You are on page 1of 5

Supplemental material to this article can be found at:

http://dmd.aspetjournals.org/content/suppl/2009/03/19/dmd.108.026203.DC1.html
0090-9556/09/3706-11521156$20.00
DRUG METABOLISM AND DISPOSITION
Copyright 2009 by The American Society for Pharmacology and Experimental Therapeutics
DMD 37:11521156, 2009

Vol. 37, No. 6


26203/3473370
Printed in U.S.A.

Short Communication
The Role of Human Hepatic Cytochrome P450 Isozymes in the
Metabolism of Racemic 3,4-Methylenedioxyethylamphetamine and
Its Single EnantiomersS
Received December 16, 2008; accepted March 12, 2009

ABSTRACT:
incubated using heterologously expressed human P450s, and
the corresponding metabolites dihydroxyethylamphetamine and
methylenedioxyamphetamine were determined by gas chromatography-mass spectrometry after chiral derivatization with S-heptafluorobutyrylprolyl chloride. The highest contributions to both
metabolic steps as calculated from the enzyme kinetic data were
obtained for CYP3A4 and CYP2D6 at substrate concentrations
corresponding to plasma concentrations of recreational users after intake of racemic MDEA. Both metabolic reactions were found
to be enantioselective with a general preference for the S-enantiomers, which was particularly pronounced in the case of CYP2C19.
In conclusion, different pharmacokinetic properties of MDEA enantiomers observed in vivo are therefore partially caused by P450dependent enantioselective metabolism.

3,4-Methylenedioxyethylamphetamine (MDEA, Eve) is chemically


and pharmacologically related to 3,4-methylenedioxymethamphetamine (MDMA, Ecstasy, Adam) (Freudenmann and Spitzer, 2004).
The MDEA enantiomers have different pharmacokinetic properties:
S-MDEA, which produces elevated mood and impairment in conceptually driven cognition, and R-MDEA, which produces increased
depression and enhanced visual feature processing. Concerning
chronic toxicity, data from animal experiments strongly suggest that
both MDMA and MDEA can cause irreversible damage to serotoninergic nerve terminals in the central nervous system (Kalant, 2001; de
la Torre et al., 2004; Freudenmann and Spitzer, 2004; Monks et al.,
2004; Easton and Marsden, 2006). However, most in vitro and animal
studies indicate that compared with MDMA, the neurotoxic potential
of MDEA is lower (Kalant, 2001; Freudenmann and Spitzer, 2004;
Monks et al., 2004).
More recently, we have published studies concerning the involvement of human cytochrome P450 (P450) in the metabolism of racemic
MDMA as well as the metabolites of its enantiomers (Meyer et al.,
2008). In this MDMA study, we were able to show that CYP2C19 (in

addition to CYP2D6) is the most enantioselective P450 isozyme


enzyme toward the two main metabolic steps, namely N-demethylation and demethylenation, with a preference for the S-enantiomer.
These results could in part explain the different pharmacokinetics of
R- and S-MDMA in vivo. As shown in Fig. 1, in vivo and in vitro
studies showed that the main metabolic steps of MDEA are the same
as those of MDMA, namely N-deethylation and demethylenation
(Maurer et al., 2000). Later, MDEA was investigated concerning
enantioselective pharmacokinetics in vivo (Brunnenberg and Kovar,
2001; Buechler et al., 2003), and the plasma half-life of R-MDEA was
found to be longer than that of S-MDEA. Accordingly, the plasma
concentrations of the S-enantiomers of the main metabolites N-ethyl4-hydroxy-3-methoxyamphetamine and 3,4-methylenedioxyamphetamine (MDA) were much higher than those of the R-enantiomers.
Enantioselective pharmacokinetics of MDEA resulting in higher
plasma concentrations of R-MDEA were also confirmed by other
authors (Spitzer et al., 2001; Meyer et al., 2002; Peters et al., 2003).
Enantioselective metabolism is the most likely explanation for the
enantioselective pharmacokinetics of MDEA. Previously published
data from inhibition studies and with recombinant P450 (Kreth et al.,
2000; Maurer et al., 2000) indicated that CYP1A2, 2B6, 2D6, and
3A4 are involved in the N-deethylation and/or demethylenation of
MDEA but did not provide any information about the possible enantioselectivity of these metabolic steps. Hence, no systematic enzyme
kinetic data are available for (enantioselective) MDEA metabolism by

Article, publication date, and citation information can be found at


http://dmd.aspetjournals.org.
doi:10.1124/dmd.108.026203.

