary measurements, only the effects at the highest dose may be of clinical significance. Received: 17 June 1997 / Accepted in revised form: 12 November 1997  相似文献   

12.
Objective: To determine the effect of the coadministration of the multidrug resistance (MDR) modulators cyclosporin A (CyA) alone or plus dexverapamil (D-Ver) on idarubicin (IDA) pharmacokinetics in patients with acute leukemia. Methods: Pharmacokinetic studies were performed in 27 patients with a diagnosis of acute myelogenous leukemia (AML), who were being treated with a combination chemotherapy regimen including idarubicin and cytarabine for the induction of a first remission (n = 14), or of a second remission (n = 7), or for remission consolidation (n = 6). Of these 27 patients, nine were coadministered CyA and seven were coadministered CyA plus D-Ver as MDR modulators. Blood was sampled at appropriate intervals after each of the three IDA daily administrations. IDA and idarubicinol (IDAOL) were assayed by HPLC. Pharmacokinetic evaluations were performed by means of a two-compartment open model with zero-order absorption and first-order elimination using the WinNonlin pharmacokinetic software package. Results: CyA markedly increased the area under the concentration time-curve (AUC) of both IDA [558.26 (197.25) μg · h · l−1 vs 315.44 (158.28) μg · h · l−1; P < 0.01] and IDAOL [2896.60 (736.38) μg · h · l−1 vs 1028.49 (603.95) μg · h · l−1; P < 0.001] when coadministered as a single modulator, due to a lower total body clearance (CL) [83.51 (52.44) l · h−1 · m−2 vs 139.65 (69.45) l · h−1 · m−2; NS]. When patients received two MDR modulators simultaneously (D-Ver plus CyA), IDA exposure was essentially the same as in those of the no inhibitor group [331.29 (95.49) μg · h · l−1 vs 315.44 (158.28) μg · h · l−1; NS], whereas the IDAOL total body exposure was greater than in the no inhibitor group [2030.32 (401.11) μg · h · l−1 vs 1028.49 (603.95) μg · h · l−1; P < 0.01], even if less than in patients receiving CyA as a single MDR modulator (IDA + CyA group) [AUC 2030.32 (401.11) μg · h · l−1 vs 2896.60 (736.38) μg · h · l−1; P < 0.05], suggesting an antagonistic effect against those of CyA on IDA and IDAOL elimination and/or an unpredictable redistribution. The main pharmacokinetic parameters of IDA, such as CL and volume of distribution at steady state (Vdss), were remarkably affected by the coadministration of CyA or CyA plus D-Ver, but no statistically significant difference was noted because of IDA pharmacokinetic interpatient variation. Conclusion: The results show that CyA alone at a dose of 10 mg · kg−1 daily significantly increased systemic body exposure to both IDA and IDAOL in acute leukemia, and suggest that these pharmacokinetic effects were at least partially decreased when D-Ver was coadministered with CyA. Our findings raise important questions concerning the need for a dosage adjustment of IDA when MDR modulators are coadministered. Received: 2 June 1998 / Accepted in revised form: 3 December 1998  相似文献   

13.
Objective: The purpose of this study was to investigate whether the disposition of fluvoxamine is associated with the CYP2D6 and CYP2C19 phenotype polymorphisms. Methods: The serum concentration of fluvoxamine was followed for 48 h after oral administration of a single dose of 50 mg fluvoxamine to five poor metabolizers of the CYP2D6 test drug dextromethorphan, five poor metabolizers of the CYP2C19 test drug mephenytoin, and five extensive metabolizers of both test drugs. Results: Poor metabolizers of dextromethorphan had significantly higher areas under the serum concentration-time curve than extensive metabolizers of dextromethorphan (mean 1.31 vs 1.00 μmol · h · l−1). There were no differences between poor and extensive metabolizers of mephenytoin (mean, 1.00 vs 1.15 μmol · h · l−1). Conclusion: The results are consistent with a possible minor to moderate role of CYP2D6, but not CYP2C19, in fluvoxamine metabolism. Received: 25 April 1996 / Accepted in revised form: 12 November 1996  相似文献   

