Original Research Open Access
Copyright ©The Author(s) 2001. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Gastroenterol. Dec 15, 2001; 7(6): 830-835
Published online Dec 15, 2001. doi: 10.3748/wjg.v7.i6.830
Chiral metabolism of propafenone in rat hepatic microsomes treated with two inducers
Quan Zhou, Tong-Wei Yao, Su Zeng, College of Pharmaceutical Sciences, Zhejiang University, Hangzhou 310031, Zhejiang Province, China
Quan Zhou, Second Hospital of Medical School, Zhejiang University, Hangzhou 310031, Zhejiang Province, China
Author contributions: All authors contributed equally to the work.
Supported by the National Natural Science Foundation of China (NO.39370805, NO.39770868) and Zhejiang Natural Science Foundation (№ RC97016)of Zhejiang Province
Correspondence to: Prof. Su Zeng, College of Pharmaceutical Sciences, Zhejiang University, Hangzhou 310031, China. zengsu@zjuem.zju.edu.cn
Telephone: +86-571-87217060, Fax: +86-571-87217086.
Received: June 2, 2001
Revised: July 19, 2001
Accepted: August 5, 2001
Published online: December 15, 2001

Abstract

AIM: To study the influence of inducers of drug metabolism enzyme, β-naphthoflavone (BNF) and dexamethasone (DEX), on the stereoselective metabolism of propafenone in the rat hepatic microsomes.

METHODS: Phase I metabolism of propafenone was studied using the microsomes induced by BNF and DEX and the non-induced microsome was used as the control. The enzymatic kinetics parameters of propafenone enantiomers were calculated by regress analysis of Eadie-Hofstee Plots. Propafenone enantiomer concentrations were assayed by a chiral HPLC.

RESULTS: The metabolite of propafenone, N-desalkylpropafenone, was found after incubation of propafenone with the rat hepatic microsomes induced by BNF and DEX. In these two groups, the stereoselectivity favoring R (-) isomer was observed in metabolism at low substrate concentrations of racemic propafenone, but lost the stereoselectivity at high substrate concentrations. However, in control group, no stereoselectivity was observed. The enzyme kinetic parameters were: ① Km. Control group: R (-) 83 ± 6, S (+) 94 ± 7; BNF group: R (-) 105 ± 6, S (+) 128 ± 14; DEX group: R (-) 86 ± 11, S (+) 118 ± 16; ② υmax. Control group: R (-) 0.75 ± 0.16, S (+) 0.72 ± 0.07; BNF group: R (-)1.04 ± 0.15, S (+)1.0 7 ± 14; DEX group: R (-) 0.93 ± 0.06, S (+) 1.04 ± 0.09; ③ Clint. Control group: R (-) 8.9 ± 1.1, S (+) 7.6 ± 0.7; BNF group: R (-)9.9 ± 0.9, S (+)8.3 ± 0.7; DEX group: R (-) 10.9 ± 0.8, S (+) 8.9 ± 0.9. The enantiomeric differences in Km and Clint were both significant, but not in υmax, in BNF and DEX group. Whereas enantiomeric differences in three parameters were all insignificant in control group. Furthermore, Km and υ max were both significantly less than those in BNF or DEX group. In the rat liver microsome in duced by DEX, nimodipine (NDP) decreased the stereoselectivity in propafenone metabolism at low substrate concentration. The inhibition of NDP on the metabolism of propafenone was stereo selective with R (-)-isomer being impaired more than S (+)-isomer. The inhibition constant (Ki) of S (+)- and R (-)-propafenone, calculated from Dixon plots, was 15.4 and 8.6 mg•L¯¹, respectively.

CONCLUSION: CYP1A subfamily (induced by BNF) and CYP3A4 (induced by DEX) have pronounced contribution to propafenone N-desalkylation which exhibited stereose lectivity depending on substrate concentration. The molecular base for this phenomenon is the stereo selectivity in affinity of substrate to the enzyme activity centers instead of at the catalyzing sites.

Key Words: propafenone/metabolism, mitochondria, liver, rat, optical rotation



INTRODUCTION

Propafenone, is a widely used antiarrhythmic agent administered as the racemic mixture of R (-) and S (+) enantiomers. The two enantiomers are equipotent in terms of sodium channel-blocking activity, but the main side effect, i.e., β-adrenoreceptor-blocking action resides in the S (+)-isomer[1], and, therefore, information on stereoselective disposition of the racemate is of clinical relevance.

