Skip to main content
Advertisement

Main menu

  • Home
  • Articles
    • Current Issue
    • Fast Forward
    • Latest Articles
    • Archive
  • Information
    • Instructions to Authors
    • Submit a Manuscript
    • FAQs
    • For Subscribers
    • Terms & Conditions of Use
    • Permissions
  • Editorial Board
  • Alerts
    • Alerts
    • RSS Feeds
  • Virtual Issues
  • Feedback
  • Other Publications
    • Drug Metabolism and Disposition
    • Journal of Pharmacology and Experimental Therapeutics
    • Molecular Pharmacology
    • Pharmacological Reviews
    • Pharmacology Research & Perspectives
    • ASPET

User menu

  • My alerts
  • Log in
  • Log out
  • My Cart

Search

  • Advanced search
Drug Metabolism & Disposition
  • Other Publications
    • Drug Metabolism and Disposition
    • Journal of Pharmacology and Experimental Therapeutics
    • Molecular Pharmacology
    • Pharmacological Reviews
    • Pharmacology Research & Perspectives
    • ASPET
  • My alerts
  • Log in
  • Log out
  • My Cart
Drug Metabolism & Disposition

Advanced Search

  • Home
  • Articles
    • Current Issue
    • Fast Forward
    • Latest Articles
    • Archive
  • Information
    • Instructions to Authors
    • Submit a Manuscript
    • FAQs
    • For Subscribers
    • Terms & Conditions of Use
    • Permissions
  • Editorial Board
  • Alerts
    • Alerts
    • RSS Feeds
  • Virtual Issues
  • Feedback
  • Visit dmd on Facebook
  • Follow dmd on Twitter
  • Follow ASPET on LinkedIn
Research ArticleArticle

The Role of Human Hepatic Cytochrome P450 Isozymes in the Metabolism of Racemic 3,4-Methylenedioxy-Methamphetamine and Its Enantiomers

Markus R. Meyer, Frank T. Peters and Hans H. Maurer
Drug Metabolism and Disposition November 2008, 36 (11) 2345-2354; DOI: https://doi.org/10.1124/dmd.108.021543
Markus R. Meyer
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Frank T. Peters
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Hans H. Maurer
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • Article
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF
Loading

Abstract

The entactogen, 3,4-methylenedioxy-methamphetamine (MDMA), is a chiral drug that is mainly metabolized by N-demethylation and demethylenation. The involvement of cytochrome P450 (P450) isozymes in these metabolic steps has been studied by inhibition assays with human liver microsomes and, in part, with heterologously expressed human P450 isozymes. However, a comprehensive study on the involvement of all relevant human P450s has not been published yet. In addition, the chirality of this drug was not considered in these in vitro studies. The aim of the present work was to study the contribution of human P450 isozymes in the N-demethylation and demethylenation of racemic MDMA and its single enantiomers. MDMA and its enantiomers were incubated using heterologously expressed human P450s, and the metabolites were quantified by gas chromatography-mass spectrometry after derivatization with S-heptafluorobutyrylprolyl chloride. The highest contribution for the N-demethylation as calculated from the enzyme kinetic data, were obtained for CYP2B6 (R,S-MDMA), CYP1A2 (R-MDMA), and CYP2B6 (S-MDMA). In the case of the demethylenation, the isozyme with the highest contribution to net clearance for R,S-MDMA, R-MDMA, and S-MDMA was CYP2D6. For the first time, marked enantioselectivity was observed for N-demethylation and demethylenation by CYP2C19 with a preference for the S-enantiomers. In addition, CYP2D6 showed preference for S-MDMA in the case of demethylenation. None of the other isozymes showed major preferences for certain enantiomers. In conclusion, therefore, the different pharmacokinetic properties of the MDMA enantiomers may be caused by enantioselective metabolism by CYP2C19 and CYP2D6.

R,S-3,4-Methylenedioxy-methamphetamine (R,S-MDMA), also known as “Adam” or “Ecstasy,” is a chiral compound and a very popular drug of abuse, leading to feelings of euphoria and energy and a desire to socialize (Kalant, 2001). However, it can induce severe acute toxic symptoms such as tachycardia, hypertension, hyperthermia, and hepatotoxicity (Kalant, 2001). Many severe or even fatal intoxications have been described (Kalant, 2001).

Concerning chronic toxicity, data from animal experiments strongly suggest that these compound can cause irreversible damage to serotoninergic nerve terminals in the central nervous system (Kalant, 2001; de la Torre et al., 2004; Monks et al., 2004; Easton and Marsden, 2006). Decreased levels of the serotonin metabolite 5-hydroxyindoleacetic acid (McCann et al., 1994) and serotonin transporters (McCann et al., 1998) found in recreational MDMA users as compared with control subjects with no history of MDMA exposure point in the same direction. However, because the history of drug abuse of the studied MDMA users relied on self-report, it cannot be excluded that these findings were (in part) attributable to concomitant abuse of other drugs of abuse (Gouzoulis-Mayfrank and Daumann, 2006).

As shown in Fig. 1, in vivo studies with MDMA revealed two main metabolic steps: N-demethylation and demethylenation (Maurer, 1996; Maurer et al., 2000; de la Torre et al., 2004). Systemic metabolism of MDMA may play a role in its neurotoxicity. This was concluded from the observation that direct injection of ecstasy into the brain fails to reproduce the neurotoxic effects seen after systemic administration (Esteban et al., 2001) and from the report that alteration of P450-mediated MDMA metabolism influences MDMA-induced neurotoxicity (Gollamudi et al., 1989; Esteban et al., 2001). Metabolites such as 3,4-dihydroxymethamphetamine (DHMA) are easily oxidized to their corresponding quinones, which can form adducts with glutathione and other thiol-containing compounds (Hiramatsu et al., 1990; Miller et al., 1997; Bai et al., 1999). Recently, such adducts have been implicated in MDMA neurotoxicity (de la Torre and Farre, 2004; Capela et al., 2007).

