Modulating factors in MDMA induced neurotoxicity - UvA Scripties

36
Literature thesis Date 17-01-2020 Student name Sebastiaan van Bruchem Student number 10213929 Email [email protected] Supervisor Prof. dr. Liesbeth Reneman Second assessor Prof. dr. Paul Lucassen Institution Amsterdam UMC Department Radiology and Nuclear Medicine Address Location AMC Meibergdreef 9 1105 AZ, Amsterdam Modulating factors in MDMA induced neurotoxicity

Transcript of Modulating factors in MDMA induced neurotoxicity - UvA Scripties

Literature thesis Date 17-01-2020 Student name Sebastiaan van Bruchem Student number 10213929 Email [email protected] Supervisor Prof. dr. Liesbeth Reneman Second assessor Prof. dr. Paul Lucassen Institution Amsterdam UMC Department Radiology and Nuclear Medicine Address Location AMC Meibergdreef 9 1105 AZ, Amsterdam

Modulating factors in MDMA induced neurotoxicity

1

2

Contents Abstract ................................................................................................................................................... 3

Introduction ............................................................................................................................................. 4

Cognitive problems as a result of MDMA use ..................................................................................... 4

Evidence of MDMA induced neurotoxicity ......................................................................................... 5

Neurotoxicity in animals .................................................................................................................. 6

Neurotoxicity in humans ................................................................................................................. 6

Thesis aim ............................................................................................................................................ 7

Metabolism of MDMA ............................................................................................................................. 9

Causes of MDMA neurotoxicity ............................................................................................................. 10

Genetic susceptibility ............................................................................................................................ 12

CYP2D6 .............................................................................................................................................. 12

Other cytochrome p450 enzymes ..................................................................................................... 13

Catechol-O-Methyltransferase .......................................................................................................... 14

Monoamine Oxidase-B ...................................................................................................................... 15

Sex ..................................................................................................................................................... 15

Ethnicity ............................................................................................................................................. 17

Environmental ....................................................................................................................................... 18

Drug-drug interactions .......................................................................................................................... 19

Caffeine ............................................................................................................................................. 19

Tobacco ............................................................................................................................................. 19

Cannabis ............................................................................................................................................ 20

Alcohol ............................................................................................................................................... 20

Stimulants .......................................................................................................................................... 20

Psychedelics....................................................................................................................................... 21

Conclusion/recommendations .............................................................................................................. 22

References ............................................................................................................................................. 25

3

Abstract MDMA is an illicit drug with widespread use in recreational and therapeutic settings. Because of its

widespread use, it is important to gain insight into the possibly damaging effects of MDMA. To date,

research into the neurotoxic effects of MDMA has been inconsistent however. Animal studies clearly

indicate selective serotonergic neurotoxicity produced by MDMA. Human studies find memory

problems in MDMA users. However, research using markers of serotonergic damage in humans give

inconsistent results. The discrepancy between animal and human research is most likely caused by

the wide variety of factors modulating an individual’s susceptibility to the neurotoxic effects of

MDMA. This thesis gives an overview of some of the most important modulating factors.

Toxicity of MDMA is mainly caused by neurotoxic metabolites of MDMA and excessive dopamine

influx into serotonergic neurons. Polymorphisms of CYP2D6 and COMT, both enzymes involved in

MDMA metabolism, greatly influence susceptibility to the neurotoxic effects of MDMA. CYP1A2,

CYP2B6, CYP3A4, and CYP2C19 polymorphisms are minor modulating factors, while there is an

indication MAO-B polymorphism is also a modulating factor in MDMA induced neurotoxicity. Women

seem to be more susceptible to men with regards to the neurotoxic effects of MDMA, at least at

higher doses of MDMA. African American people, and Asian people at higher doses of MDMA, seem

less susceptible than Caucasians. Ambient temperature and excessive physical exertion often seen at

dance parties exacerbate MDMA induced neurotoxicity. Lastly, most MDMA users are polydrug users

and this greatly influences susceptibility to the neurotoxic effects of MDMA. Cannabis and

low/moderate doses of alcohol seem protective, while stimulants and psychedelics exacerbate

MDMA induced neurotoxicity.

The information presented in this thesis can be valuable in assessing the impact of modulating

factors in susceptibility to the neurotoxic effects of MDMA. By correcting for these modulating

factors research accuracy can be greatly improved, ultimately leading to an improved understanding

of the damaging effects of MDMA. In addition, this information can be used to improve harm-

reduction practices for users which is especially relevant for use in therapeutic setting.

4

Introduction MDMA (3,4-methylenedioxymethamphetamine) or “ecstasy” is a psychoactive drug which induces

euphoria, empathy, increased energy and sometimes mild hallucinations in its users. It was first

described in a patent by Merck as a by-product and chemical intermediate in the synthesis of

hydrastinine analogues from safrole. The patent for the synthesis of MDMA by Merck stems from

1912. It was not until 1978 that it’s effects were first described in humans (Shulgin & Nichols 1978),

though reports of its appearance on the illicit drug market already appeared date back to 1972

(Gaston & Rasmussen 1972). Besides its use as a club drug, in the mid 1970’s Alexander Shulgin

began distributing MDMA to psychotherapists recognizing the value of MDMA as an adjunct in

psychotherapy (Benzenhöfer & Passie 2010). In the years following the value of MDMA for its use in

psychotherapy became more recognized (e.g. Grinspoon & Bakalar 1986, and more recently White

2014; Mithoefer et al. 2018). Its recreational use began to rise rapidly during the 1980’s as well, with

the drug being mass produced and sold openly in bars in the US in the mid 1980’s (Pentney 2001).

Because of its increasing use and concerns about the safety of MDMA, the DEA decided on an

emergency ban on MDMA in the US in 1985. In 1987 the ban was lifted to support research into

possible benefits of its use in psychotherapy, but this was reversed in 1988, citing lack of evidence

regarding the beneficial effects as an adjunct in psychotherapy and lack of toxicological studies as

among their main reasons (Lawn 1988). Despite the ban on MDMA, use of MDMA has spread with

the Global Drug Survey 2019 reporting that MDMA is the second most popular used illicit drug after

cannabis worldwide (Winstock et al. 2019). The widespread use of MDMA has prompted researchers

to gain insight into the possibly damaging effects of MDMA. Most research seems to indicate MDMA

is neurotoxic in humans. However, inconsistent research results, partially caused by the high amount

of confounding factors in MDMA research, make it harder to come to a convincing conclusion with

regards to the neurotoxic effects of MDMA.

Indirect evidence of neurotoxicity: cognitive problems due to MDMA Use of MDMA is not without its risks. After use, users can experience heavy negative side-effects like

anhedonia, lethargy, and depression (Parrott 2014). This could have especially negative

consequences when used for therapeutic purposes. There are also concerns of possible neurotoxicity

as a result of recreational MDMA use. Early research into the possibly damaging effects of MDMA

indicated MDMA users show poorer memory performance compared to controls (Gouzoulis-

Mayfrank et al. 2000; Reneman et al. 2000; Reneman et al. 2001b). As part of the Netherlands XTC

Toxicity study (NeXT study) (de Win et al. 2005) a prospective study on 188 MDMA-naive subjects

likely to use MDMA in the future was performed (Schilt et al. 2007). After a follow up of on average

11 months, 58 subjects had started using MDMA (median use at follow up was 1.5 pill) and were

compared to 60 still MDMA-naive subjects from the same cohort matched for age, sex, intelligence,

and use of other substances. At the follow-up, the researchers found a significant negative effect of

MDMA use on verbal memory compared to controls, with no effect on other cognitive functions. This

study thus provides convincing evidence of a causal relationship between MDMA use and subtle

poorer memory performance, even following low exposure. Rogers and colleagues (2009) provide a

systematic review of 114 research papers that assess the damaging neurocognitive effects of MDMA.

They found evidence for significant lower memory performance of MDMA users compared to

polydrug controls. The highest negative effect was seen on working memory (Standardized Mean

Difference (SMD) –0.391, 95% Confidence Interval (CI) –0.589 to –0.192), delayed verbal memory

(SMD –0.377, 95% CI –0.498 to –0.257) and immediate verbal memory (SMD –0.332, 95% CI –0.451

5

to –0.214). These effects were deemed small by Cohen’s rule of thumb (SMD small = 0.2, medium =

0.5, large = 0.8, Cohen 2013). When comparing MDMA users to drug-naive controls this effect was

even larger, with the largest effects on immediate verbal memory (SMD –0.840, 95% CI –0.990 to –

0.690) and delayed verbal memory (SMD –1.037, 95% CI –1.734 to –0.341). However, no dose-

response relationship is found in most literature between memory problems and MDMA use. Studies

reporting heavier average use of ecstasy did not provide more extreme effect measures than those

consisting of lighter users. In addition, lifetime MDMA use is found not to be predictive for the

severity of memory problems. Although the prospective NeXT study did observe memory problems

as a consequence of MDMA use, the lack of an observed relationship between exposure and

outcome in all other included studies make it difficult to infer any causal relation between MDMA

use and memory problems.

Direct evidence of MDMA induced neurotoxicity The exact cause of the memory deficits caused by MDMA is still to be elucidated. The mechanism of

action of MDMA is briefly described in box 1; the effects of MDMA are mainly on serotonin,

dopamine and norepinephrine. As will be elucidated later on, MDMA appears to be specifically

neurotoxic to serotonergic neurons. This serotonergic neurotoxicity could be responsible for the

observed memory problems in MDMA users, since serotonin is an important modulator in memory

processes (Schmitt et al. 2006). Biomarkers have been developed that specifically target the

serotonergic system in order to quantify serotonergic neurotoxicity. These biomarkers include

density of the serotonin transporter protein (SERT), serotonin receptors, serotonin (5-HT), the

serotonin metabolite 5-hydroxyindoleacetic acid (5-HIAA), and the enzyme responsible for serotonin

synthesis tryptophan-hydroxylase (TPH). Lower density of these markers is used as an indicator for

loss of serotonergic axons/neurons (Green et al. 2003). This section focuses on neurophysiological

evidence for serotonergic neurotoxicity in animals and humans which is assumed to be the cause of

the memory deficits seen in MDMA users.

6

Neurotoxicity of MDMA in animals Research into the neurotoxic effects of MDMA stems back almost four decades, with the first paper

on MDMA induced neurotoxicity by Schmidt and colleagues (1986) reporting lasting depletions in 5-

HT and 5-HIAA in rats after administration of MDMA indicating serotonergic neurotoxicity. Most

animal research on MDMA induced neurotoxicity uses mice, rats, and nonhuman primates. Mice

show dopaminergic neurotoxicity as a consequence of MDMA administration (e.g. Granado et al.

2008), an effect not seen in other animals (Moratalla et al. 2017). In rats and nonhuman primates

MDMA seems to induce specific serotonergic neurotoxicity (Schmidt et al. 1986; Ricaurte et al. 1988)

as evidenced by structural damage to serotonergic nerve fibers. The serotonergic neurotoxicity seen

in animals is accompanied by reductions in brain 5-HT, 5-HIAA, SERT density, 5-HT receptors (5-HT1A

and 5-HT2A), TPH, and a reduction in cerebrospinal fluid 5-HIAA (Green et al. 2003).

There has been some critique on the use of animal studies and the applied allometric scaling in

making conclusions with regards to possible neurotoxic effects of MDMA in humans (de la Torre &

Farré 2004b). Animal studies often use high doses of MDMA (10-20 mg/kg) to correct for the higher

metabolism of these animals. However, this allometric scaling has not been validated for MDMA.

Research indicates animals given an intraperitoneal injection of MDMA at a dose comparable to

humans (2.0 mg/kg) also produced an MDMA Cmax comparable to humans (210 ± 108 ng/mL

compared to 292 ± 76 ng/mL in humans at a dose of 1.6 mg/kg) (Green et al. 2012). In addition,

MDMA produces comparable neurochemical, endocrine, and behavioral actions in rats and humans

at equivalent doses. This indicates the allometric scaling used in most animal research is flawed and

results most likely do not give accurate information with regards to the neurotoxic effects of MDMA

in humans. However, serotonergic neurotoxicity has even been observed in squirrel monkeys using a

dose regimen that is comparable to human users (Mechan et al. 2006). In conclusion, animal studies

clearly indicate MDMA can be toxic to serotonergic neurons, but it is unclear whether these results

are translatable to humans.

Neurotoxicity of MDMA in humans In order to assess the neurotoxic profile of MDMA in humans, some biomarkers used in animals

cannot be used due to their invasive nature. SERT density has been used most frequently as a

biomarker in non-invasive neuroimaging studies using PET and SPECT imaging to detect serotonergic

neurotoxicity caused by MDMA use. The reduction in SERT is thought to signify loss of serotonergic

synapses/axons and possibly loss of serotonergic neurons, since SERT is a structural component of

the serotonergic axon. SERT reduction has been validated as a measure for serotonergic damage in

animals (de Win et al. 2004). Early research indeed found a reduction in SERT density in MDMA users

compared to controls (McCann et al. 1998; Reneman et al. 2001a), providing preliminary evidence

that the serotonergic neurotoxicity induced by MDMA in animals also occurs in humans. In a recent

meta-analysis on 10 imaging studies on SERT density in MDMA users, a significant reduction in SERT

density in MDMA users compared to controls was also observed (Müller et al. 2019). This reduction

in SERT density was apparent in 8 of the 13 studied brain areas, specifically the anterior cingulate,

posterior cingulate, hippocampus, occipital lobe, parietal lobe, temporal lobe, and the thalamus. No

significant reductions were observed in the caudate, frontal lobe, insula, midbrain, and putamen.

Strikingly, the authors found no relation between lifetime MDMA use and severity of SERT reduction.

This is in accordance to the lack of dose-response relationship between observed memory problems

and MDMA use. However, recovery of SERT density has been observed in humans in several studies

(e.g. Buchert et al. 2006; Selvaraj et al. 2009; Erritzoe et al. 2011) with Erritzoe and colleagues

7

estimating SERT density would recover fully after about 200 days of abstinence. Reneman and

colleagues (2001b) found that while abstinent female MDMA users showed recovery of SERT density,

memory problems in this group persisted. Since the memory problems observed in MDMA users

seem permanent and are even observed to worsen after abstinence, the validity of SERT density in

estimating MDMA induced neurotoxicity has been questioned. It is possible the reduction in SERT

signifies an adaptive response as a consequence of serotonin depletion without loss of serotonergic

axons. Another hypothesis explaining the recovery in SERT density after MDMA use is related to the

ability of serotonergic axons to undergo regenerative sprouting after MDMA use. Indeed, recovery of

serotonergic axons is observed in animals after neurotoxic doses of MDMA (Fischer et al. 1995;

Hatzidimitriou et al. 1999). However, the observed reinnervation patterns of these axons appear

abnormal, possibly explaining the lack of recovery of memory function.