S The online version of this article (available at http://dmd.aspetjournals.org)


contains supplemental material.

ABBREVIATIONS: MDEA, 3,4-methylenedioxyethylamphetamine; MDMA, 3,4-methylenedioxymethamphetamine; P450, cytochrome P450; MDA,


3,4-methylenedioxyamphetamine; CL, clearance; MAB3A4, monoclonal antibody inhibitory to 3A4.
1152

Downloaded from dmd.aspetjournals.org at ASPET Journals on January 9, 2015

The 3,4-methylenedioxy-methamphetamine (MDMA)-related designer drug 3,4-methylenedioxyethylamphetamine (MDEA, Eve) is


a chiral compound that is mainly metabolized by N-deethylation
and demethylenation during phase I metabolism. The involvement
of several cytochrome P450 (P450) isozymes in these metabolic
steps has been demonstrated by inhibition assays using human
liver microsomes. 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 first to elucidate the
contribution of the relevant human P450 isozymes in the demethylenation as well as in the N-dealkylation of racemic MDEA and its
single enantiomers and secondly to compare these findings with
recently published data concerning the enantioselective metabolism of MDMA. Racemic MDEA and its single enantiomers were

ROLE OF HUMAN P450s IN MDEA ENANTIOMERS METABOLISM

1153

recombinant isozymes. However, kinetic data for each involved


isozyme could help to explain the differences in MDMA/MDEA
metabolism as well as the different pharmacokinetics of the enantiomers in vivo. Therefore, the aim of the present study was to obtain
enantioselective enzyme kinetic data of N-deethylation and demethylenation of MDEA by the 10 P450s most relevant in human drug
metabolism and to compare these data with the kinetics of MDMA to
reveal and explain the differences in enantioselective metabolism.
Materials and Methods
In principle, materials and methods were the same as described previously
(Meyer et al., 2008). Therefore, only the differences are mentioned in the
following section. Racemic hydrochlorides of MDEA (98.5% purity) and
N-ethyl-3,4-dihydroxyamphetamine were obtained from Lipomed (Bad Saeckingen, Germany) and ReseaChem (Burgdorf, Switzerland), respectively. The
MDEA enantiomers (98% purity, each) were prepared in the authors laboratory by enantioseparation as described below.
Separation of Racemic MDEA and High-Performance Liquid Chromatography Conditions. Racemic MDEA was separated in aliquots (100 l) of
an aqueous stock solution (5 mg/ml, 30 mg in total) using a mobile phase
composition of 0.1 M ammonium acetate buffer (pH 6.5)/acetonitrile 85:15
(v/v), a flow rate of 3 ml/min, and UV detection at 263 nm. The
cyclodextrin column was mounted into a freezer at 8C.
The fractions containing the separated enantiomers were collected, and the
enantiomers were isolated from the aqueous part by liquid/liquid extraction at
pH 9 using ethyl acetate (three times using 150 ml each). The extracts were
evaporated to dryness using a Rotavapor (Buchi, Essen, Germany) under
reduced pressure and reconstituted in 1.0 ml of HCl (0.01 M). Thereafter, the
concentrations of the MDEA enantiomers in the resulting solution were determined. The recovery was approximately 75% per enantiomer.
Initial Screening Studies and Kinetic Studies. Incubations were performed with 50 M R,S-MDEA, R-MDEA, and S-MDEA. Kinetic constants of
N-deethylation (expressed as MDA formation) or demethylenation (expressed
as DHEA formation) were derived from incubations with an incubation time of
20 min and a P450 concentration of 40 pmol/ml (N-deethylation) and 30
pmol/ml (demethylenation). The substrate concentrations shown in Supplemental Table S1 were used.
Calculation of enzyme kinetic constants was similar to the previously
published method (Meyer et al., 2008). In brief, the Michaelis-Menten equation (eq. 1) was used to calculate apparent Km and Vmax values for singleenzyme systems. Eadie-Hofstee plots were used to check for biphasic kinetics.
If the Eadie-Hofstee plot indicated biphasic kinetics, eq. 1 and the alternative
eq. 2 for a two-enzyme model (Clarke, 1998) were applied to the respective
data. For eq. 2, CLint,2 represents the intrinsic clearance or Vmax/Km of the

low-affinity component (Clarke, 1998). If eq. 2 was found to fit the data
significantly better (F test, P 0.05), biphasic kinetics were assumed.
V max [S]
K m [S]