The S-enantiomer of MDMA is known to be more potent than the R-enantiomer in producing the distinctive subjective effects that are characteristic of ecstasy. They also differ in their dose-response curves for changes in serotonergic function and neurotoxicity, and their in vivo kinetics are known to be different. It has been shown that after ingestion of racemic MDMA, the S-enantiomer is eliminated at a higher rate than the R-enantiomer (Fallon et al., 1999; Kalant, 2001; Kraemer and Maurer, 2002; Peters et al., 2003, 2005; Pizarro et al., 2004). Enantioselective metabolism is the most likely explanation for the enantioselective pharmacokinetics of MDMA. Data from inhibition studies (Kreth et al., 2000; Maurer et al., 2000) and experiments with recombinant CYP2D6 (Tucker et al., 1994; Lin et al., 1997; Kreth et al., 2000) indicated that CYP2D6, CYP3A4, CYP1A2, and CYP2B6 are involved in the N-demethylation and demethylenation of MDMA. Today, there is still only limited knowledge on the enantioselective metabolism of MDMA; data are only available for demethylenation of MDMA by recombinant CYP2D6, indicating a higher affinity for the S-enantiomer (Tucker et al., 1994; Lin et al., 1997). Therefore, the aim of the presented study was to obtain enantioselective enzyme kinetic data of N-demethylation and demethylenation of MDMA by the 10 P450s most relevant in human drug metabolism.

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

The two main metabolic steps of R- and S-MDMA leading to the formation of the corresponding enantiomers of DHMA and MDA.

Materials and Methods

Methanolic solution (100 mg/l) of racemic 3,4-methylenedioxyamphetamine (MDA)-d5 was obtained from LGC Promochem (Wesel, Germany), and racemic hydrochlorides of MDMA and DHMA were from Lipomed (Bad Säckingen, Germany). Single MDMA enantiomers were obtained in the authors' lab through enantioseparation of racemic MDMA as described below. Sodium bicarbonate was from Fluka (Buchs, Switzerland); NADP+ was from BIOMOL Research Laboratories (Plymouth Meeting, PA), isocitrate, isocitrate dehydrogenase, and sodium metabisulfite were from Roth (Karlsruhe, Germany), and 3,4-dihydroxybenzylamine (DHBA) was from Sigma Chemie (Deisenhofen, Germany). Water was purified in the authors' lab using a Millipore filtration unit. All other chemicals and reagents were obtained from Merck (Darmstadt, Germany). The following microsomes were from BD Gentest (Woburn, MA) and delivered by NatuTec (Frankfurt/Main, Germany): baculovirus-infected insect cell microsomes (ICMs; Supersomes) containing 1 nmol/ml human cDNA-expressed CYP1A2, CYP2A6, CYP2B6, CYP2C8, CYP2C9, CYP2C19, CYP2D6, and CYP3A4 or 2 nmol/ml CYP2E1 and CYP3A5; wild-type baculovirus-infected ICMs (control Supersomes), and pooled human liver microsomes (HLMs; 20 mg microsomal protein/ml, 400 pmol total P450/mg protein). After delivery, the microsomes were thawed at 37°C, aliquoted, shock-frozen in liquid nitrogen, and stored at -80°C until use. All chemicals were of analytical grade or the highest purity available. The monoclonal antibody inhibitory to CYP 3A4 (MAB3A4) (10 mg/ml MAB3A4 protein) was delivered by NatuTec. MAB3A4 is an ascites-based antibody preparation from mice and was tested to be selective for inhibition toward human CYP3A4. The derivatization reagent S-heptafluorobutyrylprolyl chloride was synthesized in the authors' laboratory according to Peters et al. (2002).

Separation of Racemic MDMA and HPLC Conditions. MDMA was separated using a Hewlett Packard Series 1050 semipreparative HPLC system consisting of a pump and a variable wavelength detector (λ = 263 nm) that was coupled to an Advantec SF 2120 Super Fraction collector. The stationary phase was a Merck Hibar HPLC ChiraDex column (250 × 10 mm, 5 μm). The mobile phase consisted of a mixture of 0.1 M ammonium acetate buffer adjusted to pH 6.5 with 85% acetic acid and 15% acetonitrile at a flow rate 3 ml/min.

MDMA was separated in aliquots (100 μl) of an aqueous stock solution (5 mg/ml, 40 mg in total). The fractions were collected and checked for optical purity by gas chromatography (GC)-mass spectrometry (MS) as described in detail under Sample Preparation.

The acetonitrile part of the HPLC solvent was evaporated. The remaining part was acidified with 1 ml of 0.01 mol/l HCl, and the enantiomers were isolated from the aqueous part by solid phase extraction using Varian Bond Elut SCX HF cartridges (5 g, 20 ml; Varian, Inc., Palo Alto, CA), previously conditioned with 10 ml of methanol and 10 ml of water. After passage of the fractions, the cartridges were washed with 10 ml of 0.01 M hydrochloric acid and 10 ml of methanol. The compounds were eluted twice with 10 ml of freshly prepared mixture of methanol/aqueous ammonia [96:4 (v/v)]. The eluates were evaporated to dryness under reduced pressure and reconstituted in 1.0 ml of 0.01 M HCl and quantified according to Peters et al. (2005).

Microsomal Incubations. Incubation mixtures (final volume, 50 μl) consisted of 90 mM phosphate buffer, pH 7.4, 5 mM MgCl2, 5 mM isocitrate, 1.2 mM NADP+, 0.5 U/ml isocitrate dehydrogenase, 200 U/ml superoxide dismutase, and substrate at 37°C. The substrate was added after dilution of a 25 mM aqueous stock solution in the above-mentioned phosphate buffer. Reactions were started by addition of ice-cold microsomes and terminated with 5 μl of 60% (w/w) perchloric acid.

Initial Screening Studies. Incubations were performed with 50 μM R,S-MDMA, R-MDMA, or S-MDMA and 50 pmol/ml CYP1A2, CYP2A6, CYP2B6, CYP2C8, CYP2C9, CYP2C19, CYP2D6, CYP2E1, CYP3A4, or CYP3A5 for 30 min. For incubations with CYP2A6 or CYP2C9, phosphate buffer was replaced with 45 or 90 mM Tris buffer, respectively, according to the Gentest manual.

Kinetic Studies. Kinetic constants of N-demethylation (expressed as MDA formation) or demethylenation (expressed as DHMA formation) were derived from incubations with an incubation time of 20 min and a P450 concentration of 40 pmol/ml (N-demethylation) and 30 pmol/ml (demethylenation). The substrate concentrations as provided in Table 1 were used.