Other imaging studies have focused on density of the 5-HT2A (serotonin) receptor as a measure of

MDMA induced neurotoxicity. Research finds a decrease in 5-HT2A receptor binding a short period

after use, followed by increased 5-HT2A receptor binding after abstinence which is thought to be a

compensatory mechanism as a result of serotonin depletion (Reneman et al. 2000; Reneman et al.

2002b; Di Iorio et al. 2012; Urban et al. 2012). Interestingly, Reneman and colleagues (2000) found a

correlation between 5-HT2A density and severity of memory problems. Additionally, Di Iorio and

colleagues (2012) find a relationship between lifetime MDMA use and 5-HT2A density. Further

research is warranted because of the relatively low sample sizes used in the aforementioned studies,

but 5-HT2A density appears to be a promising biomarker in assessing MDMA induced neurotoxicity

and these studies seem to indicate serotonergic neurotoxicity as a consequence of MDMA use

happens in humans as well.

Thesis aim There is some ambiguity with regards MDMA induced serotonergic neurotoxicity in humans.

Decades of animal work clearly indicate dose-dependent serotonergic neurotoxicity induced by

MDMA. In line with this body of animal work, in humans, reduced SERT densities have been observed

in addition to effects on verbal and working memory. Intriguingly, and in contrast to the animal work,

the effects of MDMA on SERT density and memory function are not related to cumulative MDMA

consumption. The discrepancy between the effects of MDMA observed in animal research and in

human research might be caused by other factors that influence an individual’s susceptibility to the

neurotoxic effects of MDMA. Insight into these modulating factors in MDMA induced neurotoxicity

can help improve research assessing the neurotoxic effects of MDMA, especially improving accuracy

of research by correcting for important modulating factors. Potentially, this information can also be

used to improve harm-reduction practices for MDMA users in both a recreational and a therapeutic

setting.

A lot of modulating factors exist that obscure research into MDMA induced neurotoxicity. This

problem is exacerbated further by a lack of uniformity with regards to the variables for which study

groups are matched (Rogers et al. 2009). Known factors influencing MDMA induced neurotoxicity

include age, sex, ethnicity, diet, genes, polydrug use, and environment. This thesis aims to assess the

impact of some of the most important modulating factors based on earlier research and from a

mechanistic point of view. As will be elucidated on in a later section, the neurotoxic effects of MDMA

appear to be caused by its metabolites rather than MDMA itself. Therefore, this thesis will focus on

the metabolism of MDMA, as well as factors like genetic polymorphisms, ethnicity, and sex that

8

modulate activity of the enzymes involved in the metabolism of MDMA. In addition, MDMA users are

often polydrug users and use MDMA at dance parties with high ambient temperature; both

modulating factors in MDMA induced neurotoxicity.

9

Metabolism of MDMA The neurotoxic effects of MDMA depend on its metabolism which is mainly executed by the liver.

Phase 1 metabolism of MDMA is primarily executed by cytochrome p450 (CYP) enzymes and

catechol-O-methyltransferase (COMT) (de la Torre et al 2004a). MDMA can be O-demethylated to

dihydroxymethamphetamine (HHMA) (see figure 1-B) or N-demethylated to MDA (figure 1-A).

CYP2D6 is mainly responsible for O-demethylation and CYP2B6 is mainly responsible for N-

demethylation, though other enzymes are involved in these reactions as well (figures 2-A and 2-B). O-

demethylation happens one order of magnitude faster than N-deamination (Kreth et al. 2000) hence

HHMA is formed in higher quantities than MDA. MDA can also be O-demethylated to its respective

metabolite dihydroxyamphetamine (HMA) (figure 1-B). After demethylation, COMT catalyzes

methylation of HHMA and HHA to hydroxymethoxymethamphetamine (HMMA) and

hydroxymethoxyamphetamine (HMA) respectively (figure 1-C). The methylated compounds will

undergo phase 2 metabolism resulting in their respective sulphate or glucuronide conjugates before

being excreted by the kidneys. In an alternative route, HHMA and HHA can be oxidized to their

respective ortho-quinones (Hiramatsu et al. 1990). These quinone intermediates can form

glutathione (GSH) adducts (figure 1-D) which are the primary candidates for causing neurotoxicity

associated with MDMA use, as will be elucidated on in the next section.

Figure 1, Metabolism of MDMA to MDA (A), subsequent demethylation (B) and methylation (C) of metabolites, and formation of glutathionyl conjugates of HHMA and HHA via their quinone intermediates (D). Based on de la Torre & Farré 2004b, Meyer et al. 2008.

In a minor metabolic pathway, MDMA and its metabolites MDA, HHMA, HMA, HMMA, and HMA can

undergo deamination followed by conjugation to glycine before being excreted. Additionally, HHMA

and HMA can be ring-hydroxylated to 2,4,5-trihydroxymethamphetamine (THMA) 2,4,5-

trihydroxyamphetamine (THA) respectively, though this also constitutes a minor pathway in

metabolism of MDMA.

10

Causes of MDMA neurotoxicity MDMA itself does not appear to be neurotoxic as evidenced by the fact that MDMA administered

directly into the dorsal raphe or median raphe of Sprague-Dawley rats does not produce reductions

in 5-HT and its metabolite 5-HIAA (Paris & Cunningham 1992). Also, MDMA perfusion into the

hippocampus of Dark Agouti rats in concentrations matching those reached by systemic MDMA

administration (5, 10, and 15 mg/kg) did not cause long-term decreases in brain 5-HT, 5-HIAA, and

DA content (Esteban et al. 2001). Research indicates MDMA needs to be bio-activated by hepatic

metabolism for neurotoxicity to occur (Capela et al. 2009). Direct administration of HHA to the rat

brain does not produce neurotoxicity (McCann et al. 1991), therefore it is thought peripherally

formed glutathione conjugates of HHMA and HHA, (5-(GSH)-HHMA, 2,5-bis-(GSH)-HHMA, 5-(GSH)-

HHA, and 2,5-bis-(GSH)-HHA (see figure 1), cause neurotoxicity. Indeed, HHMA and HHA glutathione

conjugates have been proven to be neurotoxic in Wistar rat cortical neuronal cultures (Capela et al.

2007). Direct administration of 2,5-bis-(GSH)-HHMA to Sprague-Dawley rats caused long-term 5-HT

depletion indicating lasting serotonergic damage (Miller et al. 1997). Peripherally formed glutathione

adducts of MDMA metabolites have also been observed to be present in the rat brain after MDMA

administration (Jones et al. 2005). Inside the neuron, quinones like HHMA and HHA conjugated to

glutathione undergo redox-cycling which produces a disproportionate amount of reactive oxygen

species (ROS) and reactive nitrogen species (RNS) (Monks & Lau 1997). High amounts of oxidative

stress damages cellular components and will ultimately result in cell death (Droge 2002). High levels

of ROS and RNS have indeed been implicated in neurotoxicity caused by MDMA (Colado et al. 1997).

Inside the brain these glutathione adducts are metabolized to their corresponding N-acetylcysteine

adducts which have been proven to be selective serotonergic neurotoxins as well (Bai et al. 1999).

The N-acetylcysteine HHMA and HMA conjugates have been identified in human urine providing

supporting evidence this form of neurotoxicity also happens in humans (Perfetti et al. 2009). In

conclusion, MDMA is metabolized by the liver in various steps to its thioether metabolites (figure 1-

D). These metabolites are transported into the brain where they undergo redox cycling. As a result,

oxidative stress within the neuron increases dramatically causing damage and ultimately cell death.

Another possible mechanism of MDMA neurotoxicity has been postulated by Sprague and colleagues

(1998). MDMA causes the release serotonin resulting in subsequent serotonin depletion. MDMA also

activates the 5-HT2A receptor which results in increased dopamine synthesis and release. Dopamine is

then taken up into the serotonergic neuron by SERT where it is metabolized by monoamine oxidase B

(MAO-B). This metabolism in turn causes the formation of ROS and RNS which leads to cellular

damage and eventually cell death. In support of this hypothesis, Hrometz and colleagues (2004)

found that human choriocarcinoma cells exposed to MDMA and dopamine exhibited strongly

reduced cell viability which was blocked by either a SERT inhibitor or an MAO-B inhibitor.

Administration of an MAO-B inhibitor to Wistar rats reversed MDMA induced increases in oxidative

stress markers (Alves et al. 2007). Also, attenuated serotonergic neurotoxicity is observed in MAO-B

knockout mice and MAO-B knockdown rats (Fornai et al. 2001; Falk et al. 2002).

Lastly, it has been hypothesized increased oxidation of neurotransmitters within the presynaptic

neuron adds to MDMA induced neurotoxicity (Capela et al. 2009). Also, minor metabolites of

MDMA, THMA and THA, appear to be neurotoxic in the rat brain as well (Elayan et al. 1993).

However, these metabolites are unlikely to be involved in MDMA mediated neurotoxicity since these

metabolites also confer dopaminergic toxicity (Elayan et al. 1992) not seen with MDMA.

11

In conclusion, two main mechanisms appear to be involved in MDMA induced neurotoxicity. MDMA

is converted to the metabolites HHMA and HHA. These metabolites are conjugated to glutathione

and subsequently taken up into the serotonergic neuron. Inside the neuron these metabolites

undergo redox cycling producing a disproportionate amount of ROS. In addition, serotonin depletion

enables dopamine to be taken up into the serotonergic neuron. MAO-B catalyzes dopamine

metabolism inside the serotonergic neuron which again produces ROS. The high level of oxidative

stress caused by increased ROS production leads to damage to serotonergic neurons and ultimately

cell death.

12

Genetic susceptibility Figure 1 gives an overview of the enzymes involved in the metabolism of MDMA. The activity of

these enzymes can be highly variable between individuals. Since the enzymes are involved in MDMA

induced neurotoxicity, individual differences in enzyme activity are an important modulating factor in

MDMA induced neurotoxicity. This section gives an overview of the factors involved in determining

enzymatic activity and the implications of these differences in activity.

CYP2D6 The hepatic CYP2D6 enzyme is responsible for phase 1 metabolism of a large variety of

pharmaceutical and recreational drugs. A large number of polymorphisms of the CYP2D6 gene exist

with a total of 74 allele variants known to date (Zhou 2009). The allele variants can cause loss of

function, reduced function, normal function or increased function, though these effects can be

substrate-dependent. The genetic polymorphisms result in four discernable phenotypes: a poor

metabolizer (PM), an intermediate metabolizer (IM), an extensive metabolizer (EM), and an intensive

metabolizer (IM) phenotype. In general, most people possess the EM phenotype; for example 70-

80% of Caucasians possess the EM phenotype (Sachse et al. 1997). Polymorphism of the CYP2D6

gene is the main determinant of CYP2D6 functions, with factors such as diet, age, and sex playing a

relatively minor role (Steiner et al. 1985). Therefore, polymorphism of the CYP2D6 gene can have a

big influence on the pharmacokinetics of a drug and therefore also its toxicity.

Since CYP2D6 is one of the main enzymes responsible for MDMA metabolism, different CYP2D6

phenotypes have a big influence on the pharmacokinetics of MDMA. For example, subjects with the

PM phenotype had a higher Cmax (maximum concentration) and AUC (area under the curve) of MDMA

after MDMA administration compared to subjects with the EM phenotype (Steuer et al. 2016). Also,

PM subjects had a lower Cmax and AUC of MDMA’s metabolites HHMA and HMMA. Another study

found people with the CYP2D6 *1/*1 genotype (two functional alleles resulting in an EM phenotype)

have a higher Cmax and AUC of MDMA’s main metabolites HHMA and HMMA than people with the

CYP2D6 *4/*4 genotype (two non-functional alleles) (de la Torre et al. 2005). Also, inhibition of

CYP2D6 by either paroxetine or bupropion in people with an EM phenotype results in a PM

phenotype and subsequent higher MDMA Cmax and AUC and a lower Cmax and AUC for HHMA, HMMA,

and HMA (Segura et al. 2005; Steuer et al. 2016).

The neurotoxic effects of MDMA are probably dependent on bio-activation, most importantly by

CYP2D6. Higher CYP2D6 activity leads to a higher Cmax and AUC of HHMA and HMA. Since these

metabolites are implicated in MDMA induced neurotoxicity via their thioether conjugates, high

CYP2D6 activity could lead to a higher susceptibility to the neurotoxic effects of MDMA. Indeed, a

high CYP2D6 activity phenotype seems to be related to a lower performance on verbal fluency tests

(Cuyás et al. 2011) indicating possibly increased neurotoxic effects of MDMA in this group.

Conversely, since the CYP2D6 PM phenotype leads to a lower amount of toxic metabolites being

formed this phenotype might be protective against MDMA induced neurotoxicity (Perfetti et al.

2009).

De la Torre and colleagues found that MDMA has non-linear pharmacokinetics, probably via auto-

inhibition of CYP2D6. This inhibition might be caused by MDMA forming a complex with CYP2D6

rendering it inactive. Recovery of CYP2D6 enzymatic activity is slow (~280 hours) and depends on its

de novo synthesis. Yang and colleagues (2006) found that MDMA could inactivate most hepatic

13

CYP2D6 within an hour. Also, repeated MDMA administration causes a disproportionately high

MDMA Cmax and AUC and a disproportionately low HMMA and HMA Cmax and AUC as expected by

simple dose accumulation (Farré et al. 2004; Farré et al. 2015). This clearly indicates MDMA inhibits

its own metabolism after repeated doses in humans. Because of the ability of MDMA to inhibit its

own metabolism, de la Torre and colleagues (2012) hypothesized that regardless of genetic

polymorphism, with a high enough dose all MDMA users will exhibit a functional CYP2D6 PM

phenotype. As a result, CYP2D6 polymorphism is a genetic modulating factor in the susceptibility to

MDMA induced neurotoxicity in low to moderate doses, but at high doses this effect is abolished due

to auto-inhibition of metabolism by MDMA.