The main metabolic pathways of propafenone in vivo and in vitro involve CYP1A2 and CYP3A4 mediated N-desalkylation, CYP2D6 mediated 5-hydroxylation and UDPGT mediated glucuronidation[2-6]. N-desalkylpropafenone has the same electrophysiological potency as 5-hydroxypropafenone and propafenone, and the plasma concentrations of N-desalkylpropafenone are similar to those of 5-hydroxypropafenone during chronic administration in human, therefore, N-desalkyl propafenone contributes to the antiarrhythmic effects of propafenone, especially in patients with poor metabolizer phenotype of CYP2D6[7,8]. Although stereo selectivities in 5-hydroxylation and glucuronidation in vitro have been reported[9-11], whether N-desalkylation exhibits stereoselectivity has not been addressed. Meanwhile, rat liver microsomes pretreated by specific inducers provide sound models to study metabolism in vitro[12-16]. Considering that β-naphthoflavone (BNF) was a typical inducer of CYP1A subfamily and dexamethasone (DEX) was a typical inducer of CYP3A4[17-21], this experiment studied the stereoselective propafenone N-desalkylation in rat hepatic microsomes induced by BNF and DEX.

MATERIAL AND METHODS
Chemicals and solutions

Dexamethasone (DEX), β-naphthoflavone (BNF), 7-ethoxyresorufin (ER), triacetylole and omycin (TAO), NADPH, 2,3,4,6-tetra-O-acetyl-β-D-glucopyranosyl isothiocynate (GITC), (R, S)-propafenone, R (-) and S (+)-propafenone were supplied by Sigma Chemical Co (St. Louis, MO, US A). N-desalkylpropafenone was a generous gift from Prof. Tang YN (Xinhua Hospital, Shanghai). All other chemicals were obtained from the common commercial sources. Stock buffer (pH7.4):1 mol•L¯¹ pH7.4 Tris-HCl buffer 25 mL, 1 mol•L¯¹ KCl 75 mL and 1 mol•L¯¹ MgCl2 5 mL were mixed and diluted with water to 500 mL. NADPH solution: dissolve NADPH in ice-cold 10 g•L¯¹ NaH CO3 solution to the desired concentration of 25 mmol•L¯¹. The solution should be freshly prepared just before the incubation.

Preparation of hepatic microsomes

Sprague-Dawley rats (male, 170-210 g) were divided into three groups. One group received three daily intraperitoned injection of 80 mg•kg¯¹ BNF (dissolved in oil); the second group received three daily DEX (132 mg•kg¯¹•d¯¹, ig) and the third group was used as the non-treated control. About 24 h after the last treatment and with no food supplied for 16 h before taking the livers, the rats were sacrificed by decapitation. Liver samples were excised and perfused by the ice-cold physiological saline to remove blood and homogenized in ice-cold Tris buffer. Hepatic microsomes were prepared with the ultracentrifugation methods[22,23]. All manipulations were carried out in cold bath. Pellets were re-suspended in sucrose-Tris buffer (pH7.4) (95:5, mass to volume ratio) and immediately stored at -30 °C. Protein and cytochrome P450 contents were estimated according to the methods of Zeng et al[24] and Omura et al[25], respectively. Enzymatic activity of CYP1A was measured according to the method of Klotz et al[26], and expressed as initial velocity of O-deethylation of 7-ethoxyresorufin (activity of EROD). Enzymatic activity of CYP3A4 was determined according to the method of Wrighton et al[27], and expressed as the extent of P450-MI complex (absorbance difference per gram of protein between 456 nm and 5 10 nm) using triacetyloleandomycin as substrate. Incubation of propafenone with rat hepatic microsomes. The incubation mixture contained microsomal protein (1.6 g•L¯¹), stock buffer (pH7.4) bubbled with oxygen for 1 min and racemic propafenone as substrate. After 5 min preincubation, reaction was started by adding 10 μL NADPH solution. The final volume was 250 μL. For kine tic experiments, racemic propafenone was used at concentrations of 10, 20, 40, 80, 160, and 320 mg•L¯¹ and the incubation time was 30 min. For the time dependent experiments, the substrate co ncentration used was 10 mg•L¯¹. For inhibition experiments, nimodipine was used as inhibitors (at 0, 8, 16, 32 mg•L¯¹) and incubated simultaneously with racemic propafenone (50, 100 mg•L¯¹). After the indicated time, the reaction was terminated by adding 750 μL chlorform. The mixture was votexed for 3 min, then centrifuged at 2000 g for 10 min. The organic layer was transferred to a clean tube and evaporated to dryness under a gentle stream of air. GITC solution (in acetonitrile) and methanol containing 14 g•L¯¹ triethamine were added and the tube was capped and allowed to react for 30 min a t 35 °C. After evaporation of organic solvents, the residues were reconstutited with 100 μL me thanol, and 20 μL was injected into HPLC system.

HPLC procedure for determining propafenone enantiomer in the rat hepatic incubates

Enantiomers of propafenone were quantitated with an HPLC system with UV detection (λ = 254 nm)[28]. A 5-μm reverse phase column (Shimpack CLC- ODS 15 cm ± 4.6 mm) was used with a flow rate of 0.8 mL/min. The mobile phase was a mixture of methanol-water-glacial acetic acid (67:33:0.05).