View this table:
  • View inline
  • View popup
TABLE 1

Substrate concentration in micromolar used for microsomal incubations

Enzyme kinetic constants were estimated by nonlinear curve-fitting using GraphPad Prism 3.02 software (GraphPad Software Inc., San Diego, CA). The Michaelis-Menten equation (eq. 1) was used to calculate apparent Km and Vmax values for single-enzyme systems. Math Eadie-Hofstee plots were used to check for biphasic kinetics (Clarke, 1998). If the Eadie-Hofstee plot indicated biphasic kinetics, eq. 1 and the alternative eq. 2 for a two-site binding model (Clarke, 1998) were applied to the respective data. If eq. 2 was found to fit the data significantly better (F test, p < 0.05), biphasic kinetics were assumed. Math

Calculation of Relative Activity Factors, Contributions, and Percentages of Net Clearance. The relative activity factor (RAF) approach (Crespi and Miller, 1999; Venkatakrishnan et al., 2000; Grime and Riley, 2006) was used to account for differences in functional levels of redox partners between the two enzyme sources. The turnover rates (TRs) of CYP2C19 [probe substrate (PS) S-mephenytoin], CYP2D6 (PS bufuralol), CYP2B6 (PS 7-ethoxy-4-trifluoromethylcoumarin), and CYP1A2 (PS phenacetin) in ICMs and HLMs were taken from the supplier's data sheets. The RAFs were calculated according to eq. 3. Math The enzyme velocities Venyzme (see eq. 1) for the respective metabolic reactions were calculated at different substrate concentrations and were then multiplied with the corresponding RAF leading to a value, which is defined as “contribution.” The Vmax and Km values (eq. 1) were obtained from the incubations with cDNA-expressed P450s. Math From these corrected activities (contributions), the percentages of net clearance by a particular P450 at a certain substrate concentration can be calculated according to eq. 5:Embedded Image

Chemical Inhibition Studies. The effect of 3 μM quinidine (CYP2D6 inhibitor), 6 μM omeprazole (CYP2C19 inhibitor), or 30 μM α-naphthoflavone (CYP1A2 inhibitor) on DHMA formation was assessed in incubations containing 0.5 mg HLM protein/ml and 1 or 5 μM R,S-MDMA, R-MDMA, or S-MDMA (n = 6 each). Control incubations (n = 6) contained none of these chemical inhibitors. Significance of inhibition was tested by one-tailed unpaired Student's t tests using GraphPad Prism 3.02 software.

Inhibition Studies with MAB3A4. The effect of MAB3A4 to CYP3A4 racemic DHMA formation was assessed in incubations containing 0.4 mg HLM protein/ml and 0.02 mg MAB3A4 protein/ml for 20 min at a concentration of 1 or 5 μM of the drug (n = 6 each). Control incubations contained none of this MAB3A4 (n = 6 each). The MAB3A4 was added to the HLMs incubated for 15 min on ice according to the NatuTec manual. As a control, HLMs were incubated for 15 min on ice without MAB3A4. After this preincubation, HLMs with and without MAB3A4 were added to the incubation mixture. Significance of inhibition was tested by one-tailed unpaired Student's t tests using GraphPad Prism 3.02 software.

Sample Preparation. After termination, the incubation mixtures were centrifuged, and 45 μl of the supernatants was transferred to 1.5-ml reaction caps and diluted with 200 μl of aqueous carbonate buffer (35 g/l sodium bicarbonate and 15 g/l sodium carbonate, pH 9) containing the respective internal standard (IS) racemic MDA-d5 and DHBA (0.02 ng/ml, each) and 3% sodium metabisulfite and 3% EDTA-Na (250 mM each) for preservation of the dihydroxy compounds.

Derivatization was performed according to (Peters et al., 2005), with slight modifications. After adding 20 μl of derivatization reagent (0.1 mol/l S-heptafluorobutyrylprolyl chloride in dichloromethane), the reaction vials were sealed and left on a rotary shaker at ambient temperature for 30 min. After the addition of 100 μl of cyclohexane to the reaction vials, they were resealed and placed on a rotary shaker for 5 min. After phase separation by centrifugation (10,000g for 1 min), the cyclohexane phase was transferred to autosampler vials. Aliquots of 3 μl were injected into the GC-MS.

Enantioselective GC-NICI-MS Quantification. The samples were analyzed by an Agilent Technologies 6890 Series GC system combined with an Agilent Technologies 5973 network mass selective detector, an Agilent Technologies 7683 series injector, and an Agilent Technologies enhanced Chem Station G1701CA (version C.00.00, December 21, 1999; Agilent Technologies, Santa Clara, CA). For detection of R,S-MDA and R,S-MDA-d5, the GC conditions were as follows: splitless injection mode; column, 5% phenyl methyl siloxane (HP-5MS; 30-m × 0.25-mm i.d.; 250-nm film thickness); injection port temperature, 280°C; carrier gas, helium; flow rate, 1 ml/min; and column temperature, 100 increased to 200°C at 30°C/min to 260°C at 5°C/min and to 310°C at 30°C/min. The negative-ion chemical ionization (NICI)-MS conditions were as follows: transfer line heater, 280°C; NICI, methane (2 ml/min); source temperature, 150°C; solvent delay, 11 min; and selected-ion monitoring (SIM) mode with the following ions: m/z 432 for MDA and 437 for MDA-d5. For detection of DHMA and DHBA, the GC and NICI-MS conditions were modified as follows: column temperature, 150 increased to 280°C at 40°C/min and to 310°C at 2°C/min; solvent delay, 13.5 min; and SIM mode with the following ions: m/z 472 for DHMA and 430 for DHBA. MDA and DHMA enantiomers were quantified by comparison of their peak-area ratios (enantiomers of analyte versus corresponding enantiomer of the IS) to calibration curves in which the peak-area ratios of enriched calibrators had been plotted versus their concentrations (0.1, 0.5, 1.0, 1.5, and 5.0 μM) using a weighted (1/x) least-squares linear regression model.

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

A, separation of MDMA using a Chiradex column under the described conditions (UV λ = 263 nm). The first peak is R-MDMA, and the second peak is S-MDMA. B and C, GC-MS of aliquots of the collected R-MDMA (upper mass fragmentogram) and S-MDMA (lower mass fragmentogram) HPLC fractions after derivatization.

Results

Semipreparative Separation of R,S-MDMA. Forty milligrams of racemic MDMA-HCl was separated under the described conditions. The chiral HPLC method provided almost baseline separation of the MDMA enantiomers (Fig. 2A). As illustrated in Fig. 2, B and C, purity check of the isolated enantiomers by GC-MS indicated high optical purities of S-MDMA-HCl (97%) and R-MDMA-HCl (98%), respectively. The elution order was confirmed according to (Peters et al., 2005). The yield was approximately 60% or 12 mg/enantiomer.