Other cytochrome p450 enzymes Although CYP2D6 activity is strongly correlated to the generation of neurotoxic MDMA metabolites,

the enzyme is responsible for about 30% of total metabolism of MDMA (Segura et al. 2005). As can

be seen in figure 1, multiple other CYP enzymes are involved in the bio-activation of MDMA. CYP1A2,

CYP2B6, CYP3A4, and CYP2C19 all have significant MDMA O-demethylation activity (see figure 1). In

addition, CYP1A2 and CYP2B6 also exhibit significant MDMA N-demethylation activity. This section

gives a brief overview of factors influencing activity of the aforementioned enzymes.

No polymorphisms of CYP1A2 have been identified that cause reduced function. However, one

single-nucleotide polymorphism (SNP) at rs762551 has been identified: a transversion form C to A

(Sachse et al. 1999). The CYP1A2 rs762551 A/A polymorphism appears to be inducible by smoking.

Vizeli and colleagues (2017) found that smokers who possessed this polymorphism had a higher MDA

Cmax after MDMA administration, without any influence on the concentration of MDMA and HMMA

or the subjective effects. Also, Yubero-Lahoz and colleagues (2012) found that CYP1A2 activity was

raised after MDMA administration. Unfortunately they did not look at the genetic polymorphism of

CYP1A2 in their subjects.

A multitude of allele variants and SNP’s of CYP2B6 have been identified making this one of the most

polymorphic CYP genes in humans (Zanger et al. 2007). Subjects with the CYP2B6 rs3745274 T/T

genotype causing reduced activity had a higher MDMA Cmax, though the plasma levels of MDA and

HMMA were unaffected. Interestingly, the effects of CYP2B6 genotype on MDMA concentration

appeared later in time. Vizeli and colleagues hypothesized this could indicate that MDMA

metabolism by CYP2B6 might become more important after CYP2D6 has been inactivated by MDMA.

As with CYP2D6, CYP3A4 activity is mainly influenced by genes (Özdemir et al. 2000; Lee & Goldstein

2005). Genetic variants in CYP3A4 causing amino-acid sequence changes in the CYP3A4 enzyme are

rare and most polymorphisms influence expression of the enzyme (Wang et al. 2011). CYP3A4

activity has been correlated to the generation of neurotoxic metabolites of MDMA (Antolino-Lobo et

al. 2010) and could therefore be an important factor in MDMA mediated neurotoxicity. However,

CYP3A4 activity gets inhibited by MDMA (Yubero Lahoz et al. 2011; Jamshidfar et al. 2017) and

inhibition of CYP3A4 does not completely prevent the neurotoxic effects of MDMA (Antolino-Lobo et

al. 2011). Therefore the impact of CYP3A4 activity on MDMA induced neurotoxicity seems relatively

small.

Multiple allele variants of CYP2C19 have been identified that cause loss of function or gain of

function (Desta et al. 2002; Hicks et al. 2013). Vizeli and colleagues (2017) found that reduced activity

14

of CYP2C19 caused a higher Cmax of MDA after MDMA administration. However, lowered CYP2C19

activity did not influence the concentration of MDMA and HMMA or the subjective effects of MDMA.

It appears genetics have a strong influence on CYP2B6, CYP3A4, and CYP2C19 activity. In contrast,

only one SNP has been identified in CYP1A2 that can influence its activity. Differences in activity of

the aforementioned enzymes seem to have a mild effect on MDMA metabolism compared to

differences in CYP2D6 activity. Lowered CYP2C19 and CYP3A4 activity, and higher CYP1A2 activity

could shift the metabolic route of MDMA from O-demethylation (figure 1-B) to N-demethylation

(figure 1-A) causing higher formation of MDA. MDA appears to be more neurotoxic than MDMA

(O’Hearn et al. 1988) and so the aforementioned changes in enzymatic activity could result in more

neurotoxicity. However, lower CYP2C19 activity and higher CYP1A2 activity did not influence HMMA

plasma concentrations so the influence of altered activity of these enzymes on the metabolism of

MDMA seems comparatively small. In conclusion, activity of the aforementioned enzymes can have a

small influence on MDMA induced neurotoxicity but CYP2D6 activity is a far more important

modulating factor in susceptibility to the neurotoxic effects of MDMA.

Catechol-O-Methyltransferase Catechol-O-Methyltransferase (COMT) catalyzes the conversion of HHMA and HHA to HMMA and

HMA respectively. Since HHMA and HHA can be converted to neurotoxic thioether metabolites,

COMT activity could be protective against MDMA induced neurotoxicity. In this case, COMT will

catalyze the conversion of HHMA and HHA to HMMA and HMA before HHMA and HHA can oxidize

and get converted to neurotoxic metabolites. Two COMT isoforms are known: the short soluble S-

COMT isoform found throughout the body and the long membrane bound MB-COMT mostly found in

the brain. One important functional polymorphism is the Val158Met polymorphism which results in a

40% decrease of enzyme activity in vivo (Chen et al. 2004). It is therefore expected that people

possessing one or two Val158Met alleles will exhibit slower conversion of HHMA and HHA to HMMA

and HMA respectively. Lower COMT activity could decrease the rate at which HHMA and HHA are

converted to HMMA and HMA and increase the rate of formation of neurotoxic metabolites. Also,

COMT is involved in breaking down dopamine (Weinshilboum et al. 1999) and higher dopamine

levels might be involved in MDMA induced serotonergic neurotoxicity (Sprague et al. 1998).

Therefore, lower COMT activity might result in higher dopamine levels even more exacerbating

MDMA induced neurotoxicity.

Antolino-Lobo and colleagues (2010) found that MDMA caused cytotoxicity in cells expressing both

CYP2D6 and COMT. This cytotoxicity was significantly greater when cells were exposed to both

MDMA and COMT inhibitor, providing evidence for the protective effect of COMT activity on MDMA

induced toxicity. Pharmacological inhibition of COMT also exacerbated MDMA induced neurotoxicity

in rats as determined by brain 5-HT and 5-HIAA levels (Herndon et al. 2014). Also, COMT val/met and

COMT met/met (possessing either one or two Val158Met alleles respectively) mice were more

sensitive to MDMA-induced neurotoxicity compared to COMT val/val mice. The hypothesis that the

COMT Val158Met polymorphism leads to increased production of neurotoxic thioether metabolites

is substantiated further by the fact that after administration of MDMA, the recovery of the toxic

MDMA metabolite N-Ac-5-Cys-HHMA is two times higher in human subjects possessing the COMT

met/met polymorphism compared to subjects possessing the COMT val/val polymorphism (Perfetti

et al. 2009). This difference was not significant however (p < 0.1), possibly owing to a low sample size

(n = 14). Lastly, COMT met-allele carriers seem more susceptible to the sustained negative effects of

15

MDMA on verbal learning compared to COMT val-carriers (Schilt et al. 2009). Because of the auto-

inhibiting effect of MDMA on CYP2D6 and other cytochrome p450 enzymes, it has been hypothesized

that at higher or repeated doses often seen in users COMT polymorphism plays a more important

role in MDMA mediated neurotoxicity than cytochrome p450 enzymes (Perfetti et al. 2009). In

conclusion, people possessing one or two COMT Val158Met alleles have an increased risk to MDMA

induced neurotoxicity.

Monoamine Oxidase-B Dopamine influx into serotonergic neurons might be an additional mechanism of MDMA induced

neurotoxicity. After MDMA induces serotonin depletion, dopamine is taken up by the serotonergic

neuron where it is metabolized by MAO-B (Sprague et al. 1998). Several polymorphisms of the MAO-

B gene exists, the most common being a G/A substitution on intron 13. Since this SNP is located on

an intron, no change in MAO-B activity has been observed as a result of this polymorphism (Pivac et

al. 2007). On the other hand, the polymorphism does seem to influence expression of MAO-B

(Jakubauskiene et al. 2012). The MAO-B ‘A’ allele enhances splicing and thereby expression of MAO-B

protein. Since MAO-B is involved in MDMA induced neurotoxicity, and MAO-B inhibitors completely

abolish these neurotoxic effects, it seems likely that genetic polymorphisms of MAO-B modulate the

neurotoxic effects of MDMA. However, to date no human research has looked at MAO-B

polymorphism as a modulating factor in MDMA induced neurotoxicity so this remains speculation.

In conclusion, there is clear evidence CYP2D6 and COMT polymorphisms play an important

modulating role in MDMA induced neurotoxicity. CYP1A2, CYP2B6, CYP3A4, and CYP2C19

polymorphisms are assumed to have a relatively low influence on MDMA induced neurotoxicity since

altered activity of these enzymes have a small impact on MDMA pharmacokinetics. Lastly, there are

indications MAO-B polymorphisms could influence MDMA induced neurotoxicity, though no human

research has been done to date to support this assumption.

Sex Research into gender-related differences in susceptibility to the neurotoxic effects of MDMA is

scarce. Female rats seem to be more vulnerable to the memory impairment effects of MDMA than

male rats (Asl et al. 2015). In humans, gender-related differences in serotonergic markers have also

been observed as a result of MDMA use. Female MDMA users show a greater reduction in SERT

binding compared to males (Reneman et al. 2001; Buchert et al. 2004). McCann and colleagues

(1994) also found a greater reduction in CSF 5-HIAA and homovanillic acid (HVA), both markers of

serotonergic function, in female MDMA users compared to male users. However, no gender

differences have been observed in memory function between male and female MDMA users

(Reneman et al. 2006). In contrast, Bolla and colleagues (1998) found that males had greater memory

impairment than females after abstinence from MDMA. This could be caused by a greater ability to

recover from MDMA induced neurotoxicity in women compared to men. Reneman and colleagues

(2001) found that after abstinence SERT density in women did not differ significantly from control.

This recovery of SERT density was not observed in men. The negative effect on memory by MDMA

use also seems to reverse in women but not in men (Price et al. 2014). The greater recovery seen in

women might be caused by the effects of female hormones on the serotonergic system. For

example, the female estrogen estradiol-17β has a positive modulatory effect on SERT expression

(McQueen et al. 1997). A clear difference in susceptibility to the neurotoxic effects of MDMA

between sexes has not yet been found. On the other hand, since the damage from MDMA stems

16

mostly from its metabolism, differences in metabolism could cause a different risk profile for women

compared to men.

Research into sex-related differences in CYP2D6 activity has been ambiguous. For example, Hägg and

colleagues (2001) found that females have a slightly lower baseline CYP2D6 activity compared to

men while Llerena and colleagues (1996) found no differences between genders. These results are

not surprising considering variability of CYP2D6 activitty is mostly determined by genetics (Steiner et

al. 1985). Some research has suggested that women have a slightly lower CYP2D6 baseline activity

for MDMA metabolism compared to men (Yubero-Lahoz et al. 2011). Pardo-Lozano and colleagues

(2012) did not find discernable differences in pharmacokinetic parameters between men and

women. Kolbrich and colleagues (2008) found a higher Cmax and AUC for MDMA and MDA after

MDMA administration in women compared to men. This effect was only significant at the lower dose

MDMA (1mg/kg) and lost significance at the higher dose MDMA (1.6 mg/kg). This result was

reproduced by Hartman and colleagues (2014) and it was hypothesized women experience CYP2D6

auto inhibition by MDMA at a lower dose than men. Indeed, Yubero-Lahoz and colleagues (2011)

found that when dosing men and women (all CYP2D6 EM phenotype) 1.5 mg/kg MDMA, 100% of

women exhibited a PM functional phenotype, while 67% of men exhibited a PM functional

phenotype. In conclusion, CYP2D6 baseline activity appears to be marginally lower for women than

men. However, MDMA is able to inhibit CYP2D6 in women at a much lower dose than in men.

Because of this, it seems plausible that women are more protected for the neurotoxic effects of

MDMA, at least with regards to CYP2D6 mediated toxicity and at lower doses of MDMA.

Sex also seems to influence activity of other enzymes implicated in MDMA induced neurotoxicity.

Estrogen is a regulator of COMT activity and COMT activity have been found to be lower in females in

blood (Weinshilboum et al. 1999) and postmortem brain tissue (Chen et al. 2004). This lowered

activity was independent of COMT polymorphisms. Snell and colleagues (2002) found that females

have a higher MAO-B platelet activity compared to males. Platelet MAO-B activity has been

determined to be a reliable indicator of brain MAO-B activity (Bench et al. 1991). Since a lower COMT

activity and a higher MAO-B activity seem to confer an increased risk in the emergence of MDMA

induced neurotoxicity, it seems plausible women are more susceptible to the damaging effects of

MDMA than men, at least at higher doses of MDMA when both men and women exhibit a CYP2D6

PM phenotype.

Current research into sex-related differences in susceptibility to the neurotoxic effects of MDMA

remains inconclusive. However, there is reason to assume women are more susceptible when

looking at the mechanism behind MDMA induced neurotoxicity. There does not appear to be a big

difference in CYP2D6 function between men and women. However, women undergo CYP2D6 auto

inhibition by MDMA at a lower dose than men. Since MDMA users frequently take higher doses or

take redoses, it is expected most users, men and women, exhibit a functional PM phenotype during

use. Therefore COMT and MAO-B activity could be bigger contributors to MDMA induced

neurotoxicity. COMT activity seems lower and MAO-B activity higher in women, both contributing

factors into the development of neurotoxicity. The difference in enzyme activity seen between men

and women could render women more susceptible to the neurotoxic effects of MDMA. This would

explain the higher reduction in serotonergic markers observed in women MDMA users compared to

male users.

17

In conclusion, with regards to CYP2D6 activity women seem less susceptible to the neurotoxic effects

of MDMA, while lower COMT and higher MAO-B activity seen in women result in a higher

susceptibility to the neurotoxic effects. Since most users exhibit a functional CYP2D6 PM phenotype,

in most situations women are supposed to be more susceptible to the neurotoxic effects of MDMA

than men.