Statistical analysis

The maximum velocity (υmax) and Michaelis-Menten constant (Km) values for propafenone enantiomer were determined by regress analysis of Edie-Hofstee plots. The ¯x ± s of three determinations of υmax and Km was calculated for each substrate and metabolic reaction. Intrinsic clearance was calculated by the ratio of υmax/Km. All statistical difference was tested by unpaired t test.

RESULTS

A baseline separation between the diastereomers of S (+) and R (-)-propafenone was achieved, with the retention time being 23 min and 28 min, respectively. The HPLC system also allowed monitoring the formation of N-desalkylpropafenone. The retention time was 8 min and 10 min for diastereomer of N-desalkylpropafenone, respectively. The amount of diastereomers of N-desalkylpropafenone were increasing while those of propafenone were decreasing during 30 min incubation with the rat hepatic microsomes induced by DEX and BNF. Typical chromatograms were showed in Figure 1. Quantitation was performed by external standardization. Calibration curves were linear at a range of 0.5 to 320 mg•L¯¹ for each enantiomer of propapfenone. The LOQ was 0.5 mg•L¯¹ (S/N = 10, n = 5) for each enantiomer. The inter-assay and intra-assay variability averaged 8.5% for both enantiomers. The method recovery averaged 77.1% for both enantiomers.

Figure 1
Figure 1 Chiral high performance liquid chromatogram of racemic propafenone in rat liver microsomal incubates after 30 min incubation. A: BNF pre-treated B: without incubation. Peaks 1, 2: Diastereomers of S(+)-propafenone and R(-)-propafenone; Peaks 3, 4: Diastereomers of metabolite (N-desalkylpropafenone)
Induction of rat hepatic metabolizingenzymes

In DEX group, the extent of P450-MI complex (an indicator of activi ty of CYP3A4) was significantly more than the control or BNF group (P < 0.001, Table 1). In BNF group, the initial velocity of deethylation of 7-ethoxyresorufin (an indicator of activity of CY P1A) was significantly more than in the control or DEX group (about 20-fold, P < 0.001). Therefore, CYP1A subfamily was successful inducted by BNF and CYP3A4 by DEX, which provided sound enzymatic sources for getting information on CYP1A and CYP3A4 mediated N-desalkylation of propafenone.

Table 1 The amount and activity of P450 in rat liver microsomes (mean ± SD, n = 3).
PretreatP450 in pro/μmol•g¯¹Extent of P450-MI complex△AActivity of EROD/μmol•min-1μg-1
Control0.95 ± 0.150.5 ± 0.20.22 ± 0.04
BNF1.42 ± 0.212.2 ± 0.43.87 ± 0.20b
Dex1.11 ± 0.1718.3 ± 3.6a0.18 ± 0.02
Impact of substrate concentration on stereoselective metabolism of propafenone

At 10 mg•L¯¹ concentration of racemic propafenone, stereoselectivity was observed in DEX and BNF group, but not in control group (Table 2). The depletion of R (-)-isomer was faster than that of S (+)-isomer. However, with the substrate concentration increasing, S/R ratios of propafenone were not altered in control group (P > 0.05), but in DEX and BNF group S/R ratios were decreasing from 1.18 to 1.00 (P < 0.01), and 1.10 to 1.00 (P < 0.01), respectively.

Table 2 Ratio of S (+)/R (-) propafenone at different concentrations in rat liver microsomal incubates (mean ± SD, n = 3).
Enantiomer/mg•L¯¹Pretreat
ControlDexBNF
51.016 ± 0.0161.177 ± 0.062ab1.104 ± 0.019ab
101.029 ± 0.0121.103 ± 0.0571.069 ± 0.015
200.995 ± 0.0161.088 ± 0.0181.053 ± 0.002
400.974 ± 0.0261.057 ± 0.0301.043 ± 0.000
800.978 ± 0.0241.019 ± 0.0171.027 ± 0.005
1600.988 ± 0.0121.003 ± 0.0191.005 ± 0.005
Concentration-time curves and ratio of S (+)/R (-) propafenone concentration

The ratio of S/R was in unity in control group from the incubation time of 0 to 30 min, whereas in DEX or BNF group, the ratio of S/R increased and was significantly different with the corresponding ratio in control group at 8 and 30 min (P < 0.01, 0.05, Table 3). Moreover, the ratio of S/R in DEX group at incubation time of 30 min was significantly higher than that in BNF.