Enantioselective GC-NICI-MS Quantification. The separation of the MDA enantiomers was comparable with that described by Peters et al. (2005). The mass fragmentograms shown in Fig. 3 demonstrate that the modified GC-MS conditions applied for analysis of DHMA enantiomers provided sufficient separation. The chosen target ions were selective for the analytes under these conditions as proven with blank samples (control microsomes without substrate and IS) and zero samples (control microsomes without substrate but with IS). The method showed good linearity in a range of 0.01 to 5.0 μM DHMA or MDA (r2 > 0.99).

Initial Activity Screening.N-Demethylation of racemic MDMA and its enantiomers was catalyzed by CYP1A2, CYP2B6, CYP2C19, CYP2D6, and CYP3A4/3A5. Demethylenation of MDMA and its enantiomers was catalyzed by CYP1A2, CYP2B6, CYP2C19, CYP2D6, and CYP3A4/3A5.

Kinetic Studies. The kinetic parameters could not be determined for all P450s capable of catalyzing the monitored metabolic reactions because the activity of CYP3A4/A5 with respect to the N-demethylation and CYP2B6 and CYP3A5 with respect to demethylenation of MDMA was too low. The same was true for CYP2C19-catalyzed demethylenation of R-MDMA in racemic MDMA. The Km and Vmax values for the other P450s are listed in Table 2. The data for CYP1A2, CYP2B6, and CYP3A4 (N-demethylation) and CYP1A2, CYP2D6, CYP2C19 (demethylenation) followed the expected classical hyperbolic Michaelis-Menten kinetics as exemplified for CYP2C19 in Fig. 4A. In contrast, the Eadie-Hofstee plots (data not shown) of the CYP2D6 (N-demethylation) and CYP3A4 (demethylenation) clearly indicated biphasic kinetics, which was statistically confirmed (F test, p < 0.05). Hence, the kinetic parameters were estimated by fitting the data into eq. 2 for a two-site binding model. The resulting Km1 and Vmax1 data are also reported in Table 2. Figure 5 shows plots of the percentages of net clearance as calculated from the RAF-corrected kinetic data versus substrate concentrations up to 10 μM. The data for N-demethylation and demethylenation of racemic MDMA, R-MDMA, and S-MDMA are shown in Fig. 5, A to C, respectively. At low substrate concentrations (1 μM), N-demethylation and demethylenation were predominantly catalyzed by CYP2B6 (42-48%) and CYP2D6 (55-76%), respectively. At high substrate concentrations (10 μM), CYP2B6 (43-50%) and CYP1A2 (36-42%) showed a comparable net clearance with respect to N-demethylation. The same was true for CYP2D6 (33-34%) and CYP1A2 (38-41%) with respect to demethylenation.

View this table:
  • View inline
  • View popup
TABLE 2

Kinetic data for the two main metabolic steps of (R,S-)MDMA

Units used are: Vmax, picomoles per minute per picomole; and Km, micromolar.

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Typical mass fragmentograms of ions m/z 430 for DHBA and m/z 472 for DHMA of an incubation mixture after derivatization.

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Michaelis-Menten plots for MDA (left) and DHMA (right) formation catalyzed by CYP2C19 (A), CYP2D6 (B), CYP1A2 (C), and Michaelis-Menten plots for MDA formation catalyzed by CYP2B6 (D, left) and DHMA formation catalyzed by CYP3A4 (D, right). Data points represent means and ranges (error bars) of duplicate measurements. The solid curves represent formation of R-MDA/R-DHMA from racemic MDMA (upper part) and R-MDMA (lower part). The broken curves represent formation of S-MDA/S-DHMA from racemic MDMA (upper part) and S-MDMA (lower part).

Marked enantioselectivity [km(R-MDMA)/km(S-MDMA) > 1.5] was observed for both metabolic steps by CYP2C19 and for demethylenation additionally by CYP2D6. In addition, CYP1A2 showed a minor preference for S-MDMA with respect to demethylenation.

Chemical Inhibition Studies. The results of the experiments with the chemical inhibitors α-naphthoflavone (CYP1A2), omeprazole (CYP2C19), quinidine (CYP2D6), or the MAB3A4 are presented in Fig. 6.

Discussion

Enantioselective metabolism is the most likely explanation for the enantioselective pharmacokinetics of MDMA. Data from inhibition studies (Kreth et al., 2000; Maurer et al., 2000) and experiments with recombinant CYP2D6 (Tucker et al., 1994; Lin et al., 1997; Kreth et al., 2000) indicated that CYP2D6, CYP3A4, CYP1A2, and CYP2B6 are involved in the demethylenation and N-demethylation of MDMA. Recombinant CYP2D6 was reported to have higher affinity for the S-enantiomer with respect to MDMA demethylenation (Tucker et al., 1994; Lin et al., 1997). The study presented here is the first providing enantioselective enzyme kinetic data of N-demethylation and demethylation of MDMA by the 10 recombinant P450s, which are the most relevant in human drug metabolism. Among these enzymes was CYP2C19, which is known to be one of the five most abundant P450s involved in xenobiotic metabolism (Guengerich, 2005), which had so far not been considered in studies on MDMA metabolism.

The experiments were performed with racemic MDMA as present in ecstasy tablets and single MDMA enantiomers to check whether certain P450s specifically metabolize certain enantiomers and whether the enantiomers influence metabolism of each other. Various buffers and buffer concentrations and organic modifiers were tested for semipreparative isolation of the single enantiomers for these experiments. The best result in enantiomer separation was achieved with KH2PO4 0.1M/acetonitrile 95/5 with 0.1%TEA, but the TEA content and the phosphate buffer turned out to be problematic in the following isolation of the enantiomers by solid phase extraction. The conditions described under Materials and Methods were finally preferred because they yielded sufficient separation combined with favorable properties for further workup. The final products were obtained as yellowish powders of high optical purities. Despite rather low recoveries (∼60% per enantiomer), the isolated amounts were sufficient for further kinetic studies.

Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

A to C, plots of total calculated net clearance versus substrate concentration for the N-demethylation and demethylenation.

The described quantification method according to Peters et al. (2005) was modified to be applicable for the presented kinetic studies. It is based on the derivatization of analyte enantiomers to diastereomers using an optically pure derivatization reagent. The diastereomeric derivatives can then be separated on an achiral GC column. In the present study, the amount of derivatization reagent had to be increased to achieve a reliable and reproducible (triple) derivatization of the DHMA enantiomers. In addition, the oven temperature program, the run time, and the ions monitored in SIM had to be modified to allow separation and detection of the triple derivatives of the dihydroxy compounds, DHMA and DHBA (Fig. 2).