Ethnicity Ethnicity can have an influence on the frequency of certain polymorphisms, including the ones

discussed in this paper. For example CYP2D6 polymorphism frequency seems to be dependent on

ethnicity (Zhou 2009). In Caucasians the frequency of the CYP2D6 PM, IM, EM, and UM phenotypes is

5-10%, 10-17%, 70-80%, and 3-5% respectively (Sachse et al. 1997). The PM phenotype is almost

absent in the Asian population; PM frequency for Chinese, Korean, and Japanese people is 0%, 0.5%,

and 0.7% respectively (Sohn et al. 1991). An unusually high amount of the CYP2D6 UM phenotype

has been found in black Ethiopians: ~16% (Aklillu et al. 1996). In Caucasians the frequency of non-

functional and reduced function alleles is 29%, while almost 50% in Africans and African Americans

(Bradford 2002). There is some evidence these ethnic differences in CYP2D6 polymorphism

frequencies impact MDMA pharmacokinetics. After administration of MDMA, African Americans

have a higher MDMA Cmax and AUC (Kolbrich et al. 2008) and a lower HMMA AUC (Hartman et al.

2014). This data indicates reduced MDMA metabolism by CYP2D6 in African-Americans, possibly

reducing the risk of MDMA induced neurotoxicity.

Differences in COMT Val158Met polymorphism frequency between ethnic groups have also been

observed. In Caucasians, Southwest Asians, and Turkish population the frequency of COMT val/val,

COMT val/met, and COMT met/met have been found to be 25%, 50%, and 25% respectively (McLeod

et al. 1998; Kocabaş et al. 2001). The frequency of the COMT met/met polymorphism seems lower

however in Asian and African populations. For example 6% of Japanese people, 10% of Chinese

people, and 9% of Kenyan people possess the COMT met/met polymorphism (Kunugi et al. 1997; Li

et al 1997; McLeod et al. 1998). Since the COMT val/met and COMT met/met polymorphisms

resulting in less active COMT are associated with increased neurotoxicity, Asian and African people

seem more protected against MDMA induced neurotoxicity because of the low COMT met/met

frequency in these ethnic groups.

In conclusion, the previously mentioned research indicates that African American people have a

lowered risk of MDMA induced neurotoxicity compared to Caucasian people. African Americans have

a higher frequency of non-functional and reduced function CYP2D6 alleles and a lower frequency of

the COMT met/met polymorphism. This would result in lower CYP2D6 activity and higher COMT

activity, both being protective against MDMA induced neurotoxicity. For the Asian group the risk is

more difficult to determine. This group has a low frequency of the CYP2D6 PM polymorphism, but

also a lower frequency of the COMT met/met polymorphism. The higher average CYP2D6 activity

confers a higher risk to MDMA induced neurotoxicity while the lower frequency of the COMT

met/met polymorphism could be protective against MDMA induced neurotoxicity. Because higher

doses of MDMA result in a CYP2D6 PM phenotype regardless of genetic polymorphism, it is assumed

Asians are less susceptible to the neurotoxic effects of MDMA, at least at higher doses frequently

consumed by MDMA users.

18

Environmental Another component that seems to be involved in MDMA mediated neurotoxicity is hyperthermia.

MDMA is often used at crowded dance parties with a relatively high ambient temperature. Combined

with high physical exertion this makes users especially susceptible to the hyperthermic effects of

MDMA. The hyperthermia induced by MDMA seems to be primarily caused by an increase in

catecholamines and thyroid activation (Capela et al. 2009). For example, the D1 antagonist SCH 23390

completely reverses MDMA induced hyperthermia in rats (Mechan et al. 2002). However, 5-HT2A

antagonists have also been observed to prevent MDMA induced hyperthermia in rats and rabbits

(Blessing et al. 2003; Shioda et al. 2008). 5-HT2A mediated hyperthermia could be due to an increase

in dopamine release after 5-HT2A agonism. In addition, activation of spinal 5-HT2A receptors increases

cutaneous vasoconstriction limiting heat loss via the skin. The increase in catecholamines,

vasoconstriction, physical exertion, and high ambient temperature all contribute to MDMA induced

hyperthermia.

A strong correlation between changes in body temperature as a result of MDMA use and severity of

neurotoxicity has been found in animals (Malberg & Seiden 1998). For example, low ambient

temperatures (10°C - 11°C) appear to cause hypothermia in rats administered MDMA which results in

a lack of observed neurotoxicity (Drafters 1994; Broening et al. 1995). In contrast, high ambient

temperatures (>24°C) appear to exacerbate MDMA induced neurotoxicity (Sanchez et al. 2004).

Though hyperthermia is involved in MDMA mediated neurotoxicity, the effect seems to be additive

instead of causative. This is evidenced by the fact that compounds that protect against MDMA

mediated neurotoxicity like memantine and fluoxetine do not influence MDMA induced

hyperthermia (Chipana et al. 2008; Sanchez et al. 2001). The primary mechanism of MDMA induced

neurotoxicity involves its reactive metabolites resulting in an increase in ROS. Hyperthermia

observed with MDMA use most likely results in an additional increase in oxidative stress exacerbating

neurotoxicity (Colado et al. 1998; Colado et al. 1999).

In a controlled environment, hyperthermia as a result of MDMA use has not been observed in

humans at relatively low doses (1-1.7 mg/kg, Grob et al. 1995; Mas et al. 1999; Kolbrich et al. 2008).

However, at a higher dose of 2 mg/kg Freedman and colleagues (2005) observed a 0.4°C increase in

body temperature at 18°C ambient temperature (p < 0.02) and a 0.6°C increase in body temperature

at 30°C ambient temperature (p < 0.001). Interestingly, there was no significant difference between

the temperature increases at low and high ambient temperature. This could mean that club-goers

who consume MDMA are not necessarily more susceptible to the hyperthermic effects of MDMA due

to high ambient temperature. However, most research finds that the increase in body temperature

of MDMA users at dance parties is higher (> 1°C) than that of MDMA users in controlled

environments (Parrott 2012). In addition, Parrott and colleagues (2006) found that MDMA users who

reported to be dancing ‘all the time’ while on MDMA had higher Prospective Memory Questionnaire

long-term problem scores, indicating possible neurotoxic effects, compared to users who danced

sometimes or frequently.

In conclusion, it seems that the combination of high ambient temperature combined with high

physical exertion can exacerbate MDMA induced hyperthermia and possibly MDMA induced

neurotoxicity.

19

Drug-drug interactions MDMA users are often polydrug users, with a UK survey finding most MDMA users also use

amphetamines, cannabis, cocaine, and LSD (Winstock et al. 2001). Some of these drugs have been

proven to be neurotoxic. As a result, polydrug use obscures research into MDMA induced

neurotoxicity. In addition, besides being neurotoxic on their own, some drugs might exacerbate

MDMA induced neurotoxicity when consumed concurrently with MDMA (Gouzoulis-Mayfrank &

Daumann 2006). Other drugs have even been found to be protective against MDMA induced

neurotoxicity. This section gives an overview of drugs commonly used concurrently with MDMA and

known interactions between these drugs.

Caffeine Even though caffeine is generally considered safe for consumption, there are indications that

combined use of caffeine with MDMA can increase the neurotoxic effects of MDMA. For example,

Downey and colleagues (2010) found that MDMA combined with caffeine caused increased

cytotoxicity in vitro compared to MDMA alone. In rats, caffeine appeared to increase the

hyperthermic effects of MDMA (Vanattou et al. 2010) and this effect seemed independent on

caffeine dose (5 and 10 mg/kg, McNamara et al. 2006). These hyperthermic effects seem to be

mediated by dopamine and adenosine A1 and A2 receptors. Caffeine also exacerbated MDMA

induced 5-HT and 5-HIAA depletion. Camarasa and colleagues (2006) hypothesized caffeine could

increase neurotoxicity by increasing dopamine release and subsequent dopamine oxidation in

serotonergic neurons (Sprague et al. 1998). No human studies have been performed to date into the

interaction between caffeine and MDMA, but it seems plausible the hyperthermic effect of caffeine

can have an additive effect on MDMA induced neurotoxicity.

Tobacco MDMA is often used in combination with tobacco. For example, one study reported 64% of students

who used MDMA used it in combination with tobacco (Barrett et al. 2006). The possible interaction

between nicotine and MDMA is not well understood. Since nicotine increases oxidative stress it

seems likely concurrent use of nicotine with MDMA increases the risk of neurotoxicity compared to

MDMA use alone (Parrot 2003). Indeed, Budzynska and colleagues (2018) found that co-

administration of MDMA and nicotine to Swiss mice decreased total antioxidant status (sum of

extracellular endogenous and food-derived antioxidants, Miller et al. 1993), decreased GPx activity,

and raised malondialdehyde levels compared to MDMA alone. This indicates that co-administration

of nicotine with MDMA increases oxidative stress compared to MDMA administration alone which

could possibly lead to increased neurotoxicity. However, the value of mice research is limited since

MDMA has a different neurotoxic profile in mice, showing mainly dopaminergic damage instead of

serotonergic damage, compared to rats and humans (Granado et al. 2008; Capela et al. 2009). Since

nicotine also releases dopamine it could contribute to dopamine mediated serotonergic

neurotoxicity (Sprague et al. 1998), though the dopamine release caused by nicotine seems

insignificant compared to other drugs (Tsukada et al. 2002). In conclusion, not enough research has

been done to date to elucidate the interaction between nicotine and MDMA, but there is an

indication there could be a harmful interaction between the two compounds.

20

Cannabis Most MDMA users also use cannabis (Winstock et al. 2001). Cannabis use can have negative effects

on attention and memory but these effects appear to be reversible and not indicative of

neurotoxicity (Gouzoulis-Mayfrank & Daumann 2006). Some in vitro research indicates one of the

active compounds in cannabis, Δ9-tetrahydrocannabinol (THC), can be neurotoxic to hippocampal

neurons (Chan et al. 1998) while more recent research indicates THC could actually have

neuroprotective effects on hippocampal neurons (e.g. Gilbert et al. 2007). There is also an indication

that cannabis can protect against MDMA induced neurotoxicity. Morley and colleagues (2004) found

that cannabinoids are able to partially protect against the 5-HT depleting effects of MDMA. This

could be related to the hypothermic effects of cannabis, as well as antioxidant effects of THC and

cannabidiol (CBD) (Hampson et al. 1998). In conclusion, cannabis seems to be protective against

MDMA induced neurotoxicity.

Alcohol Alcohol has been observed to increase the Cmax of MDMA in rats (Hamida et al. 2009), as well as in

humans; a dose of 0.8 g/kg causes a rise in MDMA Cmax of 13% in humans (Hernandez-Lopez et al.

2002). Alcohol attenuates the effects of MDMA on hyperthermia and water retention in humans at a

dose of 2-3 beverages (Dumont et al. 2010). The hypothermic effect of alcohol is also seen in rodents

at 23°C ambient temperature, but this effect is abolished at 32°C ambient temperature (Cassel et al.

2007). Ros-Simó and colleagues (2012) also found that besides its hypothermic effects, alcohol also

protected against MDMA induced neuroinflammation in mice. In contrast, pre-exposing rats to a

binge regimen ethanol for four days increased MDMA induced cognitive deficits and serotonergic

neurotoxicity determined by 5-HT concentration and SERT density (Izco et al. 2007). Also, the ethanol

binge regimen did not cause the hypothermic effect observed with low doses of ethanol. It appears

alcohol might increase MDMA induced neurotoxicity at high repeated doses. However, at low doses

the hypothermic effects of alcohol might have a neuroprotective effect. In addition, consumption of

a few alcoholic drinks could mitigate the risk of MDMA induced hypernatremia, a condition that

could be fatal to MDMA users when combined with excessive water consumption (Kalantar-Zadeh et

al. 2006). In conclusion, alcohol in moderate doses could have a protective effect on MDMA induced

neurotoxicity, while higher repeated doses exacerbate MDMA induced neurotoxicity.

Stimulants Concurrent use of stimulants like amphetamine, methamphetamine, and cocaine with MDMA is high

(Winstock et al. 2001). Evidence suggests these stimulants can be neurotoxic to dopaminergic

neurons (Tung et al. 2017; Moratalla et al. 2017; Pereira et al. 2015). Reneman and colleagues

(2002a) found evidence of dopaminergic neurotoxicity in MDMA users who also used amphetamine,

while subjects who only used MDMA had no signs of dopaminergic neurotoxicity. It is not clear

however if the combined use of MDMA with other stimulants results in a synergistic effect on

neurotoxicity in humans. A synergistic effect between MDMA and stimulants on cytotoxicity in vitro

has been observed (Da Siva et al. 2013). Cocaine appears to have an additive effect on MDMA

induced ROS formation (Peraile et al. 2013). In addition, rats given a combination of

methamphetamine and MDMA showed greater serotonergic and dopaminergic toxicity and

hyperthermia compared to methamphetamine or MDMA alone (Clemens et al. 2004; Clemens et al.

2005). It seems stimulants can exacerbate MDMA induced neurotoxicity via increased hyperthermia.

Also, since MDMA induced neurotoxicity seems to be partially related to excessive dopamine release,

21

the additional dopamine release by stimulant drugs might add further to MDMA induced

neurotoxicity. Even though no human studies have been performed to date that elucidate the

interaction between stimulants and MDMA with regards to neurotoxicity, there is enough reason to

assume the combination exacerbates MDMA induced neurotoxicity.

Psychedelics Most psychedelics (e.g. magic mushrooms, LSD, mescaline) do not appear to possess any neurotoxic

effects (Johnson et al. 2008). However, there is evidence that combined use of MDMA and

psychedelics can exacerbate MDMA induced neurotoxicity. For example, combined administration of

LSD and MDMA to rats resulted in an increase in serotonergic neurotoxicity compared to MDMA

administration alone as measured by brain SERT density (Armstrong et al. 2004). In addition, the

psychedelic drugs DOI and 5-MeO-DMT also exacerbate MDMA induced neurotoxicity (Gudelsky et

al. 1994). Psychedelics mainly work via activation of the 5-HT2A receptor. As mentioned before,

activation of the 5-HT2A receptor appears to be involved in MDMA induced neurotoxicity via

increased dopamine release and increased hyperthermic effects. Indeed, the additive effects of DOI

and 5-MeO-DMT on MDMA induced neurotoxicity seem to be mediated by increased dopamine

release (Gudelsky et al. 1994). In conclusion, even though psychedelics do not appear to be

neurotoxic on their own, combined use with MDMA can exacerbate MDMA induced neurotoxicity via

increased activation of the 5-HT2A receptor.