Table 3 Ratio of S (+)/R (-) propafenone concentration in rat li ver microsomal incubates (mean ± SD, n = 3).
Groupt (incubation)/min
0382030 (min)
Control1.0001.017 ± 0.0100.997 ± 0.0161.006 ± 0.0121.016 ± 0.016
Dex1.0001.007 ± 0.0031.044 ± 0.011d1.076 ± 0.0191.170 ± 0.050abc
BNF1.0001.005 ± 0.0021.031 ± 0.012d1.068 ± 0.0231.094 ± 0.017ac
Enzymatic kinetic parameters for propafenone metabolism in hepatic microsomes

Depletion of propafenone could be described by Michaelic-Menten kinetics. Km had no statistical difference between the two enantiomers in control microsomes, whereas the enantiomeric difference in Km was significant in the microsomes induced with DEX or BNF (S > R, P < 0.05, Table 4). There was significant difference for Clint between the two enantiomers (S < R, P < 0.05, Table 4) in DEX or BNF group, but not in control group. The Km of S (+)-isomer in DEX, or S(+)- or R(-)-isomer in BNF group was significantly higher than the corresponding enantiomer in control group (P < 0.05, 0.01, Table 4). The υmax of S (+)-isomer in DEX group, or S (+)- or R (-)-isomer in BNF group, was significantly higher than the corresponding enantiomers in the control group (P < 0.05, 0.01, Table 4). Difference for Clint between the two enantiomers in DEX or BNF group and the corresponding enantiomer in control group was insignificant. Moreover, the Km of R (-)-propafenone in DEX group was significantly lower than that in BNF group (P < 0.05, Table 4).

Table 4 Enzymatic parameters in propafenone enantiomer metabolism in vitro (mean ± SD, n = 3).
PretreatEnantiomerKm/μmol•L¯¹υmax/μmol•g¯¹•min-1Clint in prot/L μmin-1•g¯¹
ControlS (+)94 ± 70.72 ± 0.077.6 ± 0.7
R (-)83 ± 60.75 ± 0.168.9 ± 1.1
DexS (+)118 ± 16ab1.04 ± 0.09c8.9 ± 0.9a
R (-)86 ± 11d0.93 ± 0.0610.9 ± 0.8b
BNFS (+)128 ± 14ac1.07 ± 0.20 b8.3 ± 0.7a
R (-)105 ± 6c1.04 ± 0.15b9.9 ± 0.9
Stereoselective inhibition of propafenone metabolism by nimodipine

Ki for S(+)- and R(-)-propafenone was 15.4 and 8.6 mg•L¯¹, respectively, which suggested that nimodipine (specific substrate of C YP3A4) inhibited metabolism of propafenone enantiomer stereoselectively (Figure 2, Figure 3). With nimodipine amount increasing, the depletion of propafenone enantiomers and the S/R ratio of the remaining amount of propafenone enantiomer were decreasing (Table 5).

Figure 3
Figure 3 Concentration-time curves for S(+)- and R(-)-propafenone metabolism in rat hepatic microsomes. A: Control; B: BNF; C: DEX.
Figure 3
Figure 3 Dixon plot for S(+)-propafenone (Left) and R(-)-propafenone (Right) wit nimodipine as inhibitor at three concentration. Ki for S (+)-and R(-)-PPF was 15.4, 8.mg•L¯¹, respectively. Each data point represents the mean of duplicate determinations.
Table 5 The stereoselective effects of nimodipine on metabolic depletion of propafenone (mean ± SD, n = 3).
GroupNimodipine/mg•L¯¹S (+)-propafenoneR (-)-propafenone/mg•L¯¹S/R
DEX02.10 ± 0.041.75 ± 0.14a1.20
DEX82.32 ± 0.262.10 ± 0.21b1.1
0DEX163.81 ± 0.11c3.62 ± 0.13c1.0
6DEX324.30 ± 0.13c4.17 ± 0.26c1.03
DISCUSSION

Due to the capabilities of highly efficient separation and sensitive determination of enantiomers in microsome incubates, chiral chromatography is extremely valuable to study stereoselectivity of racemate metabolism[29-34]. So far as we are aware, we took the lead in acquiring the information on stereochemistry of propafenone metabolism by chiral HPLC method.

Previously, we observed that the glucuronidation of propranolol in rat hepatic microsome has stereoselectivity of S (-)-propranolol, and that the induction of phenobarbital reduced this stereoselectivity[35]. The phase I metabolic stereoselectivity of propranolol was reversed by the induction of BNF and increased by the induction of phenobarbital[36]. Phenobarbital instead of BNF inducted the stereoselective difference of Clint in glucuroniodation of ofloxacin [37]. However, the induction of DEX or BNF in this study vested propafenone metabolism with stereoselectivity in rat hepatic microsomes. It is thus clear that different inducers may have different impacts on some racemate metabolism.