(1)

V max,1 [S]
CLint,2 S]
K m,1 [S]

(2)

V
V

The relative activity factor 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 corrected activities (contributions), the percentages of net clearance by
a particular P450 at a certain substrate concentration, can be calculated
according to eq. 3:
clearanceenyzme[%]

contributionenyzme
100
contributionenzyme

(3)

Inhibition Studies with Chemical Inhibitors or the MAB3A4. The effect


of 3 M quinidine and 6 M omeprazole on (R,S-)DHEA formation was
assessed in incubations containing 0.5 mg of human liver microsome protein/ml and 1 M or 5 M R,S-MDEA, R-MDEA, or S-MDEA (n 6 each).
Control incubations contained none of these chemical inhibitors. Significance
of inhibition was tested by one-tailed unpaired t test using GraphPad Prism
3.02 software (GraphPad Software Inc., San Diego, CA). The effect of MAB
inhibitory to CYP3A4 DHEA formation was assessed in the same way as
described for MDMA (Meyer et al., 2008).
Sample Preparation. Sample preparation and metabolite quantification
were also performed according to Meyer et al. (2008) with the following
modifications: For detection of (R,S-)DHEA and dihydroxybenzylamine, the
gas chromatography 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, 280C; carrier gas, helium; flow
rate, 1 ml/min; column temperature, 150C increased to 250C at 40C/min, to
290C at 2C/min, and finally to 310C. For quantification, the following
target ions (m/z) were used in the selected-ion monitoring mode: m/z 437 for
MDA-d5, 432 for MDA, 430 for dihydroxybenzylamine, and 780 for DHEA.

Results and Discussion


The applied gas chromatography-mass spectrometry conditions
provided separation of DHEA enantiomers (Supplemental Figure S1),
and the chosen target ions were selective for the analytes under these
conditions as proven with blank samples (control microsomes without

Downloaded from dmd.aspetjournals.org at ASPET Journals on January 9, 2015

FIG. 1. The two main metabolic steps of R- and S-MDEA leading to the formation of the corresponding enantiomers of dihydroxyethylamphetamine (DHEA) and MDA.

1154

MEYER ET AL.
TABLE 1
Kinetic data for the two main metabolic steps of (R,S-)MDEA

Units used are as follows: Vmax, picomoles per minute per picomole; and Km, micromolar.
MDA Formation

R,S-MDEA

R-MDEA
(racemic MDEA)

S-MDEA
(racemic MDEA)

S-MDEA
(single enantiomers)

Vmax

Km

Vmax/Km

Vmax

Km

Vmax/Km

19 0.6
23 0.8
22 0.9
3.3 0.4*
2.4 0.4*
10 0.4
12 0.5
10 0.5
3.2 0.2*
2.1 0.2*
9.0 0.4
11 0.7
12 0.4
2.2 0.2*
3.8 0.3*
13 0.5
18 1.2
15 0.4
2.5 0.5*
13 2.0*
16 0.8
19 0.8
8.7 1.5*
5.1 0.3
3.2 0.4*

91 11
124 17
110 17
4.6 1.9*
10 4.8*
65 11
82 13
126 6
6.5 0.6*
16 2.4*
32 6
53 15
31 19
11 0.2*
18 1.7*
179 21
115 34
187 19
3.6 2.7*
76 18*
126 21
99 20
1.8 0.9*
20 5
7.3 2.5*

0.21
0.19
0.20
0.72
0.24
0.15
0.15
0.08
0.49
0.13
0.28
0.21
0.39
0.20
0.21
0.07
0.16
0.08
0.69
0.17
0.13
0.19
4.83
0.26
0.44

ND
ND
3.0 0.1
9.2 0.5
0.8 0.1*
ND
ND
ND
4.4 0.3
0.4 0.2*
ND
ND
2.9 0.3
4.8 0.3
0.3 0.08*
ND
ND
2.4 0.1
9.3 0.2
1.1 0.08*
ND
ND
3.3 0.1
9.5 0.1
0.7 0.2*

ND
ND
54 7
2.8 0.9
45 16*
ND
ND
ND
2.0 0.6
21 5*
ND
ND
25 8
1.0 0.3
13 4*
ND
ND
139 20
1.4 0.2
38 7*
ND
ND
18 2.8
0.7 0.07
36 8*

ND
ND
0.06
3.29
0.02
ND
ND
ND
2.20
0.02
ND
ND
0.12
4.80
0.02
ND
ND
0.02
6.64
0.03
ND
ND
0.18
13.57
0.02

ND, not determined.