In conclusion, previous research indicates there are interaction effects between MDMA and other

often used drugs that can result in either an additive or a protective effect on MDMA induced

neurotoxicity. Drugs that appear to be protective against MDMA induced neurotoxicity are cannabis

and low/moderate doses of alcohol. On the other hand, high doses of alcohol, stimulants, and

psychedelics exacerbate MDMA induced neurotoxicity and some research suggests caffeine and

tobacco are able to contribute to the neurotoxic effects of MDMA as well. MDMA is often used in

combination with other drugs making this an important modulating factor in research into the

neurotoxic effects of MDMA (Gouzoulis-Mayfrank et al. 2006). Therefore it is important to determine

the possible interactions between other drugs and MDMA in order to accurately assess the specific

neurotoxic effects of MDMA.

22

Conclusion/recommendations The aim of this thesis was to provide an overview of the modulating factors that can influence an

individual’s susceptibility to the neurotoxic effects of MDMA, in order to better interpret existing as

well as design future research. Most of the neurotoxic effects of MDMA emerge via its metabolites

HHMA and HHA. After conjugation to glutathione these metabolites undergo redox cycling in

serotonergic neurons producing a high amount of ROS and RNS ultimately leading to neurotoxicity. In

addition, there is evidence that increased dopamine influx in serotonergic neurons followed by

metabolism by MAO-B, as well as the hyperthermic effects of MDMA, add to the generation of ROS

and RNS and subsequent neurotoxicity.

The main enzyme involved in the formation of HHMA and HHA, CYP2D6, is highly polymorphic in

humans resulting in multiple different phenotypes. People with the lowest (PM) CYP2D6 activity

phenotype appear to be less vulnerable to the neurotoxic effects of MDMA since the rate of

formation of HHMA and HHA is lower in this group. Following the same logic, people with the highest

CYP2D6 activity phenotype appear to be more vulnerable to the neurotoxic effects of MDMA. High

COMT activity protects against MDMA induced neurotoxicity by converting the neurotoxic

metabolites of MDMA into non-neurotoxic metabolites. Therefore, people possessing the highly

active COMT val/val enzyme polymorphism are less susceptible to the neurotoxic effects than people

possessing the COMT val/met and COMT met/met polymorphism. Lastly, there is an indication that

MAO-B polymorphisms could influence MDMA induced neurotoxicity, though no empirical evidence

exists to date to support this claim.

There is also an indication that susceptibility to the neurotoxic effects of MDMA differs between

ethnic groups and sex. This mostly has to do with differences in frequencies in CYP2D6 and COMT

allele variants between ethnic groups, and differences in expression of CYP2D6 and COMT between

sexes. Because of a relatively higher frequency of low activity CYP2D6 and high activity COMT

phenotypes, African-Americans are less susceptible to the neurotoxic effects of MDMA compared to

Caucasians and Asians. In addition, at higher doses that result in a functional CYP2D6 phenotype,

Asians also appear to be less susceptible to the neurotoxic effects of MDMA due to the lower

frequency of the COMT met/met polymorphism. Women experience auto inhibition of CYP2D6 by

MDMA at lower doses of MDMA. Therefore women appear to be less susceptible to the neurotoxic

effects of MDMA at low doses. At higher doses that result in a functional PM CYP2D6 phenotype, this

difference between sexes is abolished. In this case, women are more susceptible to MDMA induced

neurotoxicity due to lower COMT activity and higher MAO-B activity. However, women also show

higher recovery of neurotoxic markers compared to men.

Ambient temperature does not appear to have a strong effect on MDMA induced hyperthermia.

However, high ambient temperature combined with high physical exertion like dancing leads to a

significant increase in body temperature. Since hyperthermia exacerbates MDMA induced

neurotoxicity, MDMA users who frequent dance parties have an increased risk of neurotoxicity. In

addition to environmental factors, polydrug use also has a significant effect on MDMA induced

neurotoxicity, mostly by altering the hyperthermic effects of MDMA. Cannabis and low doses of

alcohol can be neuroprotective via their hypothermic effects. Cannabis also has an added antioxidant

effect. In contrast, stimulants and psychedelics increase the hyperthermic effects of MDMA and

subsequent neurotoxicity. There is also evidence that tobacco and caffeine contribute to the

neurotoxic effects of MDMA, either via increased oxidative stress or an increase in hyperthermia.

23

An overview of the modulating factors in MDMA induced neurotoxicity that have been discussed in

this thesis is provided in figure 2 below:

Figure 2, discussed modulating factors in MDMA induced neurotoxicity. Neurotoxicity of MDMA is presumed to occur because of buildup of toxic metabolites and dopamine influx into the serotonergic neuron ultimately leading to increased production of ROS. Cytochrome p450 and COMT polymorphisms influence enzyme activity and thereby formation of neurotoxic metabolites. Interethnic differences in polymorphism frequency have been observed for the aforementioned enzymes. Individual differences in expression and activity of these enzymes also influence formation of metabolites, and sex is an important modulator of activity of these enzymes. Hyperthermia has been observed to both increase the rate of formation of neurotoxic metabolites, as well as directly increasing ROS. Polydrug use can both positively and negatively influence hyperthermia, and can cause additional dopamine influx into the serotonergic neuron. Lastly, MAO-B polymorphisms might modulate the rate at which dopamine is metabolized and the level of ROS produced.

The information presented in this thesis can be helpful in improving research that assesses the

neurotoxic effects of MDMA. Insight into the modulating factors of MDMA induced neurotoxicity is

important in order to determine for which variables to correct in research. First of all, CYP2D6 and

COMT allele variants are important modulating factors and genetic testing should be applied to a

subject pool to correct for these factors. Also, the influence of MAO-B polymorphisms on MDMA

induced neurotoxicity should be assessed in future research. Ethnicity and sex are also important

modulating factors; African Americans seem less susceptible to the neurotoxic effects of MDMA than

Caucasians, and women seem more susceptible than men. With regards to correcting for the

hyperthermic effects of high ambient temperature combined with high physical exertion often seen

at dance parties, a method based on research by Parrott and colleagues (2006) might prove valuable.

Parrott and colleagues divided subjects in three groups: those who danced ‘all the time’, ‘frequently’,

or ‘sometimes’ at parties and found a clear difference between these groups with regards to memory

performance. This indicates this method of subdividing subjects might be valuable in correcting for

the hyperthermic effects of high physical exertion. Lastly, concurrent use of MDMA with cannabis

24

and low doses of alcohol can limit the neurotoxic effects of MDMA while stimulant and psychedelic

use can exacerbate the neurotoxic effects of MDMA.

This thesis also provides some directions on improving harm-reduction practices for MDMA users.

Users should first of all take care when attending dance parties by taking enough breaks in between

dancing to cool off. Even though there is evidence low/moderate doses of alcohol and cannabis can

be protective against MDMA induced neurotoxicity, there might be an increased risk of detrimental

cardiovascular effects as a result of this combination. Therefore more research is necessary to assess

the safety of this combination. On the other hand, previous research indicates an interaction effect

between stimulants/psychedelics and MDMA exists which could lead to exacerbated serotonergic

neurotoxicity. Therefore, users should avoid combining MDMA with stimulant or psychedelic drugs.

This thesis has discussed some of the most important modulating factors on MDMA induced

neurotoxicity. However, this thesis does not provide information with regards to possible

interaction/cumulative effects between the described modulating factors. A combination between,

for example, genetic and ethnic modulating factors that increase an individual’s risk to MDMA

induced neurotoxicity might work synergistically in exacerbating the neurotoxic effects of MDMA.

Future research should therefore also aim at elucidating the interaction effects between these

modulating factors.

25

References Aklillu, E., Persson, I., Bertilsson, L., Johansson, I., Rodrigues, F., & Ingelman-Sundberg, M. (1996). Frequent

distribution of ultrarapid metabolizers of debrisoquine in an ethiopian population carrying duplicated and

multiduplicated functional CYP2D6 alleles. Journal of pharmacology and experimental

therapeutics, 278(1), 441-446.

Alves, E., Summavielle, T., Alves, C. J., Gomes-da-Silva, J., Barata, J. C., Fernandes, E., ... & Carvalho, F. (2007).

Monoamine oxidase-B mediates ecstasy-induced neurotoxic effects to adolescent rat brain

mitochondria. Journal of Neuroscience, 27(38), 10203-10210.

Antolino-Lobo, I., Meulenbelt, J., Nijmeijer, S. M., Scherpenisse, P., van den Berg, M., & van Duursen, M. B.

(2010). Differential roles of phase I and phase II enzymes in 3, 4-methylendioxymethamphetamine-

induced cytotoxicity. Drug Metabolism and Disposition, 38(7), 1105-1112.

Antolino-Lobo, I., Meulenbelt, J., van den Berg, M., & van Duursen, M. B. (2011). A mechanistic insight into 3, 4-

methylenedioxymethamphetamine (“ecstasy”)-mediated hepatotoxicity. Veterinary quarterly, 31(4), 193-

205.

Asl, S. S., Mehdizadeh, M., Shahraki, S. H., Artimani, T., & Joghataei, M. T. (2015). Sex differences in MDMA-

induced toxicity in Sprague-Dawley rats. Functional neurology, 30(2), 131.

Armstrong, B. D., Paik, E., Chhith, S., Lelievre, V., Waschek, J. A., & Howard, S. G. (2004). Potentiation of (DL)‐3,

4‐methylenedioxymethamphetamine (MDMA)‐induced toxicity by the serotonin 2A receptior partial

agonist d‐lysergic acid diethylamide (LSD), and the protection of same by the serotonin 2A/2C receptor

antagonist MDL 11,939. Neuroscience Research Communications, 35(2), 83-95.

Bai, F., Lau, S. S., & Monks, T. J. (1999). Glutathione and N-acetylcysteine conjugates of α-methyldopamine

produce serotonergic neurotoxicity: possible role in methylenedioxyamphetamine-mediated

neurotoxicity. Chemical research in toxicology, 12(12), 1150-1157.

Barrett, S. P., Darredeau, C., & Pihl, R. O. (2006). Patterns of simultaneous polysubstance use in drug using

university students. Human Psychopharmacology: Clinical and Experimental, 21(4), 255-263.

Baumann, M. H., Wang, X., & Rothman, R. B. (2007). 3, 4-Methylenedioxymethamphetamine (MDMA)

neurotoxicity in rats: a reappraisal of past and present findings. Psychopharmacology, 189(4), 407-424.

Bench, C. J., Price, G. W., Lammertsma, A. A., Cremer, J. C., Luthra, S. K., Turton, D., ... & Biziere, K. (1991).

Measurement of human cerebral monoamine oxidase type B (MAO-B) activity with positron emission

tomography (PET): a dose ranging study with the reversible inhibitor Ro 19-6327. European journal of

clinical pharmacology, 40(2), 169-173.

Benzenhöfer, U., & Passie, T. (2010). Rediscovering MDMA (ecstasy): the role of the American chemist

Alexander T. Shulgin. Addiction, 105(8), 1355-1361.

Bernschneider-Reif, S., Öxler, F., & Freudenmann, R. W. (2006). The origin of MDMA (‘ecstasy’)–separating the

facts from the myth. Die Pharmazie-An International Journal of Pharmaceutical Sciences, 61(11), 966-972.

Blessing, W. W., Seaman, B., Pedersen, N. P., & Ootsuka, Y. (2003). Clozapine reverses hyperthermia and

sympathetically mediated cutaneous vasoconstriction induced by 3, 4-methylenedioxymethamphetamine

(ecstasy) in rabbits and rats. Journal of Neuroscience, 23(15), 6385-6391.

Bock, K. W., Schrenk, D., Forster, A., Griese, E. U., Mörike, K., Brockmeier, D., & Eichelbaum, M. (1994). The

influence of environmental and genetic factors on CYP2D6, CYP1A2 and UDP-glucuronosyltransferases in

man using sparteine, caffeine, and paracetamol as probes. Pharmacogenetics, 4(4), 209-218.

Bolla, K. I., McCann, U. D., & Ricaurte, G. A. (1998). Memory impairment in abstinent MDMA (" Ecstasy")

users. Neurology, 51(6), 1532-1537.

Bradford, L. D. (2002). CYP2D6 allele frequency in European Caucasians, Asians, Africans and their

descendants. Pharmacogenomics, 3(2), 229-243.

26

Broening, H. W., Bowyer, J. F., & Slikker, W. (1995). Age-dependent sensitivity of rats to the long-term effects

of the serotonergic neurotoxicant (+/-)-3, 4-methylenedioxymethamphetamine (MDMA) correlates with

the magnitude of the MDMA-induced thermal response. Journal of Pharmacology and Experimental

Therapeutics, 275(1), 325-333.

Buchert, R., Thomasius, R., Wilke, F., Petersen, K., Nebeling, B., Obrocki, J., ... & Clausen, M. (2004). A voxel-

based PET investigation of the long-term effects of “Ecstasy” consumption on brain serotonin

transporters. American Journal of Psychiatry, 161(7), 1181-1189.

Buchert, R., Thomasius, R., Petersen, K., Wilke, F., Obrocki, J., Nebeling, B., ... & Clausen, M. (2006). Reversibility

of ecstasy-induced reduction in serotonin transporter availability in polydrug ecstasy users. European

journal of nuclear medicine and molecular imaging, 33(2), 188-199.

Budzynska, B., Wnorowski, A., Kaszubska, K., Biala, G., Kruk-Slomka, M., Kurzepa, J., & Boguszewska-Czubara, A.

(2018). Acute MDMA and nicotine co-administration: behavioral effects and oxidative stress processes in

mice. Frontiers in behavioral neuroscience, 12, 149.

Camarasa, J., Pubill, D., & Escubedo, E. (2006). Association of caffeine to MDMA does not increase

antinociception but potentiates adverse effects of this recreational drug. Brain research, 1111(1), 72-82.

Capela, J. P., Macedo, C., Branco, P. S., Ferreira, L. M., Lobo, A. M., Fernandes, E., ... & Carvalho, F. (2007).

Neurotoxicity mechanisms of thioether ecstasy metabolites. Neuroscience, 146(4), 1743-1757.

Capela, J. P., Carmo, H., Remião, F., Bastos, M. L., Meisel, A., & Carvalho, F. (2009). Molecular and cellular

mechanisms of ecstasy-induced neurotoxicity: an overview. Molecular neurobiology, 39(3), 210-271.

Cassel, J. C., Hamida, S. B., & Jones, B. C. (2007). Attenuation of MDMA-induced hyperthermia by ethanol in

rats depends on ambient temperature. European journal of pharmacology, 571(2-3), 152-155.