The enantiomers of a racemic drug may differ in metabolic behavior as a consequence of stereoselective interaction with hepatic microsomes[38-42]. The underlying mechanism of stereoselectivity in metabolism, as many studies have shown, was enantiomeric difference in υmax (an indice of enzymatic catalyzing ability) and/or in Km (an index of enzyme affinity to the substrate). For example, the stereoselective N-demethylation of chlorpheniramine was due to enantiomeric differences in Km[43]. Whereas there were little or no difference in Km of the enantiomers of ofloxacin, the stereoselectivities in glucuronidation were caused by enantiomeric differences in υmax[44]. The υmax of the O-demethylation of (-)-tramadol was 1.6 times that of (+)-isomer, but the Km for both enantiomers was same, thus resulted in its stereos elective O-demethylation[45]. Recently, we have also proved that stereoselectivity of propranolol cytochrome P450 metabolism in the rat hepatic microsomes was due to the stereoselectivity of the catalyzing function in enzyme[35]. In this in vitro study, stereoselectivity of propafenone occurred in Km and Clint in the rat hepatic microsomes induced by DEX or BNF, but not in υmax. Combining with the interesting results of Table 2 that stereoselectivity depends on substrate concentration, we suppose that stereoselectivity at low substrate concentration was mainly due to the enantiomeric difference of the enzyme affinity to the substrate, and that insignificant enantiomeric difference in catalyzing abilities resulted in the abolished stereoselectivity at high substrate concentration. Fujita et al[46] also reported that stereoselectivity of propranolol in rat liver microsomes was sometimes altered when the substrate concentration was varied. Augustijns et al[38] observed that the enantiomeric ratio (R/S) of desethylchloroquine was dependent on concentration, and ranged from 8 at 1 microM to 1 at 300 microM. Mutual enantiomer-enantiomer interaction studies at low concentration (1-5 microM) revealed that the formation of (R)-desethylchloroquine was strongly inhibited by (S)-chloroquine. In this in vitro metabolism, enantiomer-enantiomer interaction at enzyme activity centers may also exist at low concentration, resulting in enantiomeric difference of the enzyme affinity to the substrate. This needs to be addressed by additional experiments.

Table 3 indicated that the stereoselectivity in DEX was stronger than in BNF. It maybe explained by the difference in Km of R (-)-propafenone between DEX and BNF group and that the affinity of R (-)-PPF with CYP3A4 was higher than that with CYP1A, and that of S (+)-PPF with CYP3A4 was similar with CYP1A. Tab le 1 showed that CYP1A and CYP3A4 were significantly inducted by BNF and DEX, respectively, and this agreed with the well known documents. In BNF or DEX group, the υmax was also significantly higher than that in the control group (about 1.5-fold), which indicated that CY P1A and CYP3A4 contributed to the metabolism of propafenone. This substantiated the methods used by Botsch et al[47]. In their study, CYP1A2 and CYP3A4 were identified involved in N-desalkylation using specific antibodies and inhibitors and stably expressed cytochrome P450. Km in the control group was significantly lower than that in DEX or BNF group, which indicated that other enzyme with high affinity to substrate involved in metabolism of propafenone. CYP2D6 which had very low value of Km might be one of such enzymes. Due to the lower value of both Km and υmax in control group, the Clint of propafenone enantiomer was not different from that in DEX or BNF group.

The competitive inhibition model (propafenone/nimodipine) suggested that propafenone and nimodipine were both substrates of the same coenzyme. Because nimodipine was a specific substrate of CYP3A4[48,49], the results of inhibition experiment also proved that CYP3A4 contributed to propafenone metabolism. Drug interaction of enantiomer with specific inhibitor of P450 is an important tool in the search for detailed information on the stereoselective metabolism of xenobiotics[1]. Because fluoxetine impeded in vivo met abolism of R-methadone more than that of S-methadone, Eap et al[50] concluded that CYP2D6-mediated methadone metabolism exhibited stereoselectivity. The fact that the AUC ratio for the two enantiomers of reboxetine was minimally affected by ketoconazole treatment indicates similar affinities of the enantiomers for CYP3A4[51]. In the present study, the phenomenon that nimodipine inhibited S (+)-propafenone more than R (-)-isomer also implies that CYP3A4-mediated propafenone metabolism existed stereoselectivity.