* Kinetic data estimated according to eq. 2.

substrate and standard) and zero samples (control microsomes without


substrate but with standard). The method showed good linearity in a
range of 0.05 to 5.0 M DHEA or MDA.
In the initial activity screening, CYP1A2, CYP2B6, CYP2C19,
CYP2D6, and CYP3A4/3A5 were found to be capable of catalyzing
the N-deethylation, and CYP1A2, CYP2C19, CYP2D6, and CYP3A4
were found to be capable of catalyzing the demethylenation of racemic MDEA and its single enantiomers. The kinetic parameters could
not be determined for N-deethylation by CYP3A5 and for demethylenation by CYP1A2 because of insufficient activity of these isozymes
with respect to these reactions. The resulting Km and Vmax values of
the other isozymes are listed in Table 1. Figure 2 shows plots of the
percentages of net clearance as calculated from the relative activity
factor-corrected kinetic data versus substrate concentrations up to 10
M. The data for total metabolite formation are shown in Fig. 2.
Most kinetic data followed the expected classic hyperbolic Michaelis-Menten kinetics as shown in Supplemental Figure S2, a d. Visual
inspection of the data in these figures gave evidence of biphasic
kinetics in some cases, which led to significantly better results for eq.
2 (F test, P 0.05). The corresponding Eadie-Hofstee plot clearly
confirmed this result (data not shown). Hence, the respective kinetic
parameters were estimated by fitting the data into eq. 2 for a biphasic
kinetic. The second enzyme component is represented by CLint,2
[S]. The corresponding Vmax1/Km1 values are presented in Table 1.
Enantioselective phase I metabolism might be the most likely
explanation for the enantioselective pharmacokinetics of racemic
MDEA. Data from inhibition studies and with recombinant P450
(Kreth et al., 2000; Maurer et al., 2000) indicated that 1A2, 2B6, 2D6,
and 3A4 are involved in the N-deethylation and/or demethylenation of
MDEA. CYP2D6 expressed in Escherichia coli was investigated
toward racemic MDEA by Keizers et al. (2005), and the Km determined for DHEA formation was in the same range but lower than in
our study (1.1 versus 2.8 M), which might be explained by different

expression systems and/or different incubation conditions, particularly


incubation time, which was rather long in the aforementioned study
(10 min). In the present study, we have observed that DHEA formation by CYP2D6 was only linear in a time range up to 2.5 min. In
addition, Keizer et al. (2005) did not observe formation of MDA after
incubation of CYP2D6. The reason for this finding could be insufficient sensitivity of the used analytical procedure.
The study presented here is the first to provide enantioselective
enzyme kinetic data of N-deethylation and demethylenation of MDEA
by 10 recombinant P450s, which are the most relevant in human drug
metabolism. CYP2C19 was among these enzymes, which is known to
be one of the five most abundant P450s involved in xenobiotic
metabolism (Guengerich, 2005), but had so far not been considered in
studies on MDEA metabolism.
The experiments were performed with racemic MDEA as consumed by recreational users as well as single MDEA enantiomers to
check whether certain P450s specifically metabolize certain enantiomers and whether the enantiomers influence each others metabolism. Duration and protein content of all incubations in the experiments on enzyme kinetics 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.
As already discussed for MDMA in our previous article (Meyer et
al., 2008), there are inherent differences in the Km and Vmax values of
the single enantiomer kinetics versus the values of racemic MDEA in
incubations with racemic MDEA. In brief, in incubations of racemic
MDEA, R- and S-MDEA are competitors for the limited number of
active sites in the incubation mixture. Hence, at saturation, approximately half of the active sites are busy transforming R-MDEA while
the other half is busy transforming S-MDEA. However, in incubations
of single enantiomers, these enantiomers must occupy all active sites,

Downloaded from dmd.aspetjournals.org at ASPET Journals on January 9, 2015

R-MDEA
(single enantiomers)

CYP1A2
CYP2B6
CYP2C19
CYP2D6
CYP3A4
CYP1A2
CYP2B6
CYP2C19
CYP2D6
CYP3A4
CYP1A2
CYP2B6
CYP2C19
CYP2D6
CYP3A4
CYP1A2
CYP2B6
CYP2C19
CYP2D6
CYP3A4
CYP1A2
CYP2B6
CYP2C19
CYP2D6
CYP3A4