Chan, G. C. K., Hinds, T. R., Impey, S., & Storm, D. R. (1998). Hippocampal neurotoxicity of Δ9-

tetrahydrocannabinol. Journal of Neuroscience, 18(14), 5322-5332.

Chen, J., Lipska, B. K., Halim, N., Ma, Q. D., Matsumoto, M., Melhem, S., ... & Egan, M. F. (2004). Functional

analysis of genetic variation in catechol-O-methyltransferase (COMT): effects on mRNA, protein, and

enzyme activity in postmortem human brain. The American Journal of Human Genetics, 75(5), 807-821.

Chipana, C., Camarasa, J., Pubill, D., & Escubedo, E. (2008). Memantine prevents MDMA-induced

neurotoxicity. Neurotoxicology, 29(1), 179-183.

Clemens, K. J., Van Nieuwenhuyzen, P. S., Li, K. M., Cornish, J. L., Hunt, G. E., & McGregor, I. S. (2004). MDMA

(“ecstasy”), methamphetamine and their combination: long-term changes in social interaction and

neurochemistry in the rat. Psychopharmacology, 173(3-4), 318-325.

Clemens, K. J., Cornish, J. L., Li, K. M., Hunt, G. E., & McGregor, I. S. (2005). MDMA (‘Ecstasy’) and

methamphetamine combined: order of administration influences hyperthermic and long-term adverse

effects in female rats. Neuropharmacology, 49(2), 195-207.

Cohen, J. (2013). Statistical power analysis for the behavioral sciences. Routledge

Colado, M. I., O'shea, E., Granados, R., Murray, T. K., & Green, A. R. (1997). In vivo evidence for free radical

involvement in the degeneration of rat brain 5‐HT following administration of MDMA (‘ecstasy’) and p‐

chloroamphetamine but not the degeneration following fenfluramine. British journal of

pharmacology, 121(5), 889-900.

Colado, M. I., Granados, R., O'Shea, E., Esteban, B., & Green, A. R. (1998). Role of hyperthermia in the

protective action of clomethiazole against MDMA (‘ecstasy’)‐induced neurodegeneration, comparison

with the novel NMDA channel blocker AR‐R15896AR. British journal of pharmacology, 124(3), 479-484.

Colado, M. I., O’Shea, E., Esteban, B., Granados, R., & Green, A. R. (1999). In vivo evidence against

clomethiazole being neuroprotective against MDMA (‘ecstasy’)-induced degeneration of rat brain 5-HT

nerve terminals by a free radical scavenging mechanism. Neuropharmacology, 38(2), 307-314.

27

Cuyàs, E., Verdejo-García, A., Fagundo, A. B., Khymenets, O., Rodríguez, J., Cuenca, A., ... & Martín-Santos, R.

(2011). The influence of genetic and environmental factors among MDMA users in cognitive

performance. PloS one, 6(11), e27206.

Da Silva, D. D., Carmo, H., & Silva, E. (2013). The risky cocktail: what combination effects can we expect

between ecstasy and other amphetamines?. Archives of toxicology, 87(1), 111-122.

Dafters, R. I. (1994). Effect of ambient temperature on hyperthermia and hyperkinesis induced by 3, 4-

methylenedioxymethamphetamine (MDMA or “ecstasy”) in rats. Psychopharmacology, 114(3), 505-508.

de la Torre, R., Farre, M., Ortuno, J., Mas, M., Brenneisen, R., Roset, P. N., ... & Cami, J. (2000). Non‐linear

pharmacokinetics of MDMA (‘ecstasy’) in humans. British journal of clinical pharmacology, 49(2), 104-109.

de la Torre, R., Farré, M., Roset, P. N., Pizarro, N., Abanades, S., Segura, M., ... & Camí, J. (2004a). Human

pharmacology of MDMA: pharmacokinetics, metabolism, and disposition. Therapeutic drug

monitoring, 26(2), 137-144.

de la Torre, R., & Farré, M. (2004b). Neurotoxicity of MDMA (ecstasy): the limitations of scaling from animals to

humans. Trends in pharmacological sciences, 25(10), 505-508.

de la Torre, R., Farré, M., Mathúna, B. Ó., Roset, P. N., Pizarro, N., Segura, M., ... & Camí, J. (2005). MDMA

(ecstasy) pharmacokinetics in a CYP2D6 poor metaboliser and in nine CYP2D6 extensive

metabolisers. European journal of clinical pharmacology, 61(7), 551-554.

de la Torre, R., Yubero-Lahoz, S., Pardo-Lozano, R., & Farré, M. (2012). MDMA, methamphetamine, and CYP2D6

pharmacogenetics: what is clinically relevant?. Frontiers in genetics, 3, 235.

de Win, M. M., de Jeu, R. A., de Bruin, K., Habraken, J. B., Reneman, L., Booij, J., & den Heeten, G. J. (2004).

Validity of in vivo [123I] β-CIT SPECT in detecting MDMA-induced neurotoxicity in rats. European

neuropsychopharmacology, 14(3), 185-189.

de Win, M. M., Jager, G., Vervaeke, H. K., Schilt, T., Reneman, L., Booij, J., ... & Van Den Brink, W. (2005). The

Netherlands XTC Toxicity (NeXT) study: objectives and methods of a study investigating causality, course,

and clinical relevance. International journal of methods in psychiatric research, 14(4), 167-185.

Desta, Z., Zhao, X., Shin, J. G., & Flockhart, D. A. (2002). Clinical significance of the cytochrome P450 2C19

genetic polymorphism. Clinical pharmacokinetics, 41(12), 913-958.

Di Iorio, C. R., Watkins, T. J., Dietrich, M. S., Cao, A., Blackford, J. U., Rogers, B., ... & Salomon, R. M. (2012).

Evidence for chronically altered serotonin function in the cerebral cortex of female 3, 4-

methylenedioxymethamphetamine polydrug users. Archives of general psychiatry, 69(4), 399-409.

Downey, C., Daly, F., & O’Boyle, K. M. (2014). An in vitro approach to assessing a potential drug interaction

between MDMA (ecstasy) and caffeine. Toxicology in Vitro, 28(2), 231-239.

Droge, W. (2002). Free radicals in the physiological control of cell function. Physiological reviews, 82(1), 47-95.

Dumont, G. J. H., Kramers, C., Sweep, F. C. G. J., Willemsen, J. J., Touw, D. J., Schoemaker, R. C., ... & Verkes, R.

J. (2010). Ethanol co-administration moderates 3, 4-methylenedioxymethamphetamine effects on human

physiology. Journal of psychopharmacology, 24(2), 165-174.

Elayan, I., Gibb, J. W., Hanson, G. R., Foltz, R. L., Lim, H. K., & Johnson, M. (1992). Long-term alteration in the

central monoaminergic systems of the rat by 2, 4, 5-trihydroxyamphetamine but not by 2-hydroxy-4, 5-

methylenedioxymethamphetamine or2-hydroxy-4, 5-methylenedioxyamphetamine. European journal of

pharmacology, 221(2-3), 281-288.

Elayan, I. K. R. A. M., Gibb, J. W., Hanson, G. R., Lim, H. K., Foltz, R. L., & Johnson, M. (1993). Short-term effects

of 2, 4, 5-trihydroxyamphetamine, 2, 4, 5-trihydroxymethamphetamine and 3, 4-

dihydroxymethamphetamine on central tryptophan hydroxylase activity. Journal of Pharmacology and

Experimental Therapeutics, 265(2), 813-818.

28

Erritzoe, D., Frokjaer, V. G., Holst, K. K., Christoffersen, M., Johansen, S. S., Svarer, C., ... & Knudsen, G. M.

(2011). In vivo imaging of cerebral serotonin transporter and serotonin2A receptor binding in 3, 4-

methylenedioxymethamphetamine (MDMA or “ecstasy”) and hallucinogen users. Archives of general

psychiatry, 68(6), 562-576.

Esteban, B., O'Shea, E., Camarero, J., Sanchez, V., Green, A. R., & Colado, M. I. (2001). 3, 4-

Methylenedioxymethamphetamine induces monoamine release, but not toxicity, when administered

centrally at a concentration occurring following a peripherally injected neurotoxic

dose. Psychopharmacology, 154(3), 251-260.

Falk, E. M., Cook, V. J., Nichols, D. E., & Sprague, J. E. (2002). An antisense oligonucleotide targeted at MAO-B

attenuates rat striatal serotonergic neurotoxicity induced by MDMA. Pharmacology Biochemistry and

Behavior, 72(3), 617-622.

Farré, M., de la Torre, R., Mathúna, B. Ó., Roset, P. N., Peiro, A. M., Torrens, M., ... & Cami, J. (2004). Repeated

doses administration of MDMA in humans: pharmacological effects and

pharmacokinetics. Psychopharmacology, 173(3-4), 364-375.

Farré, M., Tomillero, A., Pérez-Mañá, C., Yubero, S., Papaseit, E., Roset, P. N., ... & de la Torre, R. (2015). Human

pharmacology of 3, 4-methylenedioxymethamphetamine (MDMA, ecstasy) after repeated doses taken 4 h

apart Human pharmacology of MDMA after repeated doses taken 4 h apart. European

Neuropsychopharmacology, 25(10), 1637-1649.

Fischer, C., Hatzidimitriou, G., Wlos, J., Katz, J., & Ricaurte, G. (1995). Reorganization of ascending 5-HT axon

projections in animals previously exposed to the recreational drug (+/-) 3, 4-

methylenedioxymethamphetamine (MDMA," ecstasy"). Journal of Neuroscience, 15(8), 5476-5485.

Flockhart DA. Drug Interactions: Cytochrome P450 Drug Interaction Table. Indiana University School of

Medicine (2007). "https://drug-interactions.medicine.iu.edu" Accessed 3-1-2020

Fornai, F., Giorgi, F. S., Gesi, M., Chen, K., Alessrì, M. G., & Shih, J. C. (2001). Biochemical effects of the

monoamine neurotoxins DSP‐4 and MDMA in specific brain regions of MAO‐B‐deficient

mice. Synapse, 39(3), 213-221.

Freedman, R. R., Johanson, C. E., & Tancer, M. E. (2005). Thermoregulatory effects of 3, 4-

methylenedioxymethamphetamine (MDMA) in humans. Psychopharmacology, 183(2), 248-256.

Freyberg, Z., Sonders, M. S., Aguilar, J. I., Hiranita, T., Karam, C. S., Flores, J., ... & Martin, C. A. (2016).

Mechanisms of amphetamine action illuminated through optical monitoring of dopamine synaptic vesicles

in Drosophila brain. Nature communications, 7, 10652.

Gaston, T. R., & Rasmussen, G. T. (1972). Identification of 3, 4-

methylenedioxymethamphetamine. Microgram, 5, 60.

Gilbert, G. L., Kim, H. J., Waataja, J. J., & Thayer, S. A. (2007). Δ9-Tetrahydrocannabinol protects hippocampal

neurons from excitotoxicity. Brain research, 1128, 61-69.

Gouzoulis-Mayfrank, E., Daumann, J., Tuchtenhagen, F., Pelz, S., Becker, S., Kunert, H. J., ... & Sass, H. (2000).

Impaired cognitive performance in drug free users of recreational ecstasy (MDMA). Journal of Neurology,

Neurosurgery & Psychiatry, 68(6), 719-725.

Gouzoulis-Mayfrank, E., & Daumann, J. (2006). The confounding problem of polydrug use in recreational

ecstasy/MDMA users: a brief overview. Journal of Psychopharmacology, 20(2), 188-193.

Granado, N., O’Shea, E., Bove, J., Vila, M., Colado, M. I., & Moratalla, R. (2008). Persistent MDMA‐induced

dopaminergic neurotoxicity in the striatum and substantia nigra of mice. Journal of

neurochemistry, 107(4), 1102-1112.

Green, A. R., Mechan, A. O., Elliott, J. M., O'Shea, E., & Colado, M. I. (2003). The pharmacology and clinical

pharmacology of 3, 4-methylenedioxymethamphetamine (MDMA,“ecstasy”). Pharmacological reviews,

55(3), 463-508.

29

Green, A. R., King, M. V., Shortall, S. E., & Fone, K. C. F. (2012). Lost in translation: preclinical studies on 3, 4‐

methylenedioxymethamphetamine provide information on mechanisms of action, but do not allow

accurate prediction of adverse events in humans. British journal of pharmacology, 166(5), 1523-1536.

Grinspoon, L., & Bakalar, J. B. (1986). Can drugs be used to enhance the psychotherapeutic process?. American

Journal of Psychotherapy, 40(3), 393-404.

Grob, C. S., Poland, R. E., Chang, L., & Ernst, T. (1995). Psychobiologic effects of 3, 4-

methylenedioxymethamphetamine in humans: methodological considerations and preliminary

observations. Behavioural brain research, 73(1-2), 103-107.

Gudelsky, G. A., Yamamoto, B. K., & Nash, J. F. (1994). Potentiation of 3, 4-methylenedioxymethamphetamine-

induced dopamine release and serotonin neurotoxicity by 5-HT2 receptor agonists. European journal of

pharmacology, 264(3), 325-330.

Hägg, S., Spigset, O., & Dahlqvist, R. (2001). Influence of gender and oral contraceptives on CYP2D6 and

CYP2C19 activity in healthy volunteers. British journal of clinical pharmacology, 51(2), 169-173.

Hamida, S. B., Tracqui, A., De Vasconcelos, A. P., Szwarc, E., Lazarus, C., Kelche, C., ... & Cassel, J. C. (2009).

Ethanol increases the distribution of MDMA to the rat brain: possible implications in the ethanol-induced

potentiation of the psychostimulant effects of MDMA. International Journal of

Neuropsychopharmacology, 12(6), 749-759.

Hampson, A. J., Grimaldi, M., Axelrod, J., & Wink, D. (1998). Cannabidiol and (−) Δ9-tetrahydrocannabinol are

neuroprotective antioxidants. Proceedings of the National Academy of Sciences, 95(14), 8268-8273.

Han, D. D., & Gu, H. H. (2006). Comparison of the monoamine transporters from human and mouse in their

sensitivities to psychostimulant drugs. BMC pharmacology, 6(1), 6.

Hartman, R. L., Desrosiers, N. A., Barnes, A. J., Yun, K., Scheidweiler, K. B., Kolbrich-Spargo, E. A., ... & Huestis,

M. A. (2014). 3, 4-Methylenedioxymethamphetamine (MDMA) and metabolites disposition in blood and

plasma following controlled oral administration. Analytical and bioanalytical chemistry, 406(2), 587-599.