Footnotes

Edited by Xu JY

References
1.  Zhou Q, Yao TW, Zeng S. Chiral drug interaction. Zhongguo Yaoxue Zazhi. 1999;34:721-724.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Hofmann U, Pecia M, Heinkele G, Dilger K, Kroemer HK, Eichelbaum M. Determination of propafenone and its phase I and phase II metabolites in plasma and urine by high-performance liquid chromatography-electrospray ionization mass spectrometry. J Chromatogr B Biomed Sci Appl. 2000;748:113-123.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 26]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
3.  Hemeryck A, De Vriendt C, Belpaire FM. Effect of selective serotonin reuptake inhibitors on the oxidative metabolism of propafenone: in vitro studies using human liver microsomes. J Clin Psychopharmacol. 2000;20:428-434.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 26]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
4.  Tan W, Li Q, McKay G, Semple HA. Identification and determination of phase I metabolites of propafenone in rat liver perfusate. J Pharm Biomed Anal. 1998;16:991-1003.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 13]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
5.  Bonato PS, de Abreu LR, de Gaitani CM, Lanchote VL, Bertucci C. Enantioselective HPLC analysis of propafenone and of its main metabolites using polysaccharide and protein-based chiral stationary phases. Biomed Chromatogr. 2000;14:227-233.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 1]  [Reference Citation Analysis (0)]
6.  Dilger K, Hofmann U, Klotz U. Enzyme induction in the elderly: effect of rifampin on the pharmacokinetics and pharmacodynamics of propafenone. Clin Pharmacol Ther. 2000;67:512-520.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 45]  [Cited by in F6Publishing: 45]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
7.  Fonck K, Haenebalcke C, Hemeryck A, Belpaire F, Jordaens L, Calle P, Buylaert W. ECG changes and plasma concentrations of propafenone and its metabolites in a case of severe poisoning. J Toxicol Clin Toxicol. 1998;36:247-251.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8]  [Cited by in F6Publishing: 8]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
8.  Yao TW, Zhou Q, Hu XY, Zeng S. Stereoselectivity of racemic propafenone in N -depropy lation with rat liver microsomes pretreated by different inducers. Zhongguo Yaolixue Yu Dulixue Zazhi. 2000;14:440-444.  [PubMed]  [DOI]  [Cited in This Article: ]
9.  Chen X, Zhong D, Blume H. Stereoselective pharmacokinetics of propafenone and its major metabolites in healthy Chinese volunteers. Eur J Pharm Sci. 2000;10:11-16.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 25]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
10.  Cai WM, Chen B, Cai MH, Chen Y, Zhang YD. The influence of CYP2D6 activity on the kinetics of propafenone enantiomers in Chinese subjects. Br J Clin Pharmacol. 1999;47:553-556.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 14]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
11.  Cai WM, Chen B, Zhou Y, Zhang YD. Fluoxetine impairs the CYP2D6-mediated metabolism of propafenone enantiomers in healthy Chinese volunteers. Clin Pharmacol Ther. 1999;66:516-521.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 24]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
12.  Nebbia C, Ceppa L, Dacasto M, Carletti M, Nachtmann C. Oxidative metabolism of monensin in rat liver microsomes and interactions with tiamulin and other chemotherapeutic agents: evidence for the involvement of cytochrome P-450 3A subfamily. Drug Metab Dispos. 1999;27:1039-1044.  [PubMed]  [DOI]  [Cited in This Article: ]
13.  Yamazaki H, Komatsu T, Takemoto K, Shimada N, Nakajima M, Yokoi T. Rat cytochrome p450 1A and 3A enzymes involved in bioactivation of tegafur to 5-fluorouracil and autoinduced by tegafur in liver microsomes. Drug Metab Dispos. 2001;29:794-797.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Fontaine F, de Sousa G, Burcham PC, Duchêne P, Rahmani R. Role of cytochrome P450 3A in the metabolism of mefloquine in human and animal hepatocytes. Life Sci. 2000;66:2193-2212.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 44]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
15.  Watanabe M, Tateishi T, Asoh M, Nakura H, Tanaka M, Kumai T, Kobayashi S. Role of CYP3A in haloperidol N-dealkylation and pharmacokinetics in rats. Fundam Clin Pharmacol. 1999;13:337-342.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6]  [Cited by in F6Publishing: 7]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
16.  Sarver JG, Bachmann KA, Zhu D, Klis WA. Ethosuximide is primarily metabolized by CYP3A when incubated with isolated rat liver microsomes. Drug Metab Dispos. 1998;26:78-82.  [PubMed]  [DOI]  [Cited in This Article: ]