According to the supplier's advice, the incubation conditions chosen in the initial screening experiment are applicable for checking the general involvement of particular P450 enzymes. Because of the very low activity of CYP3A4/3A5 with respect to N-demethylation and the very low activity of CYP2B6 and CYP3A5 to demethylenation, only the other P450s that had shown activity in the initial screening could be characterized with respect to their kinetic profiles. Duration and protein content of all incubations in these studies were within the linear range of metabolite formation (data not shown). Less than 20% of substrate was metabolized in all incubations with exception of the lowest substrate concentrations.

There are inherent differences in the Km and Vmax values of the single enantiomer kinetics versus the values of racemic MDMA in incubations with racemic MDMA, which are discussed in the following. R- and S-MDMA are competitors for the limited number of active sites in the incubation mixture. Each binding site can only transform one molecule at a time, either R-MDMA or S-MDMA. Hence, at saturation, approximately half of the active sites are busy transforming R-MDMA, whereas the other half is busy transforming S-MDMA. In incubations of single enantiomers, however, all active sites are available for one enantiomer. Hence, at saturation, approximately twice as many molecules of the respective enantiomer can be transformed at the same time as in incubations of racemate. This clearly explains the higher Vmax values of the single enantiomer kinetics.

The differences of the Km values can be explained in the same way. Saturation is reached when all available active sites are busy transforming MDMA molecules. In incubations of the racemate, half of the active sites are occupied with R-MDMA and half with S-MDMA. In incubations of single enantiomers, this enantiomer must occupy all active sites; hence, its concentrations must be much higher as in the respective incubations of the racemate to reach saturation. Considering that the Km is defined as the substrate concentration where transformation is half-maximal, it is obvious that Km values obtained from single enantiomer incubations must inherently be higher than those of the respective enantiomer in incubations of the racemate.

CYP2D6 turned out to have the highest affinity toward both main metabolic steps of racemic MDMA and for its enantiomers (Table 2). The obvious difference in the Km values of racemic MDMA and the respective enantiomers might be caused by interactions of R- and S-MDMA in incubations of the racemate. For the difference in the Vmax values observed in the case of CYP1A2, there seems to be no straightforward explanation. It might be attributable to a decreased turnover and complex interactions, e.g., cobinding at the active site of the enzyme, during incubation of racemic MDMA. However, for the time being, interpretation of this obvious difference remains speculative. CYP2D6 also had the highest capacity for demethylenation of R,S-MDMA and R-MDMA, whereas the highest Vmax value for S-MDMA was observed for CYP2C19.

Fig. 6.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 6.

Effect of the chemical inhibitors α-naphthoflavone, 30 μM (CYP1A2), omeprazole, 10 μM (CYP2C19), quinidine, 3 μM (CYP2D6), and the MAB3A4 on demethylenation in incubation mixtures containing 1 or 5 μM R,S-, R-, or S-MDMA. Metabolite formation of the control incubations was set to 100%. Each bar represents the mean of six incubations ± S.E.M. (*, p < 0.5; **, p < 0.05; ***, p < 0.005).

The enzyme kinetic data reported here are considerably different to those reported by Kreth et al. (2000). This can be explained by the fact that the Km and Vmax data of Kreth et al. (2000) were derived from inhibition experiments with human liver microsomes rather than cDNA-expressed single enzymes as described in our study. More precisely, Kreth et al. (2000) had observed biphasic enzyme kinetics in human liver microsomes and assumed the high-affinity component to represent CYP2D6 and the low-affinity component to represent CYP1A2. Based on this assumption, they derived their reported Km and Vmax from the two kinetic components. Considering that not only these two enzymes, but also CYP2B6, CYP3A4, and, as shown in our study, also CYP2C19 are involved in MDMA metabolism, the approach by Kreth et al. (2000) is certainly an oversimplification. In addition, Kreth et al. (2000) assessed the contribution of the various isoenzymes involved in MDMA metabolism solely by inhibition experiments with supposedly specific chemical inhibitors. However, orphenadrine that was used as “specific” inhibitor of CYP2B6 also decreases the marker activity of CYP2D6 and CYP2C9 (Guo et al., 1997), which may have affected the contribution data reported by Kreth et al. (2000).

Common methods of calculating the net clearance are based on the assumption that substrate concentrations are lower than 10% of Km of the respective isozymes (Crespi, 1995; Venkatakrishnan et al., 2000). This is critical if the isozymes involved in the metabolic reaction have a very low Km value, which is considerably below expected plasma concentrations. In the present study, Km values of 0.2 to 0.3 μM were found for MDMA demethylenation by CYP2D6. These were well below expected plasma concentrations of MDMA after recreational doses, which are usually in the range of 1 μM (194 μg/l) (Fallon et al., 1999; de la Torre et al., 2000, 2004; Logan and Couper, 2001; Peters et al., 2003, 2005; Pizarro et al., 2004) but can reach 10 μM and higher in severe intoxications (Peters et al., 2003; Schifano, 2004). In the present study, we chose to calculate the percentages of net clearance for substrate concentrations ranging from 1 to 10 μM (Fig. 5) to model the involvement of the studied P450s over the relevant concentration range. At 1 and 5 μM, CYP2D6 accounts for 55 to 78% and 42 to 46% of net clearance with respect to demethylenation, respectively. This percentage continuously decreases, with increasing substrate concentrations reaching 33 to 34% at 10 μM. In contrast, CYP1A2 becomes increasingly important, with rising substrate concentrations reaching 38 to 41% of net clearance at 10 μM. A similar situation is observed for N-demethylation, where the net clearance of CYP2B6 remains rather constant from 38 to 48% (1 μM) over 41 to 49% (5 μM) to 43 to 50% (10 μM), whereas that of CYP1A2 moderately increases from 33 to 34% over 35 to 39% (5 μM) to 36 to 41%, respectively. This shows that in intoxication cases, the activity of CYP1A2 may play an important role in MDMA metabolism. The difference in the percentage of net clearance observed for CYP2D6 in the case of racemic MDMA and its enantiomers is caused by differences in the Km values for the respective compounds because the Km value takes part in calculation of the net clearance (see eqs. 1 and 5).