DHEA Formation

ROLE OF HUMAN P450s IN MDEA ENANTIOMERS METABOLISM

and hence their concentrations must be much higher as in the respective incubations of the racemate to reach saturation.
CYP2D6 turned out to be the isozyme with the highest affinity
toward racemic MDEA and its enantiomers (Table 1). In addition,
there is a generally higher affinity toward S-MDEA than R-MDEA
concerning all involved isozymes. The obvious difference in the Km
values of racemic MDEA and the respective enantiomers might be
caused by interactions of R- and S-MDEA in incubations of the

racemate. CYP2D6 also had the highest capacity for demethylenation


of racemic MDEA and its enantiomers. The enzyme kinetic data
reported here are in part different to those reported by Kreth et al.
(2000). This finding can be explained by 1) the substrate concentrations used by Kreth et al. (2000), which were 200 M and therefore
far away from the concentrations expected as plasma concentration of
recreational users as used in our study, and 2) by the fact that some
data were derived from inhibition experiments with supposedly specific chemical inhibitors in human liver microsomes rather than
cDNA-expressed single enzymes as described here. The latter is the
most probable reason why the in vivo contribution of some isozymes
(e.g., CYP1A2) was considerably overestimated by Kreth et al.
(2000).
In the present study, we calculated the percentages of net clearance
for substrate concentrations ranging from 1 to 10 M according to
Meyer et al. (2008) to model the involvement of the studied P450s
over the relevant concentration range (Fig. 2). At low substrate
concentrations (1 M), MDEA biotransformation was predominantly
catalyzed by CYP2D6 (40 50%) and CYP3A4 (30 40%). Whereas,
monitoring only the N-deethylation, relevance of CYP2D6 decreased
to 20% and contribution of CYP3A4 increased to 60%. In addition, at
high substrate concentrations (up to 10 M), the relevance of
CYP2D6 in total biotransformation decreased and the other isozymes
became increasingly important.
To confirm the role of CYP2D6, CYP2C19, and CYP3A4 in the
most important step in MDEA metabolism that leads to the potentially
toxic metabolite DHEA, inhibition experiments were performed using
pooled human liver microsomes according to the described procedure
(Meyer et al., 2008). Quinones might be formed in vivo via oxidation
of DHEA, and two major mechanisms of quinone-induced cytotoxicity have been identified ##. Quinones are reduced to radicals by
cellular reductases, and these radicals undergo rapid autoxidation with
formation of superoxide and H2O2. In addition, quinones can react
with cellular nucleophiles.
The inhibition experiments were performed at two substrate concentrations representing concentrations expected in recreational users,
namely 1 and 5 M to account for the above-mentioned concentration
dependence of the involvement of individual P450s in MDEA demethylenation. In general, the demethylenation was inhibited significantly at both substrate concentrations. Quinidine had the strongest
inhibition effects at low substrate concentrations, whereas inhibition
by MAB3A4 had the strongest effect at higher concentrations (Supplemental Data, Figure S3). This observation is in line with the
calculated percentages of net clearance as shown in Fig. 2. The
importance of CYP2D6 and CYP3A4 is critical from the toxicological
point of view, because both isozymes can be inhibited by quite a
number of different xenobiotics. This effect means that in case of
coingestion, the metabolism of MDEA might be considerably
reduced.
Marked enantioselectivity for the S-enantiomer [km(R-MDMA)/km(SMDMA) 1.5] was observed for both metabolic steps by CYP2C19,
for demethylenation by CYP2D6, and for N-deethylation by CYP3A4.
In addition, CYP2D6 showed a preference for R-MDMA with respect
to deethylation.
Considering these findings along with the fact that demethylenation
is the major metabolic step of MDEA metabolism in vivo, the different pharmacokinetic properties of the MDEA enantiomers are therefore most likely attributable to enantioselective metabolism by various
P450s. CYP2D6 and CYP3A4 should be most important in this
context, because together they account for approximately 80% of net
clearance especially at low substrate concentrations. This result must
be considered when trying to estimate the time of ingestion from

Downloaded from dmd.aspetjournals.org at ASPET Journals on January 9, 2015

FIG. 2. Plots of calculated total net clearance (in %) versus substrate concentration
for the N-deethylation and demethylenation.