17.  Hansen T, Borlak J, Bader A. Cytochrome P450 enzyme activity and protein expression in primary porcine enterocyte and hepatocyte cultures. Xenobiotica. 2000;30:27-46.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 30]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
18.  Price RJ, Surry D, Renwick AB, Meneses-Lorente G, Lake BG, Evans DC. CYP isoform induction screening in 96-well plates: use of 7-benzyloxy-4-trifluoromethylcoumarin as a substrate for studies with rat hepatocytes. Xenobiotica. 2000;30:781-795.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 22]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
19.  Shoda T, Mitsumori K, Onodera H, Toyoda K, Uneyama C, Takada K, Hirose M. Liver tumor-promoting effect of beta-naphthoflavone, a strong CYP 1A1/2 inducer, and the relationship between CYP 1A1/2 induction and Cx32 decrease in its hepatocarcinogenesis in the rat. Toxicol Pathol. 2000;28:540-547.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 48]  [Cited by in F6Publishing: 53]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
20.  Zhao C, Sheryl D, Zhou YX. Effects of combined use of diallyl disulfide and Nacetyl-cysteine on acetaminophen hepatotoxicity in beta-naphthoflavone pretreated mice. World J Gastroenterol. 1998;4:112-116.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  McCune JS, Hawke RL, LeCluyse EL, Gillenwater HH, Hamilton G, Ritchie J, Lindley C. In vivo and in vitro induction of human cytochrome P4503A4 by dexamethasone. Clin Pharmacol Ther. 2000;68:356-366.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 113]  [Cited by in F6Publishing: 110]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
22.  Gibson GG, Skett P.  Introduction to drug metabolism (Second edition). London: Blackie Aca demic & Professional 1994; 217-221.  [PubMed]  [DOI]  [Cited in This Article: ]
23.  Tang YN, Zhang SG, Li L, Zhang WL. Preparation of human liver microsom es and determination of cytochrome P450. Zhongguo Yiyuan Yaoxue Zazhi. 1998;1 8:535-536.  [PubMed]  [DOI]  [Cited in This Article: ]
24.  Zeng S, Li Yan. Determination of stereoselectivity for phase I metabolism of S-(-)- and R-()-ofloxacin by RP-HPLC. Zhejiang Daxue Xuebao (. Yixue Ban). 1999;28:265-267.  [PubMed]  [DOI]  [Cited in This Article: ]
25.  OMURA T, SATO R. THE CARBON MONOXIDE-BINDING PIGMENT OF LIVER MICROSOMES. I. EVIDENCE FOR ITS HEMOPROTEIN NATURE. J Biol Chem. 1964;239:2370-2378.  [PubMed]  [DOI]  [Cited in This Article: ]
26.  Klotz AV, Stegeman JJ, Walsh C. An alternative 7-ethoxyresorufin O-deethylase activity assay: a continuous visible spectrophotometric method for measurement of cytochrome P-450 monooxygenase activity. Anal Biochem. 1984;140:138-145.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 250]  [Cited by in F6Publishing: 184]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
27.  Wrighton SA, Maurel P, Schuetz EG, Watkins PB, Young B, Guzelian PS. Identification of the cytochrome P-450 induced by macrolide antibiotics in rat liver as the glucocorticoid responsive cytochrome P-450p. Biochemistry. 1985;24:2171-2178.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 249]  [Cited by in F6Publishing: 262]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
28.  Zhou Quan, Yao TW, Zeng S. Chiral RP-HPLC assay for propafenone enantiomers in rat liver microsomal incubates and its application in vitro metabolism. Yaoxue Xuebao. 2000;35:370-373.  [PubMed]  [DOI]  [Cited in This Article: ]
29.  Zeng S, Zhong J, Pan L, Li Y. High-performance liquid chromatography separation and quantitation of ofloxacin enantiomers in rat microsomes. J Chromatogr B Biomed Sci Appl. 1999;728:151-155.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 45]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
30.  Li X, Yao TW, Zeng S. Reversed-phase high-performance liquid chromatographic analysis of atenolol enantiomers in rat hepatic microsome after chiral derivatizaton with 2, 3, 4, 6, -tetra-O-acetyl-beta-D-glycopyranosyl isothiocyanate. J Chromatogr B Biomed Sci Appl. 2000;742:433-439.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 17]  [Article Influence: 0.7]  [Reference Citation Analysis (0)]
31.  Zeng S, Mao HQ. Stereoselective determination of fenfluramine enantiomers in rat liver microsomal incubates. J Chromatogr B Biomed Sci Appl. 1999;727:107-112.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5]  [Cited by in F6Publishing: 5]  [Article Influence: 0.2]  [Reference Citation Analysis (0)]
32.  Zeng S, Zhang L, Chen YZ. Chiral gas chromatographic assay with flame ionization detection for amphetamine enantiomers in microsomal incubates. Biomed Chromatogr. 1999;13:33-36.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 1]  [Reference Citation Analysis (0)]
33.  Yao TW, Zeng S. Stereoselective determination of p-hydroxyphenyl-phenylhydantoin enantiomers in rat liver microsomal incubates by reversed-phase high-performance liquid chromatography using beta-cyclodextrin as chiral mobile phase additives. Biomed Chromatogr. 2001;15:141-144.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8]  [Cited by in F6Publishing: 8]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