To confirm the role of CYP1A2, CYP2D6, CYP2C19, and CYP3A4 in the most important step in MDMA metabolism, which leads to the toxic metabolite DHMA, inhibition experiments were performed using pooled HLMs, with the chemical inhibitors α-naphthoflavone, quinidine, and omeprazole at concentrations according to the literature (Newton et al., 1995; Bourrie et al., 1996; Ko et al., 1997; Clarke, 1998; Venkatakrishnan et al., 2001) and the MAB3A4. These experiments were performed at two substrate concentrations representing concentrations expected in recreational users, namely 1 and 5 μM to account for the above-mentioned concentration dependence of the involvement of individual P450s in MDMA demethylenation. Generally, the demethylenation was inhibited significantly at both substrate concentrations. The only exceptions were CYP1A2 at 1 μM racemic MDMA and CYP3A4 at 5 μM S-MDMA and 1 μM R-MDMA (Fig. 6). Quinidine had the strongest inhibition effects at both substrate concentrations (Fig. 6). This is in line with the calculated percentages of net clearance as shown in Fig. 5, although the extent of inhibition by quinidine is somewhat larger as expected based on the calculations. The importance of CYP2D6 is critical from the toxicological point of view because MDMA is known to be a mechanism-based inhibitor of this isozyme (Heydari et al., 2004). This means that in case of repeated doses, the metabolism of MDMA should be considerably reduced. Indeed, the area under the curve and the maximum plasma concentration values increased by 77 and 29%, respectively, after repeated doses of MDMA (de la Torre et al., 2004). The stronger inhibition effects of the CYP1A2 inhibitor α-naphthoflavone at higher substrate concentrations is also in good agreement with the calculated percentages of net clearance increasing with substrate concentrations. As expected from the calculations of net clearance, inhibition with the CYP2C19 inhibitor omeprazole and the MAB3A4 did not show clear trends from the lower to the higher substrate concentrations. However, the observed inhibition effects were again somewhat larger than the respective percentages of net clearance.

CYP2C19 showed the largest extent of enantioselectivity in both metabolic reactions with a marked preference for the S-enantiomer. The ratios of the Km values of the MDMA enantiomers (R versus S) with respect to demethylenation and N-demethylation were 2.1 and 2.0, respectively. CYP2D6 also showed marked enantioselectivity, but here the respective ratios were 1.6 and 1.7. The other isozymes showed little or no enantioselectivity. Considering these findings along with the fact that demethylenation is the major metabolic step of MDMA metabolism in vivo, the different pharmacokinetic properties of the MDMA enantiomers are therefore most likely attributable to enantioselective demethylenation by CYP2C19 and CYP2D6. CYP2D6 should be most important in this context because it accounts for more 33% of net clearance even at high substrate concentrations as compared with a maximum of 17% for CYP2C19. This must be considered when trying to estimate the time of ingestion from enantiomer ratios in plasma as proposed by (Fallon et al., 1999; Peters et al., 2003) because the time course of such ratios may be considerably different in CYP2D6 poor metabolizers or in the case of inhibition of CYP2D6 by MDMA or other coingested drugs. In addition, it must be considered that correlation of the presented in vitro data with the in vivo situation is not straightforward because in vivo DHMA is further metabolized by O-methylation and/or glucuronidation/sulfation. Enantioselectivity of these phase II reactions might of course also influence the enantiomer ratios in plasma samples, especially those of DHMA.

Acknowledgments

We thank Prof. Dr. Thomas Kraemer, Armin A. Weber, Carsten Schroeder, Christoph Sauer, Andrea E. Schwaninger, and Dirk K. Wissenbach for assistance and helpful discussions.

Footnotes

  • doi:10.1124/dmd.108.021543.

  • ABBREVIATIONS: MDMA, 3,4-methylenedioxy-methamphetamine; P450, cytochrome P450; DHMA, 3,4-dihydroxymethamphetamine; MDA, 3,4-methylenedioxyamphetamine; DHBA, 3,4-dihydroxybenzylamine; ICM, insect cell microsome; GC, gas chromatography; MS, mass spectrometry; RAF, relative activity factor; TR, turnover rate; PS, probe substrate; HLM, human liver microsome; IS, internal standard; NICI, negative-ion chemical ionization; MAB3A4, monoclonal antibody inhibitory to 3A4; HPLC, high-performance liquid chromatography; SIM, selected-ion monitoring.

    • Received March 20, 2008.
    • Accepted August 21, 2008.
  • The American Society for Pharmacology and Experimental Therapeutics