Hatzidimitriou, G., McCann, U. D., & Ricaurte, G. A. (1999). Altered Serotonin Innervation Patterns in the

Forebrain of Monkeys Treated with () 3, 4-Methylenedioxymethamphetamine Seven Years Previously:

Factors Influencing Abnormal Recovery. The Journal of Neuroscience, 19(12), 5096-5107.

Hernandez-Lopez, C., Farré, M., Roset, P. N., Menoyo, E., Pizarro, N., Ortuno, J., ... & de la Torre, R. (2002). 3, 4-

Methylenedioxymethamphetamine (ecstasy) and alcohol interactions in humans: psychomotor

performance, subjective effects, and pharmacokinetics. Journal of Pharmacology and Experimental

Therapeutics, 300(1), 236-244.

Herndon, J. M., Cholanians, A. B., Lizarraga, L. E., Lau, S. S., & Monks, T. J. (2014). Catechol-O-methyltransferase

and 3, 4-(±)-methylenedioxymethamphetamine toxicity. Toxicological Sciences, 139(1), 162-173.

Hicks, J. K., Swen, J. J., Thorn, C. F., Sangkuhl, K., Kharasch, E. D., Ellingrod, V. L., ... & Stingl, J. C. (2013). Clinical

Pharmacogenetics Implementation Consortium guideline for CYP2D6 and CYP2C19 genotypes and dosing

of tricyclic antidepressants. Clinical Pharmacology & Therapeutics, 93(5), 402-408.

Hiramatsu, M., Kumagai, Y., Unger, S. E., & Cho, A. K. (1990). Metabolism of

methylenedioxymethamphetamine: formation of dihydroxymethamphetamine and a quinone identified

as its glutathione adduct. Journal of Pharmacology and Experimental Therapeutics, 254(2), 521-527.

Hrometz, S. L., Brown, A. W., Nichols, D. E., & Sprague, J. E. (2004). 3, 4-methylenedioxymethamphetamine

(MDMA, ecstasy)-mediated production of hydrogen peroxide in an in vitro model: the role of dopamine,

the serotonin-reuptake transporter, and monoamine oxidase-B. Neuroscience letters, 367(1), 56-59.

Izco, M., Orio, L., O’Shea, E., & Colado, M. I. (2007). Binge ethanol administration enhances the MDMA-induced

long-term 5-HT neurotoxicity in rat brain. Psychopharmacology, 189(4), 459-470.

30

Jakubauskiene, E., Janaviciute, V., Peciuliene, I., Söderkvist, P., & Kanopka, A. (2012). G/A polymorphism in

intronic sequence affects the processing of MAO-B gene in patients with Parkinson disease. FEBS letters,

586(20), 3698-3704.

Jamshidfar, S., Ardakani, Y. H., Lavasani, H., & Rouini, M. (2017). Inhibition of mirtazapine metabolism by

Ecstasy (MDMA) in isolated perfused rat liver model. DARU Journal of Pharmaceutical Sciences, 25(1), 16.

Johnson, M. P., Hoffman, A. J., & Nichols, D. E. (1986). Effects of enantiomers of MDA, MDMA and related

analogues on [3H] serotonin and [3H] dopamine release from superfused rat brain slices. European journal

of pharmacology, 132(2-3), 269-276.

Johnson, M. W., Richards, W. A., & Griffiths, R. R. (2008). Human hallucinogen research: guidelines for

safety. Journal of psychopharmacology, 22(6), 603-620.

Jones, D. C., Duvauchelle, C., Ikegami, A., Olsen, C. M., Lau, S. S., de la Torre, R., & Monks, T. J. (2005).

Serotonergic neurotoxic metabolites of ecstasy identified in rat brain. Journal of Pharmacology and

Experimental Therapeutics, 313(1), 422-431.

Kalantar-Zadeh, K., Nguyen, M. K., Chang, R., & Kurtz, I. (2006). Fatal hyponatremia in a young woman after

ecstasy ingestion. Nature Reviews Nephrology, 2(5), 283.

Kocabaş, N., Karakaya, A., Cholerton, S., & Şardaş, Ş. (2001). Catechol-O-methyltransferase (COMT) genetic

polymorphism in a Turkish population. Archives of toxicology, 75(7), 407-409.

Kolbrich, E. A., Goodwin, R. S., Gorelick, D. A., Hayes, R. J., Stein, E. A., & Huestis, M. A. (2008). Plasma

pharmacokinetics of 3, 4-methylenedioxymethamphetamine after controlled oral administration to young

adults. Therapeutic drug monitoring, 30(3), 320.

Kreth, K. P., Kovar, K. A., Schwab, M., & Zanger, U. M. (2000). Identification of the human cytochromes P450

involved in the oxidative metabolism of “Ecstasy”-related designer drugs. Biochemical

pharmacology, 59(12), 1563-1571.

Kunugi, H., Nanko, S., Ueki, A., Otsuka, E., Hattori, M., Hoda, F., ... & Collier, D. A. (1997). High and low activity

alleles of catechol-O-methyltransferase gene: ethnic difference and possible association with Parkinson's

disease. Neuroscience letters, 221(2-3), 202-204.

Lawn, J. C. (1988). Scheduling of 3, 4-methylenedioxymethamphetamine (MDMA) into Schedule I of the

Controlled Substances Act. Fed. Reg, 53, 5156-5158.

Lee, S. J., & Goldstein, J. A. (2005). Functionally defective or altered CYP3A4 and CYP3A5 single nucleotide

polymorphisms and their detection with genotyping tests.

Lee, L. O., & Prescott, C. A. (2014). Association of the Catechol-O-Methyltransferase (COMT) Val158Met

polymorphism and anxiety-related traits: a meta-analysis. Psychiatric genetics, 24(2), 52.

Llerena, A., Cobaleda, J., Martinez, C., & Benitez, J. (1996). Interethnic differences in drug metabolism:

influence of genetic and environmental factors on debrisoquine hydroxylation phenotype. European

journal of drug metabolism and pharmacokinetics, 21(2), 129-138.

Li, T., Vallada, H., Curtis, D., Arranz, M., Xu, K., Cai, G., ... & Collier, D. A. (1997). Catechol-O-methyltransferase

Val158Met polymorphism: frequency analysis in Han Chinese subjects and allelic association of the low

activity allele with bipolar affective disorder. Pharmacogenetics, 7(5), 349-353.

Malberg, J. E., & Seiden, L. S. (1998). Small changes in ambient temperature cause large changes in 3, 4-

methylenedioxymethamphetamine (MDMA)-induced serotonin neurotoxicity and core body temperature

in the rat. Journal of Neuroscience, 18(13), 5086-5094.

Mas, M., Farré, M., de la Torre, R., Roset, P. N., Ortuño, J., Segura, J., & Camí, J. (1999). Cardiovascular and

neuroendocrine effects and pharmacokinetics of 3, 4-methylenedioxymethamphetamine in

humans. Journal of Pharmacology and Experimental Therapeutics, 290(1), 136-145.

31

McCann, U. D., & Ricaurte, G. A. (1991). Major metabolites of (±) 3, 4-methylenedioxyamphetamine (MDA) do

not mediate its toxic effects on brain serotonin neurons. Brain research, 545(1-2), 279-282.

McCann, U. D., Ridenour, A., Shaham, Y., & Ricaurte, G. A. (1994). Serotonin neurotoxicity after (±) 3, 4-

methylenedioxymethamphetamine (MDMA;“Ecstasy”): a controlled study in

humans. Neuropsychopharmacology, 10(2), 129.

McCann, U. D., Szabo, Z., Scheffel, U., Dannals, R. F., & Ricaurte, G. A. (1998). Positron emission tomographic

evidence of toxic effect of MDMA (“Ecstasy”) on brain serotonin neurons in human beings. The Lancet,

352(9138), 1433-1437.

McLeod, H. L., Syvänen, A. C., Githang'a, J., Indalo, A., Ismail, D., Dewar, K., ... & Sludden, J. (1998). Ethnic

differences in catechol O-methyltransferase pharmacogenetics: frequency of the codon 108/158 low

activity allele is lower in Kenyan than Caucasian or South-west Asian individuals. Pharmacogenetics, 8(3),

195-199.

McNamara, R., Kerans, A., O'Neill, B., & Harkin, A. (2006). Caffeine promotes hyperthermia and serotonergic

loss following co-administration of the substituted amphetamines, MDMA (“Ecstasy”) and MDA

(“Love”). Neuropharmacology, 50(1), 69-80.

McQueen, J. K., Wilson, H., & Fink, G. (1997). Estradiol-17β increase serotonin transporter (SERT) mRNA levels

and the density of SERT-binding sites in female rat brain. Molecular Brain Research, 45(1), 13-23.

Mechan, A. O., Esteban, B., O'Shea, E., Elliott, J. M., Colado, M. I., & Green, A. R. (2002). The pharmacology of

the acute hyperthermic response that follows administration of 3, 4‐methylenedioxymethamphetamine

(MDMA,‘ecstasy’) to rats. British journal of pharmacology, 135(1), 170-180.

Mechan, A., Yuan, J., Hatzidimitriou, G., Irvine, R. J., McCann, U. D., & Ricaurte, G. A. (2006). Pharmacokinetic

profile of single and repeated oral doses of MDMA in squirrel monkeys: relationship to lasting effects on

brain serotonin neurons. Neuropsychopharmacology, 31(2), 339.

Meyer, M. R., Peters, F. T., & Maurer, H. H. (2008). The role of human hepatic cytochrome P450 isozymes in the

metabolism of racemic 3, 4-methylenedioxy-methamphetamine and its enantiomers. Drug Metabolism

and Disposition, 36(11), 2345-2354.

Miller, N. J., Rice-Evans, C., Davies, M. J., Gopinathan, V., & Milner, A. (1993). A novel method for measuring

antioxidant capacity and its application to monitoring the antioxidant status in premature

neonates. Clinical science, 84(4), 407-412.

Miller, R. T., Lau, S. S., & Monks, T. J. (1997). 2, 5-Bis-(glutathion-S-yl)-α-methyldopamine, a putative metabolite

of (±)-3, 4-methylenedioxyamphetamine, decreases brain serotonin concentrations. European journal of

pharmacology, 323(2-3), 173-180.

Miller, G. M. (2011). The emerging role of trace amine‐associated receptor 1 in the functional regulation of

monoamine transporters and dopaminergic activity. Journal of neurochemistry, 116(2), 164-176.

Mithoefer, M. C., Mithoefer, A. T., Feduccia, A. A., Jerome, L., Wagner, M., Wymer, J., ... & Doblin, R. (2018). 3,

4-methylenedioxymethamphetamine (MDMA)-assisted psychotherapy for post-traumatic stress disorder

in military veterans, firefighters, and police officers: a randomised, double-blind, dose-response, phase 2

clinical trial. The Lancet Psychiatry, 5(6), 486-497.

Monks, T. J., & Lau, S. S. (1997). Biological reactivity of polyphenolic− glutathione conjugates. Chemical research

in toxicology, 10(12), 1296-1313.

Moratalla, R., Khairnar, A., Simola, N., Granado, N., García-Montes, J. R., Porceddu, P. F., ... & Morelli, M.

(2017). Amphetamine-related drugs neurotoxicity in humans and in experimental animals: main

mechanisms. Progress in neurobiology, 155, 149-170.

Morley, K. C., Li, K. M., Hunt, G. E., Mallet, P. E., & McGregor, I. S. (2004). Cannabinoids prevent the acute

hyperthermia and partially protect against the 5-HT depleting effects of MDMA (“Ecstasy”) in

rats. Neuropharmacology, 46(7), 954-965.

32

Müller, F., Brändle, R., Liechti, M. E., & Borgwardt, S. (2019). Neuroimaging of chronic MDMA (" ecstasy")

effects: A meta-analysis. Neuroscience and biobehavioral reviews, 96, 10-20.

O'Hearn, E., Battaglia, G., De Souza, E. B., Kuhar, M. J., & Molliver, M. E. (1988). Methylenedioxyamphetamine

(MDA) and methylenedioxymethamphetamine (MDMA) cause selective ablation of serotonergic axon

terminals in forebrain: immunocytochemical evidence for neurotoxicity. Journal of Neuroscience, 8(8),

2788-2803.

Ootsuka, Y., Nalivaiko, E., & Blessing, W. W. (2004). Spinal 5-HT2A receptors regulate cutaneous sympathetic

vasomotor outflow in rabbits and rats; relevance for cutaneous vasoconstriction elicited by MDMA (3, 4-

methylenedioxymethamphetamine,“Ecstasy”) and its reversal by clozapine. Brain research, 1014(1-2), 34-

44.

Özdemir, V., Kalow, W., Tang, B. K., Paterson, A. D., Walker, S. E., Endrenyi, L., & Kashuba, A. D. (2000).

Evaluation of the genetic component of variability in CYP3A4 activity: a repeated drug administration

method. Pharmacogenetics and Genomics, 10(5), 373-388.

Pardo-Lozano, R., Farré, M., Yubero-Lahoz, S., O’Mathúna, B., Torrens, M., Mustata, C., ... & de la Torre, R.

(2012). Clinical pharmacology of 3, 4-methylenedioxymethamphetamine (MDMA,“ecstasy”): the influence

of gender and genetics (CYP2D6, COMT, 5-HTT). PloS one, 7(10), e47599.

Paris, J. M., & Cunningham, K. A. (1992). Lack of serotonin neurotoxicity after intraraphe microinjection of (+)-3,

4-methylenedioxymethamphetamine (MDMA). Brain research bulletin, 28(1), 115-119.

Parrott, A. C., Rodgers, J., Buchanan, T., Ling, J., Heffernan, T., & Scholey, A. B. (2006). Dancing hot on Ecstasy:

physical activity and thermal comfort ratings are associated with the memory and other psychobiological

problems reported by recreational MDMA users. Human Psychopharmacology: Clinical and

Experimental, 21(5), 285-298.

Parrott, A. C. (2012). MDMA and temperature: a review of the thermal effects of ‘Ecstasy’in humans. Drug and

alcohol dependence, 121(1-2), 1-9.

Parrott, A. C. (2014). The potential dangers of using MDMA for psychotherapy. Journal of psychoactive

drugs, 46(1), 37-43.