1155

1156

MEYER ET AL.

enantiomer ratios in plasma as proposed for, e.g., MDMA by Fallon


et al. (1999) and Peters et al. (2003), because the time course of such
ratios may be considerably different in CYP2D6 poor metabolizers or
in case of inhibition of CYP2D6/CYP3A4 by 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 DHEA 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 DHEA.
Acknowledgments. We thank Armin A. Weber, Carsten Schroeder, Andrea E. Schwaninger, and Katja Niesel for assistance and
helpful discussions.

References
Brunnenberg M and Kovar KA (2001) Stereospecific analysis of ecstasy-like N-ethyl-3,4methylenedioxyamphetamine and its metabolites in humans. J Chromatogr B Biomed Sci Appl
751:9 18.
Buechler J, Schwab M, Mikus G, Fischer B, Hermle L, Marx C, Gron G, Spitzer M, and Kovar
KA (2003) Enantioselective quantitation of the ecstasy compound (R)- and (S)-N-ethyl-3,4methylenedioxyamphetamine and its major metabolites in human plasma and urine. J Chromatogr B Analyt Technol Biomed Life Sci 793:207222.
Clarke SE (1998) In vitro assessment of human cytochrome P450. Xenobiotica 28:11671202.
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 84:121131.
de la Torre R, Farre M, Roset PN, Pizarro N, Abanades S, Segura M, Segura J, and Cam J (2004)
Human pharmacology of MDMA: pharmacokinetics, metabolism, and disposition. Ther Drug
Monit 26:137144.
Easton N and Marsden CA (2006) Ecstasy: are animal data consistent between species and can
they translate to humans? J Psychopharmacol 20:194 210.
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 [Erratum in Clin Chem 1999 Sep;45:1585].

Address correspondence to: Dr. Hans H. Maurer, Department of Experimental and Clinical Toxicology, Institute of Experimental and Clinical Pharmacology
and Toxicology, Saarland University, Building 46, D-66421 Homburg (Saar), Germany. E-mail: hans.maurer@uks.eu

Downloaded from dmd.aspetjournals.org at ASPET Journals on January 9, 2015

MARKUS R. MEYER
Department of Experimental and
FRANK T. PETERS
Clinical Toxicology, Institute of Experimental
HANS H. MAURER
and Clinical Pharmacology and Toxicology,
Saarland University, Homburg (Saar), Germany

Freudenmann RW and Spitzer M (2004) The neuropsychopharmacology and toxicology of


3,4-methylenedioxy-N-ethyl-amphetamine (MDEA). CNS Drug Rev 10:89 116.
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
7:251264.
Guengerich FP (2005) Human cytochrome P450 enzymes, in Cytochrome P450 - Structure,
Mechanism, and Biochemistry (Ortiz-de-Montellano PR ed) pp 377530, Kluwer Academic/
Plenum Publishers, New York.
Kalant H (2001) The pharmacology and toxicology of ecstasy (MDMA) and related drugs.
CMAJ 165:917928.
Keizers PH, de Graaf C, de Kanter FJ, Oostenbrink C, Feenstra KA, Commandeur JN, and
Vermeulen NP (2005) Metabolic regio- and stereoselectivity of cytochrome P450 2D6 towards
3,4-methylenedioxy-N-alkylamphetamines: in silico predictions and experimental validation.
J Med Chem 48:6117 6127.
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 59:15631571.
Maurer HH, Bickeboeller-Friedrich J, Kraemer T, and Peters FT (2000) Toxicokinetics and
analytical toxicology of amphetamine-derived designer drugs (Ecstasy). Toxicol Lett 112:
133142.
Meyer A, Mayerhofer A, Kovar KA, and Schmidt WJ (2002) Enantioselective metabolism of the
designer drugs 3,4-methylenedioxymethamphetamine (ecstasy) and 3,4-methylenedioxyethylamphetamine (eve) isomers in rat brain and blood. Neurosci Lett 330:193197.
Meyer MR, Peters FT, and Maurer HH (2008) The role of human hepatic cytochrome P450
isozymes in the metabolism of racemic MDMA and its enantiomers. Drug Metab Dispos
36:23452354.
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 26:132136.
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 27:552559.
Spitzer M, Franke B, Walter H, Buechler J, Wunderlich AP, Schwab M, Kovar KA, Hermle L,
and Gron G (2001) Enantio-selective cognitive and brain activation effects of N-ethyl-3,4methylenedioxyamphetamine in humans. Neuropharmacology 41:263271.
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.

You might also like