References

  1. ↵
    Bai F, Lau SS, and Monks TJ (1999) Glutathione and N-acetylcysteine conjugates of alpha-methyldopamine produce serotonergic neurotoxicity: possible role in methylenedioxyamphetamine-mediated neurotoxicity. Chem Res Toxicol 12: 1150-1157.
    OpenUrlCrossRefPubMed
  2. ↵
    Bourrié M, Meunier V, Berger Y, and Fabre G (1996) Cytochrome P450 isoform inhibitors as a tool for the investigation of metabolic reactions catalyzed by human liver microsomes. J Pharmacol Exp Ther 277: 321-332.
    OpenUrlAbstract/FREE Full Text
  3. ↵
    Capela JP, Macedo C, Branco PS, Ferreira LM, Lobo AM, Fernandes E, Remião F, Bastos ML, Dirnagl U, Meisel A, et al. (2007) Neurotoxicity mechanisms of thioether ecstasy metabolites. Neuroscience 146: 1743-1757.
    OpenUrlCrossRefPubMed
  4. ↵
    Clarke SE (1998) In vitro assessment of human cytochrome P450. Xenobiotica 28: 1167-1202.
    OpenUrlCrossRefPubMed
  5. ↵
    Crespi CL (1995) Xenobiotic-metabolizing human cells as tools for pharmacological and toxicological research, in Advances in Drug Research (Testa B and Meyer UA eds) pp 179-235, Academic Press, London.
  6. ↵
    Crespi CL and Miller VP (1999) The use of heterologously expressed drug metabolizing enzymes-state of the art and prospects for the future. Pharmacol Ther 84: 121-131.
    OpenUrlCrossRefPubMed
  7. ↵
    de la Torre R and Farré M (2004) Neurotoxicity of MDMA (ecstasy): the limitations of scaling from animals to humans. Trends Pharmacol Sci 25: 505-508.
    OpenUrlCrossRefPubMed
  8. ↵
    de la Torre R, Farré M, Ortuño J, Mas M, Brenneisen R, Roset PN, Segura J, and Camí J (2000) Non-linear pharmacokinetics of MDMA (“ecstasy”) in humans. Br J Clin Pharmacol 49: 104-109.
    OpenUrlCrossRefPubMed
  9. de la Torre R, Farré M, Roset PN, Pizarro N, Abanades S, Segura M, Segura J, and Camí J (2004) Human pharmacology of MDMA: pharmacokinetics, metabolism, and disposition. Ther Drug Monit 26: 137-144.
    OpenUrlCrossRefPubMed
  10. ↵
    Easton N and Marsden CA (2006) Ecstasy: are animal data consistent between species and can they translate to humans? J Psychopharmacol 20: 194-210.
    OpenUrlAbstract/FREE Full Text
  11. ↵
    Esteban B, O'Shea E, Camarero J, Sanchez V, Green AR, and Colado MI (2001) 3,4-Methylenedioxymethamphetamine induces monoamine release, but not toxicity, when administered centrally at a concentration occurring following a peripherally injected neurotoxic dose. Psychopharmacology (Berl) 154: 251-260.
    OpenUrlCrossRefPubMed
  12. ↵
    Fallon JK, Kicman AT, Henry JA, Milligan PJ, Cowan DA, and Hutt AJ (1999) Stereospecific analysis and enantiomeric disposition of 3, 4-methylenedioxymethamphetamine (Ecstasy) in humans. Clin Chem 45: 1058-1069.
    OpenUrlAbstract/FREE Full Text
  13. ↵
    Gollamudi R, Ali SF, Lipe G, Newport G, Webb P, Lopez M, Leakey JE, Kolta M, and Slikker W Jr (1989) Influence of inducers and inhibitors on the metabolism in vitro and neurochemical effects in vivo of MDMA. Neurotoxicology 10: 455-466.
    OpenUrlPubMed
  14. ↵
    Gouzoulis-Mayfrank E and Daumann J (2006) The confounding problem of polydrug use in recreational ecstasy/MDMA users: a brief overview. J Psychopharmacol 20: 188-193.
    OpenUrlAbstract/FREE Full Text
  15. ↵
    Grime K and Riley RJ (2006) The impact of in vitro binding on in vitro-in vivo extrapolations, projections of metabolic clearance and clinical drug-drug interactions. Curr Drug Metab 7: 251-264.
    OpenUrlCrossRefPubMed
  16. ↵
    Guengerich FP (2005) Human cytochrome P450 enzymes, in Cytochrome P450: Structure, Mechanism, and Biochemistry (Ortiz-de-Montellano PR ed) pp 377-530, Kluwer Academic/Plenum Publishers, New York.
  17. ↵
    Guo Z, Raeissi S, White RB, and Stevens JC (1997) Orphenadrine and methimazole inhibit multiple cytochrome P450 enzymes in human liver microsomes. Drug Metab Dispos 25: 390-393.
    OpenUrlAbstract/FREE Full Text
  18. ↵
    Heydari A, Yeo KR, Lennard MS, Ellis SW, Tucker GT, and Rostami-Hodjegan A (2004) Mechanism-based inactivation of CYP2D6 by methylenedioxymethamphetamine. Drug Metab Dispos 32: 1213-1217.
    OpenUrlAbstract/FREE Full Text
  19. ↵
    Hiramatsu M, Kumagai Y, Unger SE, and Cho AK (1990) Metabolism of methylenedioxymethamphetamine: formation of dihydroxymethamphetamine and a quinone identified as its glutathione adduct. J Pharmacol Exp Ther 254: 521-527.
    OpenUrlAbstract/FREE Full Text
  20. ↵
    Kalant H (2001) The pharmacology and toxicology of “ecstasy” (MDMA) and related drugs. CMAJ 165: 917-928.
    OpenUrlAbstract/FREE Full Text
  21. ↵
    Ko JW, Sukhova N, Thacker D, Chen P, and Flockhart DA (1997) Evaluation of omeprazole and lansoprazole as inhibitors of cytochrome P450 isoforms. Drug Metab Dispos 25: 853-862.
    OpenUrlAbstract/FREE Full Text
  22. ↵
    Kraemer T and Maurer HH (2002) Toxicokinetics of amphetamines: metabolism and toxicokinetic data of designer drugs, of amphetamine, methamphetamine and their N-alkyl derivatives. Ther Drug Monit 24: 277-289.
    OpenUrlCrossRefPubMed
  23. ↵
    Kreth K, Kovar K, Schwab M, and Zanger UM (2000) Identification of the human cytochromes P450 involved in the oxidative metabolism of “Ecstasy”-related designer drugs. Biochem Pharmacol 59: 1563-1571.
    OpenUrlCrossRefPubMed
  24. ↵
    Lin LY, Di Stefano EW, Schmitz DA, Hsu L, Ellis SW, Lennard MS, Tucker GT, and Cho AK (1997) Oxidation of methamphetamine and methylenedioxymethamphetamine by CYP2D6. Drug Metab Dispos 25: 1059-1064.
    OpenUrlAbstract/FREE Full Text
  25. ↵
    Logan BK and Couper FJ (2001) 3,4-Methylenedioxymethamphetamine (MDMA, ecstasy) and driving impairment. J Forensic Sci 46: 1426-1433.
    OpenUrlPubMed
  26. ↵
    Maurer HH (1996) On the metabolism and the toxicological analysis of methylenedioxyphenylalkylamine designer drugs by gas chromatography-mass spectrometry. Ther Drug Monit 18: 465-470.
    OpenUrlCrossRefPubMed
  27. ↵
    Maurer HH, Bickeboeller-Friedrich J, Kraemer T, and Peters FT (2000) Toxicokinetics and analytical toxicology of amphetamine-derived designer drugs (“Ecstasy”). Toxicol Lett 112: 133-142.
    OpenUrlCrossRefPubMed
  28. ↵
    McCann UD, Ridenour A, Shaham Y, and Ricaurte GA (1994) Serotonin neurotoxicity after (+/-)3,4-methylenedioxymethamphetamine (MDMA; “Ecstasy”): a controlled study in humans. Neuropsychopharmacology 10: 129-138.
    OpenUrlCrossRefPubMed
  29. ↵
    McCann UD, Szabo Z, Scheffel U, Dannals RF, and Ricaurte GA (1998) Positron emission tomographic evidence of toxic effect of MDMA (“Ecstasy”) on brain serotonin neurons in human beings. Lancet 352: 1433-1437.
    OpenUrlCrossRefPubMed
  30. ↵
    Miller RT, Lau SS, and Monks TJ (1997) 2,5-Bis-(glutathion-S-yl)-alpha-methyldopamine, a putative metabolite of (+/-)-3,4-methylenedioxyamphetamine, decreases brain serotonin concentrations. Eur J Pharmacol 323: 173-180.
    OpenUrlCrossRefPubMed
  31. ↵
    Monks TJ, Jones DC, Bai F, and Lau SS (2004) The role of metabolism in 3,4-(+)-methylenedioxyamphetamine and 3,4-(+)-methylenedioxymethamphetamine (ecstasy) toxicity. Ther Drug Monit 26: 132-136.
    OpenUrlCrossRefPubMed
  32. ↵
    Newton DJ, Wang RW, and Lu AY (1995) Cytochrome P450 inhibitors: evaluation of specificities in the in vitro metabolism of therapeutic agents by human liver microsomes. Drug Metab Dispos 23: 154-158.
    OpenUrlAbstract
  33. ↵
    Peters FT, Kraemer T, and Maurer HH (2002) Drug testing in blood: validated negative-ion chemical ionization gas chromatographic-mass spectrometric assay for determination of amphetamine and methamphetamine enantiomers and its application to toxicology cases. Clin Chem 48: 1472-1485.
    OpenUrlAbstract/FREE Full Text
  34. ↵
    Peters FT, Samyn N, Lamers CT, Riedel WJ, Kraemer T, de Boeck G, and Maurer HH (2005) Drug testing in blood: validated negative-ion chemical ionization gas chromatographic-mass spectrometric assay for enantioselective determination of the designer drugs MDA, MDMA (Ecstasy) and MDEA and its application to samples from a controlled study with MDMA. Clin Chem 51: 1811-1822.
    OpenUrlAbstract/FREE Full Text
  35. ↵
    Peters FT, Samyn N, Wahl M, Kraemer T, De Boeck G, and Maurer HH (2003) Concentrations and ratios of amphetamine, methamphetamine, MDA, MDMA, and MDEA enantiomers determined in plasma samples from clinical toxicology and driving under the influence of drugs cases by GC-NICI-MS. J Anal Toxicol 27: 552-559.
    OpenUrlAbstract/FREE Full Text
  36. ↵
    Pizarro N, Farré M, Pujadas M, Peiró AM, Roset PN, Joglar J, and de la Torre R (2004) Stereochemical analysis of 3,4-methylenedioxymethamphetamine and its main metabolites in human samples including the catechol-type metabolite (3,4-dihydroxymethamphetamine). Drug Metab Dispos 32: 1001-1007.
    OpenUrlAbstract/FREE Full Text
  37. ↵
    Schifano F (2004) A bitter pill: overview of ecstasy (MDMA, MDA) related fatalities. Psychopharmacology (Berl) 173: 242-248.
    OpenUrlCrossRefPubMed
  38. ↵
    Tucker GT, Lennard MS, Ellis SW, Woods HF, Cho AK, Lin LY, Hiratsuka A, Schmitz DA, and Chu TY (1994) The demethylenation of methylenedioxymethamphetamine (“ecstasy”) by debrisoquine hydroxylase (CYP2D6). Biochem Pharmacol 47: 1151-1156.
    OpenUrlCrossRefPubMed
  39. ↵
    Venkatakrishnan K, von Moltke LL, Court MH, Harmatz JS, Crespi CL, and Greenblatt DJ (2000) Comparison between cytochrome P450 (CYP) content and relative activity approaches to scaling from cDNA-expressed CYPs to human liver microsomes: ratios of accessory proteins as sources of discrepancies between the approaches. Drug Metab Dispos 28: 1493-1504.
    OpenUrlPubMed
  40. ↵
    Venkatakrishnan K, Von Moltke LL, and Greenblatt DJ (2001) Human drug metabolism and the cytochromes P450: application and relevance of in vitro models. J Clin Pharmacol 41: 1149-1179.
    OpenUrlCrossRefPubMed
PreviousNext
Back to top