Partilla, J. S., Dempsey, A. G., Nagpal, A. S., Blough, B. E., Baumann, M. H., & Rothman, R. B. (2006). Interaction

of amphetamines and related compounds at the vesicular monoamine transporter. Journal of

Pharmacology and Experimental Therapeutics, 319(1), 237-246.

Pentney, A. R. (2001). An exploration of the history and controversies surrounding MDMA and MDA. Journal of

psychoactive drugs, 33(3), 213-221.

Peraile, I., Granado, N., Torres, E., Gutiérrez-López, M. D., Moratalla, R., Colado, M. I., & O’Shea, E. (2013).

Cocaine potentiates MDMA-induced oxidative stress but not dopaminergic neurotoxicity in mice:

implications for the pathogenesis of free radical-induced neurodegenerative

disorders. Psychopharmacology, 230(1), 125-135.

Pereira, R. B., Andrade, P. B., & Valentão, P. (2015). A comprehensive view of the neurotoxicity mechanisms of

cocaine and ethanol. Neurotoxicity research, 28(3), 253-267.

Perfetti, X., O'Mathúna, B., Pizarro, N., Cuyàs, E., Khymenets, O., Almeida, B., ... & Farré, M. (2009). Neurotoxic

thioether adducts of 3, 4-methylenedioxymethamphetamine identified in human urine after ecstasy

ingestion. Drug Metabolism and Disposition, 37(7), 1448-1455.

Pivac, N., Knezevic, J., Kozaric-Kovacic, D., Dezeljin, M., Mustapic, M., Rak, D., ... & Muck-Seler, D. (2007).

Monoamine oxidase (MAO) intron 13 polymorphism and platelet MAO-B activity in combat-related

posttraumatic stress disorder. Journal of affective disorders, 103(1-3), 131-138.

Reneman, L., Booij, J., Schmand, B., van den Brink, W., & Gunning, B. (2000). Memory disturbances in” Ecstasy”

users are correlated with an altered brain serotonin neurotransmission. Psychopharmacology, 148(3),

322-324.

33

Reneman, L., Booij, J., de Bruin, K., Reitsma, J. B., de Wolff, F. A., Gunning, W. B., ... & van den Brink, W.

(2001a). Effects of dose, sex, and long-term abstention from use on toxic effects of MDMA (ecstasy) on

brain serotonin neurons. The Lancet, 358(9296), 1864-1869.

Reneman, L., Lavalaye, J., Schmand, B., de Wolff, F. A., van den Brink, W., den Heeten, G. J., & Booij, J. (2001b).

Cortical serotonin transporter density and verbal memory in individuals who stopped using 3, 4-

methylenedioxymethamphetamine (MDMA or ecstasy): preliminary findings. Archives of General

Psychiatry, 58(10), 901-906.

Reneman, L., Booij, J., Lavalaye, J., de Bruin, K., Reitsma, J. B., Gunning, B. W., ... & van den Brink, W. (2002a).

Use of amphetamine by recreational users of ecstasy (MDMA) is associated with reduced striatal

dopamine transporter densities: a [123 I] β-CIT SPECT study–preliminary

report. Psychopharmacology, 159(3), 335-340.

Reneman, L., Endert, E., de Bruin, K., Lavalaye, J., Feenstra, M. G., de Wolff, F. A., & Booij, J. (2002b). The acute

and chronic effects of MDMA (“ecstasy”) on cortical 5-HT 2A receptors in rat and human brain.

Neuropsychopharmacology, 26(3), 387.

Reneman, L., Schilt, T., de Win, M. M., Booij, J., Schmand, B., den Brink, W. V., & Bakker, O. (2006). Memory

function and serotonin transporter promoter gene polymorphism in ecstasy (MDMA) users. Journal of

Psychopharmacology, 20(3), 389-399.

Ricaurte, G. A., Forno, L. S., Wilson, M. A., DeLanney, L. E., Irwin, I., Molliver, M. E., & Langston, J. W. (1988).

(±3, 4-Methylenedioxymethamphetamine Selectively Damages Central Serotonergic Neurons in

Nonhuman Primates. Jama, 260(1), 51-55.

Rogers, G., Elston, J., Garside, R., Roome, C., Taylor, R. S., Younger, P., ... & Somerville, M. (2009). The harmful

health effects of recreational ecstasy: a systematic review of observational evidence.

Ros-Simó, C., Ruiz-Medina, J., & Valverde, O. (2012). Behavioural and neuroinflammatory effects of the

combination of binge ethanol and MDMA in mice. Psychopharmacology, 221(3), 511-525.

Rudnick, G., & Wall, S. C. (1992). The molecular mechanism of" ecstasy"[3, 4-methylenedioxy-

methamphetamine (MDMA)]: serotonin transporters are targets for MDMA-induced serotonin release.

Proceedings of the national academy of sciences, 89(5), 1817-1821.

Sachse, C., Brockmöller, J., Bauer, S., & Roots, I. (1997). Cytochrome P450 2D6 variants in a Caucasian

population: allele frequencies and phenotypic consequences. American journal of human genetics, 60(2),

284.

Sachse, C., Brockmöller, J., Bauer, S., & Roots, I. (1999). Functional significance of a C→ A polymorphism in

intron 1 of the cytochrome P450 CYP1A2 gene tested with caffeine. British journal of clinical

pharmacology, 47(4), 445-449.

Sanchez, V., Camarero, J., Esteban, B., Peter, M. J., Green, A. R., & Colado, M. I. (2001). The mechanisms

involved in the long‐lasting neuroprotective effect of fluoxetine against MDMA (‘ecstasy’)‐induced

degeneration of 5‐HT nerve endings in rat brain. British journal of pharmacology, 134(1), 46-57.

Sanchez, V., O’shea, E., Saadat, K. S., Elliott, J. M., Colado, M. I., & Green, A. R. (2004). Effect of Repeated

(‘Binge’) Dosing of MDMA to Rats Housed at Normal and High Temperature on Neurotoxicdamage to

Cerebral 5-Ht and Dopamine Neurones. Journal of Psychopharmacology, 18(3), 412-416.

Schilt, T., de Win, M. M., Koeter, M., Jager, G., Korf, D. J., van den Brink, W., & Schmand, B. (2007). Cognition in

novice ecstasy users with minimal exposure to other drugs: a prospective cohort study. Archives of

General Psychiatry, 64(6), 728-736.

Schilt, T., Koeter, M. W., de Win, M. M., Zinkstok, J. R., van Amelsvoort, T. A., Schmand, B., & van den Brink, W.

(2009). The effect of Ecstasy on memory is moderated by a functional polymorphism in the cathechol-O-

methyltransferase (COMT) gene. European Neuropsychopharmacology, 19(2), 116-124.

34

Schmidt, C. J., Wu, L., & Lovenberg, W. (1986). Methylenedioxymethamphetamine: a potentially neurotoxic

amphetamine analogue. European journal of pharmacology, 124(1-2), 175-178.

Schmitt, J. A. J., Wingen, M., Ramaekers, J. G., Evers, E. A. T., & Riedel, W. J. (2006). Serotonin and human

cognitive performance. Current pharmaceutical design, 12(20), 2473-2486.

Segura, M., Farré, M., Pichini, S., Peiró, A. M., Roset, P. N., Ramirez, A., ... & de la Torre, R. (2005). Contribution

of cytochrome P450 2D6 to 3, 4-methylenedioxymethamphetamine disposition in humans. Clinical

pharmacokinetics, 44(6), 649-660.

Selvaraj, S., Hoshi, R., Bhagwagar, Z., Murthy, N. V., Hinz, R., Cowen, P., ... & Grasby, P. (2009). Brain serotonin

transporter binding in former users of MDMA (‘ecstasy’). The British Journal of Psychiatry, 194(4), 355-

359.

Shioda, K., Nisijima, K., Yoshino, T., Kuboshima, K., Iwamura, T., Yui, K., & Kato, S. (2008). Risperidone

attenuates and reverses hyperthermia induced by 3, 4-methylenedioxymethamphetamine (MDMA) in

rats. Neurotoxicology, 29(6), 1030-1036.

Shulgin, A. T., & Nichols, D. E. (1978). Characterization of three new psychotomimetics. The

psychopharmacology of hallucinogens, 74-83.

Simmler, L. D., Buchy, D., Chaboz, S., Hoener, M. C., & Liechti, M. E. (2016). In vitro characterization of

psychoactive substances at rat, mouse, and human trace amine-associated receptor 1. Journal of

Pharmacology and Experimental Therapeutics, 357(1), 134-144.

Snell, L. D., Glanz, J., & Tabakoff, B. (2002). Relationships Between Effects of Smoking, Gender, and Alcohol

Dependence on Platelet Monoamine Oxidase‐B: Activity, Affinity Labeling, and Protein

Measurements. Alcoholism: Clinical and Experimental Research, 26(7), 1105-1113.

Sohn, D. R., Shin, S. G., Park, C. W., Kusaka, M., Chiba, K., & Ishizaki, T. (1991). Metoprolol oxidation

polymorphism in a Korean population: comparison with native Japanese and Chinese populations. British

journal of clinical pharmacology, 32(4), 504-507.

Sprague, J. E., Everman, S. L., & Nichols, D. E. (1998). An integrated hypothesis for the serotonergic axonal loss

induced by 3, 4-methylenedioxymethamphetamine. Neurotoxicology, 19(3), 427-441.

Steiner, E., Iselius, L., Alvan, G., Lindsten, J., & Sjöqvist, F. (1985). A family study of genetic and environmental

factors determining polymorphic hydroxylation of debrisoquin. Clinical Pharmacology &

Therapeutics, 38(4), 394-401.

Steuer, A. E., Schmidhauser, C., Tingelhoff, E. H., Schmid, Y., Rickli, A., Kraemer, T., & Liechti, M. E. (2016).

Impact of cytochrome P450 2D6 function on the chiral blood plasma pharmacokinetics of 3, 4-

methylenedioxymethamphetamine (MDMA) and its phase I and II metabolites in humans. PloS one, 11(3),

e0150955.

Sulzer, D., & Rayport, S. (1990). Amphetamine and other psychostimulants reduce pH gradients in midbrain

dopaminergic neurons and chromaffin granules: a mechanism of action. Neuron, 5(6), 797-808.

Tsukada, H., Miyasato, K., Kakiuchi, T., Nishiyama, S., Harada, N., & Domino, E. F. (2002). Comparative effects of

methamphetamine and nicotine on the striatal [11C] raclopride binding in unanesthetized

monkeys. Synapse, 45(4), 207-212.

Tung, C. S., Chang, S. T., Huang, C. L., & Huang, N. K. (2017). The neurotoxic mechanisms of amphetamine: step

by step for striatal dopamine depletion. Neuroscience letters, 639, 185-191.

Urban, N. B., Girgis, R. R., Talbot, P. S., Kegeles, L. S., Xu, X., Frankle, W. G., ... & Laruelle, M. (2012). Sustained

recreational use of ecstasy is associated with altered pre and postsynaptic markers of serotonin

transmission in neocortical areas: a PET study with [11 C] DASB and [11 C] MDL

100907. Neuropsychopharmacology, 37(6), 1465.

35

Vanattou‐Saïfoudine, N., McNamara, R., & Harkin, A. (2010). Mechanisms mediating the ability of caffeine to

influence MDMA (‘Ecstasy’)‐induced hyperthermia in rats. British journal of pharmacology, 160(4), 860-

877.

Vizeli, P., Schmid, Y., Prestin, K., zu Schwabedissen, H. E. M., & Liechti, M. E. (2017). Pharmacogenetics of

ecstasy: CYP1A2, CYP2C19, and CYP2B6 polymorphisms moderate pharmacokinetics of MDMA in healthy

subjects. European Neuropsychopharmacology, 27(3), 232-238.

Wang, D., Guo, Y., Wrighton, S. A., Cooke, G. E., & Sadee, W. (2011). Intronic polymorphism in CYP3A4 affects

hepatic expression and response to statin drugs. The pharmacogenomics journal, 11(4), 274.

Weinshilboum, R. M., Otterness, D. M., & Szumlanski, C. L. (1999). Methylation pharmacogenetics: catechol O-

methyltransferase, thiopurine methyltransferase, and histamine N-methyltransferase. Annual review of

pharmacology and toxicology, 39(1), 19-52.

White, C. M. (2014). 3, 4-Methylenedioxymethamphetamine’s (MDMA’s) impact on posttraumatic stress

disorder. Annals of Pharmacotherapy, 48(7), 908-915.

Winstock, A. R., Griffiths, P., & Stewart, D. (2001). Drugs and the dance music scene: a survey of current drug

use patterns among a sample of dance music enthusiasts in the UK. Drug and alcohol dependence, 64(1),

9-17.

Winstock, A.R., Barratt, M. J., Maier, L.J., Aldridge A., Zhuparris, A., Davies, E., Hughes, C., Johnson, M.,

Kowalski, M., Ferris, J.A. (2019). Global Drug Survey (GDS) 2019 Key Findings Report.

Yang, J., Jamei, M., Heydari, A., Yeo, K. R., de la Torre, R., Farré, M., ... & Rostami-Hodjegan, A. (2006).

Implications of mechanism-based inhibition of CYP2D6 for the pharmacokinetics and toxicity of

MDMA. Journal of Psychopharmacology, 20(6), 842-849.

Yubero-Lahoz, S., Pardo, R., Farré, M., O’Mahony, B., Torrens, M., Mustata, C., ... & de la Torre, R. (2011). Sex

differences in 3, 4-methylenedioxymethamphetamine (MDMA; ecstasy)-induced cytochrome P450 2D6

inhibition in humans. Clinical pharmacokinetics, 50(5), 319-329.

Yubero-Lahoz, S., Pardo, R., Farre, M., Mathuna, B. Ó., Torrens, M., Mustata, C., ... & de la Torre, R. (2012).

Changes in CYP1A2 activity in humans after 3, 4-methylenedioxymethamphetamine (MDMA, ecstasy)

administration using caffeine as a probe drug. Drug metabolism and pharmacokinetics, 27(6), 605-613.

Zanger, U. M., Klein, K., Saussele, T., Blievernicht, J., H Hofmann, M., & Schwab, M. (2007). Polymorphic

CYP2B6: molecular mechanisms and emerging clinical significance.

Zhou, S. F. (2009). Polymorphism of human cytochrome P450 2D6 and its clinical significance. Clinical

pharmacokinetics, 48(12), 761-804.