14.
Objective: The aim of this investigation was to study the variation of catechol-O-methyltransferase (COMT) activity in the human liver, duodenal mucosa and renal cortex, and to investigate the inhibition of COMT by entacapone and tolcapone. This study included 87 samples of human liver, 94 samples of the duodenum and 72 samples of the renal cortex. Results: The activity of COMT was measured with 3,4-dihydroxybenzoic acid (242 μmol · l−1), the methyl acceptor substrate, and adenosyl-l-methionine (44 μmol · l−1), the methyl donor substrate. The hepatic activity of COMT activity was significantly higher in men than in women, whereas it was not sex-dependent in the duodenum or renal cortex. The activity of COMT varied 4.4-fold in the liver of men, 2.6-fold in the duodenum and 5.3-fold in the renal cortex. The median estimates of COMT activity were 577, 499, 103 and 159 pmol · min−1 · mg−1 in the liver of men and women, in the duodenum and in the renal cortex, respectively. Conclusion: Entacapone and tolcapone were powerful inhibitors of COMT and their IC50 estimates were 151 and 773 nM (P = 0.008), respectively, in the liver; consistent results were obtained with the other tissues.  相似文献   

15.
Objective: The pharmacokinetics of 6-mercaptopurine, including cerebrospinal-fluid (CSF) distribution, and the erythrocyte 6-thioguanine nucleotide concentrations were determined in children randomised to receive intravenous mercaptopurine for acute lymphoblastic leukaemia (ALL), according to the EORTC protocol ALL n°58881. Results: After 1 month of oral treatment at a dose of 50 mg · m−2 · day−1, the pharmacokinetic parameters were determined after the first i.v. administration of 1 g · m−2 (bolus dose of 0.2 g · m−2 followed by an 8-h infusion of 0.8 g · m−2) in 11 patients: systemic clearance was 23.02 l · h−1, volume of distribution was 0.75 l · kg−1, and elimination half-life was 1.64 h. The erythrocyte thioguanine concentrations were measured in the same 11 patients and increased significantly between the beginning and the end of infusion (10 pmol × 108 packed RBC) or within 24 h of infusion (223 pmol × 108 packed RBC). The CSF concentration was 3.78 μmol · l−1, 1–6 h after the beginning of infusion (n=28) and the CSF to plasma ratio was 0.15 (n=16). In patients receiving the oral dose of 50–165 mg · m−2 · day−1 of 6-mercaptopurine, CSF concentrations were below 0.18 μmol · l−1, 1–24 h after drug intake (n=67), and the CSF to plasma ratio was not calculated. Conclusion: Following the i.v. administration of 6-mercaptopurine, we observed high CSF concentrations of 6-mercaptopurine and an acute increase of erythrocyte thioguanine nucleotide concentrations. The clinical trial (EORTC protocol ALL n°5881), comparing the oral and i.v. administrations of mercaptopurine, will demonstrate if the i.v. administration reduces the incidence of CNS relapses. Received: 15 August 1996 / Accepted in revised form: 8 April 1997  相似文献   

16.
Objective: The aim of this study was to evaluate the steady-state effects of once-daily inhaled fluticasone propionate (FP) and budesonide (BUD) on adrenocortical activity in asthmatic patients. Methods: Ten asthmatic patients with a mean age of 31.2 years, a mean forced expiratory volume in 1 s (FEV1) of 91% predicted and a forced mid-expiratory flow (FEF25–75) of 62.3% predicted were studied in a single-blind randomised crossover design comparing placebo (PL), FP (375 μg per day and 750 μg per day) and BUD (400 μg per day and 800 μg per day) all given once daily for 4 days at each dose via a pressurised metered dose inhaler (pMDI) at 0800 hours. After 4 days of treatment, plasma cortisol was measured at 0800 hours (24 h after the last dose) and a 10-h overnight urine collection was taken, 14 h after the last dose (2200–0800 hours) for analysis of cortisol and creatinine excretion. Results: Plasma cortisol levels (nmol · l−1, as geometric mean) at 0800 hours demonstrated a significant difference between the highest doses of FP and BUD (424.1 vs 510.3 nmol · l−1, respectively) but not between the low doses (506.8 vs 514.9 nmol · l−1; PL 532.2 nmol · l−1). For the highest dose FP (750 μg) this equated to 20% suppression of 0800 hours plasma cortisol. Likewise, for overnight urinary cortisol output (nmol · 10 h−1, as geometric mean), there was a significant difference at the high doses of FP and BUD (25.5 vs 38.2 nmol · 10 h−1), but not at the low doses 31.3 vs 34.8 nmol · 10 h−1; PL 32.0 nmol · 10 h−1. For the overnight urinary cortisol/creatinine ratio (nmol · mmol−1, as geometric mean) there was a similar trend; 4.5 vs 6.1 nmol · mmol−1 for high dose and 5.6 vs 6.3 nmol · mmol−1 for low dose; PL 5.9 nmol · mmol−1. Conclusion: Repeated doses of FP 750 μg once daily caused greater adrenal suppression than BUD 800 μg once daily, when comparing effects on plasma cortisol levels at 0800 hours, 24 h after the last dose, as well as effects on overnight urinary cortisol output. Neither FP 375 μg once daily nor BUD 400 μg once daily produced detectable adrenal suppression. Received: 29 April 1997 / Accepted in revised form: 5 July 1997  相似文献   