In this issue

Drug Metabolism and Disposition: 36 (11)
Drug Metabolism and Disposition
Vol. 36, Issue 11
1 Nov 2008
  • Table of Contents
  • Table of Contents (PDF)
  • About the Cover
  • Index by author
  • Back Matter (PDF)
  • Editorial Board (PDF)
  • Front Matter (PDF)
Download PDF
Article Alerts
Sign In to Email Alerts with your Email Address
Email Article

Thank you for sharing this Drug Metabolism & Disposition article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
The Role of Human Hepatic Cytochrome P450 Isozymes in the Metabolism of Racemic 3,4-Methylenedioxy-Methamphetamine and Its Enantiomers
(Your Name) has forwarded a page to you from Drug Metabolism & Disposition
(Your Name) thought you would be interested in this article in Drug Metabolism & Disposition.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
Research ArticleArticle

The Role of Human Hepatic Cytochrome P450 Isozymes in the Metabolism of Racemic 3,4-Methylenedioxy-Methamphetamine and Its Enantiomers

Markus R. Meyer, Frank T. Peters and Hans H. Maurer
Drug Metabolism and Disposition November 1, 2008, 36 (11) 2345-2354; DOI: https://doi.org/10.1124/dmd.108.021543

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Share
Research ArticleArticle

The Role of Human Hepatic Cytochrome P450 Isozymes in the Metabolism of Racemic 3,4-Methylenedioxy-Methamphetamine and Its Enantiomers

Markus R. Meyer, Frank T. Peters and Hans H. Maurer
Drug Metabolism and Disposition November 1, 2008, 36 (11) 2345-2354; DOI: https://doi.org/10.1124/dmd.108.021543
del.icio.us logo Digg logo Reddit logo Twitter logo CiteULike logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • Materials and Methods
    • Results
    • Discussion
    • Acknowledgments
    • Footnotes
    • References
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF

Related Articles

Cited By...

More in this TOC Section

  • In Vivo Functional Effects of CYP2C9 M1L
  • Clearance pathways: fevipiprant with probenecid perpetrator
  • Predicting Volume of Distribution from In Vitro Parameters
Show more Articles

Similar Articles

  • Home
  • Alerts
Facebook   Twitter   LinkedIn   RSS

Navigate

  • Current Issue
  • Fast Forward by date
  • Fast Forward by section
  • Latest Articles
  • Archive
  • Search for Articles
  • Feedback
  • ASPET

More Information

  • About DMD
  • Editorial Board
  • Instructions to Authors
  • Submit a Manuscript
  • Customized Alerts
  • RSS Feeds
  • Subscriptions
  • Permissions
  • Terms & Conditions of Use

ASPET's Other Journals

  • Journal of Pharmacology and Experimental Therapeutics
  • Molecular Pharmacology
  • Pharmacological Reviews
  • Pharmacology Research & Perspectives
ISSN 1521-009X (Online)

Copyright © 2021 by the American Society for Pharmacology and Experimental Therapeutics