34.  Yao TW, Zhou Q, Zeng S. Stereoselective determination of propafenone enantiomers in transgenic Chinese hamster CHL cells expressing human cytochrome P450. Biomed Chromatogr. 2000;14:498-501.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 1]  [Reference Citation Analysis (0)]
35.  Li X, Liu Q, Zeng S. Influence of Phenobarbital induction the stereoselectivity of propranolol glucuronidation in rat hepatic microsome. Zhejiang Daxue Xuebao (. Yixue Ban). 1999;28:202-205.  [PubMed]  [DOI]  [Cited in This Article: ]
36.  Li X, Zeng S. Stereoselective propranolol metabolism in two drug induced rat hepatic microsomes. World J Gastroenterol. 2000;6:74-78.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 9]  [Cited by in F6Publishing: 10]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
37.  Yao TW, Zeng S, Li X. Glucuronidation metabolism of ofloxacin enantiom ers in rat microsomes treated with different inducers. Zhongguo Yaolixue Yu Dulixue Zazhi. 2000;14:268-272.  [PubMed]  [DOI]  [Cited in This Article: ]
38.  Augustijns P, Peeters L, Croes K, Verbeeck RK. Stereoselective de-ethylation of chloroquine in rat liver microsomes. Eur J Drug Metab Pharmacokinet. 1999;24:105-108.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 3]  [Article Influence: 0.1]  [Reference Citation Analysis (0)]
39.  Echizen H, Tanizaki M, Tatsuno J, Chiba K, Berwick T, Tani M, Gonzalez FJ, Ishizaki T. Identification of CYP3A4 as the enzyme involved in the mono-N-dealkylation of disopyramide enantiomers in humans. Drug Metab Dispos. 2000;28:937-944.  [PubMed]  [DOI]  [Cited in This Article: ]
40.  Desta Z, Soukhova N, Morocho AM, Flockhart DA. Stereoselective metabolism of cisapride and enantiomer-enantiomer interaction in human cytochrome P450 enzymes: major role of CYP3A. J Pharmacol Exp Ther. 2001;298:508-520.  [PubMed]  [DOI]  [Cited in This Article: ]
41.  Wienkers LC, Allievi C, Hauer MJ, Wynalda MA. Cytochrome P-450-mediated metabolism of the individual enantiomers of the antidepressant agent reboxetine in human liver microsomes. Drug Metab Dispos. 1999;27:1334-1340.  [PubMed]  [DOI]  [Cited in This Article: ]
42.  Li X, Zeng Su. Metabolic characteristics of the optical isomers of propranolol in the cytochrome P450 of rat liver microsomes treated with different inducers. Zhongguo. Yaolixue Yu Dulixue Zazhi. 1999;13:53-56.  [PubMed]  [DOI]  [Cited in This Article: ]
43.  Nomura A, Sakurai E, Hikichi N. [Stereoselective N-demethylation of chlorpheniramine in rat liver microsomes: studies on age and sex differences]. Yakugaku Zasshi. 1995;115:633-640.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 11]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
44.  Okazaki O, Kurata T, Hakusui H, Tachizawa H. Stereoselective glucuronidation of ofloxacin in rat liver microsomes. Drug Metab Dispos. 1991;19:376-380.  [PubMed]  [DOI]  [Cited in This Article: ]
45.  Paar WD, Frankus P, Dengler HJ. The metabolism of tramadol by human liver microsomes. Clin Investig. 1992;70:708-710.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 75]  [Cited by in F6Publishing: 82]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
46.  Fujita S, Umeda S, Funae Y, Imaoka S, Abe H, Ishida R, Adachi T, Masuda M, Kazusaka A, Suzuki T. Regio- and stereoselective propranolol metabolism by 15 forms of purified cytochromes P-450 from rat liver. J Pharmacol Exp Ther. 1993;264:226-233.  [PubMed]  [DOI]  [Cited in This Article: ]
47.  Botsch S, Gautier JC, Beaune P, Eichelbaum M, Kroemer HK. Identification and characterization of the cytochrome P450 enzymes involved in N-dealkylation of propafenone: molecular base for interaction potential and variable disposition of active metabolites. Mol Pharmacol. 1993;43:120-126.  [PubMed]  [DOI]  [Cited in This Article: ]
48.  Liu XQ, Ren YL, Qian ZY, Wang GJ. Enzyme kinetics and inhibition of nimodipine metabolism in human liver microsomes. Acta Pharmacol Sin. 2000;21:690-694.  [PubMed]  [DOI]  [Cited in This Article: ]
49.  Michalets EL. Update: clinically significant cytochrome P-450 drug interactions. Pharmacotherapy. 1998;18:84-112.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Eap CB, Bertschy G, Powell K, Baumann P. Fluvoxamine and fluoxetine do not interact in the same way with the metabolism of the enantiomers of methadone. J Clin Psychopharmacol. 1997;17:113-117.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 57]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
51.  Herman BD, Fleishaker JC, Brown MT. Ketoconazole inhibits the clearance of the enantiomers of the antidepressant reboxetine in humans. Clin Pharmacol Ther. 1999;66:374-379.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 35]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]