17.
Objective: In a randomized cross-over study, we assessed pharmacokinetics and intracellular concentrations in polymorphonuclear leukocytes (PMN) and saliva of erythromycin and erythromycylamine, the active metabolite of dirithromycin. Methods: Ten healthy volunteers received 1 g erythromycin b.i.d. or 500 mg dirithromycin qd for 5 days (wash out period, 35 days). Concentrations of erythromycin and erythromycylamine were measured in serum, urine, saliva, and granulocytes by bioassay and high-performance liquid chromatography (HPLC) on days 1, 3, and 5 of each study period, respectively. Results: While maximal serum concentrations (Cmax) and the area under the data (AUDtot) of erythromycin were significantly higher (Cmax 1.44 mg · l−1, AUDtot 5.66 mg · h · l−1) than those of erythromycylamine (Cmax 0.29 mg · l−1, AUDtot 1.96 mg · h · l−1), erythromycylamine had a significantly higher mean residence time (21 h) than erythromycin (5.5 h). Erythromycylamine accumulated significantly more in PMN than erythromycin;␣the accumulation factor of erythromycylamine was 100 with a maximal intracellular concentration of 13.4 mg · l−1, whereas the maximal accumulation factor of erythromycin was 4 with a maximal intracellular concentration of 6.1 mg · l−1. There were no significant differences in maximal saliva concentrations (erythromycin 0.35 mg · l−1, erythromycylamine 0.31 mg · l−1). Received: 16 September 1996 / Accepted in revised form: 12 February 1997  相似文献   

18.
Objective: The pharmacokinetics of N G-nitro-l-arginine methyl ester (l-NAME), an inhibitor of nitric oxide (NO) synthesis, was investigated in patients with septic shock. Methods: Blood was sampled at intervals before, during and after 12-h infusion of l-NAME 1 mg · kg−1 · h−1 in nine septic shock patients for determination of plasma concentrations by high-performance liquid chromatography (HPLC). In three patients the renal clearance of the drug was determined. Results: Incubation of l-NAME with plasma and blood in vitro revealed hydrolysis to N G-nitro-l-arginine (l-NOARG), the active inhibitor of NO synthesis. l-NOARG did not undergo further degradation. Continuous intravenous infusion of 1 mg · kg−1 · h−1 of l-NAME for 12 h in patients with septic shock increased blood pressure and resulted in increasing plasma concentrations of l-NOARG (Cmax 6.2 μg · ml−1 at 12 h) whereas l-NAME concentrations reached a plateau within 1.5 h (Cmax 1.0 μg · ml−1). After the infusion was stopped l-NAME disappeared from the plasma rapidly (half-life 19.2 min) whereas l-NOARG concentration declined slowly (half-life 22.9 h). The calculated volume of distribution for l-NAME was 0.45 l · kg−1 body weight and 1.96 l · kg−1 for l-NOARG. The renal clearance for l-NOARG was 3.5% of total body clearance for l-NOARG, whereas l-NAME could not be detected in urine. Conclusion: We conclude that vasoconstriction with l-NAME in septic patients may result from hydrolysis to l-NOARG, the active inhibitor of NO synthesis. The long plasma half-life and large volume of distribution for l-NOARG suggests extensive distribution to extravascular tissues. Since renal excretion is minimal, elimination of the metabolite l-NOARG follows other pathways. Received: 13 March 1998 / Accepted in revised form: 30 June 1998  相似文献   

19.
Objectives: The aim of this study was to compare the rate of absorption between ordinary paracetamol tablets and effervescent paracetamol tablets. Methods: Twenty healthy volunteers participated in an open randomised crossover study and were given a 1000-mg dose of either ordinary paracetamol tablets (2 × 500 mg Panodil tablets, SmithKline Beecham) or effervescent paracetamol tablets (2 × 500 mg Pinex Brusetablett, Alpharma AS) with a 3-week washout period in between. Blood samples were collected for 3 h. Maximum serum concentration (Cmax) and the time to maximum serum concentration (tmax) were recorded and the area under the concentration versus time curve (AUC) was calculated. Results: The mean tmax was significantly shorter when paracetamol effervescent tablets were taken (27 min) rather than ordinary paracetamol tablets (45 min) (P=0.004). There was no significant difference between the mean Cmax of 143 μmol/l with effervescent tablets and that of 131 μmol/l with ordinary tablets. The mean AUC0–3 h was significantly higher with paracetamol effervescent tablets (223.8 μmol · h · l−1) than with ordinary tablets (198.2 μmol · h · l−1; P=0.003). After 15 min, 17 (85%) subjects in the effervescent group had a serum concentration of 70 μmol/l (lower therapeutic serum concentration) or higher relative to only 2 (10%) subjects in the ordinary tablet group (P=0.001). Conclusion: Paracetamol effervescent tablets are absorbed significantly faster than ordinary paracetamol. Thus, effervescent tablets might offer significantly faster pain relief when paracetamol is used. Received: 4 October 1999 ;/ Accepted in revised form: 15 February 2000  相似文献   

20.
Objective: Iloprost is a potent PGI2 mimetic, which has been shown to be therapeutically effective in several vascular disorders. Due to its rapid clearance from the central compartment, iloprost is administered mainly by i.v. infusion, which limits its use to hospitalized patients. In order to improve pharmacotherapeutic use of this PGI2 mimetic, an oral extended-release (ER) dosage form has been developed, which should mimic plasma level profiles as observed after i.v. infusion and serve as a therapeutic equivalent. Methods: This trial was performed to investigate the tolerability and pharmacokinetics of iloprost administered perorally, compared with i.v. infusion, in 12 patients suffering from thromboangiitis obliterans (TAO). A dose titration was carried out for 1 week with i.v. iloprost, followed by a p.o. titration and treatment phase of 3 weeks' duration. Pharmacokinetics was investigated at the individually tolerated dose levels; i.e., on days 5–7 (i.v. infusion at 2, 2.5 and 3 ng · kg−1 · min−1), and twice during p.o. treatment after b.i.d. administration of 50, 100, 150, 200 or 300 μg. Results: Individual tolerability of iloprost varied: 7 patients out of 12 tolerated the maximum i.v. dose of 3 ng · kg−1 · min−1; six tolerated the maximum oral dose of 600 μg. No patients withdrew from the study due to adverse events. Flush and headache were the most common adverse events and seemed to be related to the study drug. After i.v. infusion of iloprost, dose-normalized (3 ng · kg−1 · min−1), steady-state plasma levels were 260 pg · ml−1. Terminal half-life was 0.57 h. Total clearance ranged from 8 to 17 ml · min−1 · kg−1. Peroral administration of the ER formulation resulted in dose-dependent Cmax and AUC values. AUC values of the first and second daily dose interval, i.e., 0–5 h and 5–11 h after first dosing, were almost identical. Absolute bioavailability was 24%, with the exception of two patients who tolerated only 50 μg b.i.d. and exhibited a bioavailability of approx. 60%. The AUC values observed in weeks 2 and 4 were identical, demonstrating low day-to-day variability of iloprost plasma level profiles in TAO patients. Conclusion: Based upon pharmacokinetic data, the ER formulation provides an equivalent to the i.v. infusion of iloprost and broadens the range of therapy to non-hospitalized patients. The availability of capsules with 50 and 100 μg iloprost enables individual dose titration and pharmacotherapy. Beneficial effects, as observed with i.v. iloprost in TAO patients, should therefore be achievable by peroral pharmacotherapy using the new ER formulation. Received: 18 July 1996 / Accepted in revised form: 2 April 1997  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