CBD Cannabidiol wissenschaftliche Informationen

WHO on CBD

Der älteste bisher bekannte Beweis für die Verwendung von Cannabis in China ist mehr als 2500 Jahre alt.

Sehen ~Sie: https://www.sciencemag.org/news/2019/06/oldest-evidence-marijuana-use-discovered-2500-year-old-cemetery-peaks-western-china

Als Hersteller dürfen wir keine Aussagen über den Betrieb eines Produkts machen. Sie können Ihre eigene Forschung auf Pubmed durchführen, wo alle medizinischen Untersuchungen veröffentlicht werden. Lesen Sie den folgenden Artikel darüber, was das Who (Weltgesundheitsorganisation) über Cannabis Diol (CBD-Öl) sagt. Seit 2017 wird Cannabis in immer mehr Ländern veröffentlicht.

Die folgenden Informationen sind öffentliche Informationen von der wer und sagt nichts über unsere Produkte.

Das Originaldokument kann hier heruntergeladen werden: www.who.int/medicines/access/controlled-substances/5.2_CBD.pdf

CANNABIDIOL (CBD)
Pre-Review Report
Agenda Item 5.2
Expert Committee on Drug Dependence
Thirty-ninth Meeting
Geneva, 6-10 November 2017
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 2 of 27
Contents
Acknowledgements ……………………………………………………………………………………………………4
Summary…………………………………………………………………………………………………………………..5
1. Substance identification…………………………………………………………………………………………..6
A. International Nonproprietary Name (INN)………………………………………………………………….6
B. Chemical Abstract Service (CAS) Registry Number ……………………………………………………..6
C. Other Chemical Names …………………………………………………………………………………………….6
D. Trade Names …………………………………………………………………………………………………………..6
E. Street Names……………………………………………………………………………………………………………6
F. Physical Appearance………………………………………………………………………………………………..6
G. WHO Review History ……………………………………………………………………………………………….6
2. Chemistry……………………………………………………………………………………………………………….6
A. Chemical Name ……………………………………………………………………………………………………….6
B. Chemical Structure…………………………………………………………………………………………………..7
C. Stereoisomers………………………………………………………………………………………………………….7
D. Methods and Ease of Illicit Manufacturing………………………………………………………………….7
E. Chemical Properties…………………………………………………………………………………………………9
F. Identification and Analysis………………………………………………………………………………………..9
3. Ease of Convertibility Into Controlled Substances ……………………………………………………10
4. General Pharmacology…………………………………………………………………………………………..11
A. Routes of administration and dosage ………………………………………………………………………..11
B. Pharmacokinetics…………………………………………………………………………………………………..11
C. Pharmacodynamics………………………………………………………………………………………………..12
5. Toxicology…………………………………………………………………………………………………………….13
6. Adverse Reactions in Humans………………………………………………………………………………..13
7. Dependence Potential…………………………………………………………………………………………….14
A. Animal Studies……………………………………………………………………………………………………….14
B. Human Studies……………………………………………………………………………………………………….14
8. Abuse Potential……………………………………………………………………………………………………..14
A. Animal Studies……………………………………………………………………………………………………….14
B. Human Studies……………………………………………………………………………………………………….14
9. Therapeutic Applications and Extent of Therapeutic Use and Epidemiology of Medical
Use……………………………………………………………………………………………………………………….15
10. Listing on the WHO Model List of Essential Medicines…………………………………………….19
11. Marketing Authorizations (as a Medicinal Product) …………………………………………………19
12. Industrial Use ……………………………………………………………………………………………………….19
13. Non-Medical Use, Abuse and Dependence ………………………………………………………………19
14. Nature and Magnitude of Public Health Problems Related to Misuse, Abuse and
Dependence…………………………………………………………………………………………………………..20
15. Licit Production, Consumption and International Trade…………………………………………..20
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 3 of 27
16. Illicit Manufacture and Traffic and Related Information………………………………………….20
17. Current International Controls and Their Impact…………………………………………………….20
18. Current and Past National Controls………………………………………………………………………..20
19. Other Medical and Scientific Matters Relevant for a Recommendation on the Scheduling
of the Substance…………………………………………………………………………………………………….21
References……………………………………………………………………………………………………………….22
Annex 1: Report on WHO Questionnaire for Review of Psychoactive Substances for the
39th ECDD: Evaluation of Cannabidiol ……………………………………………………………………27
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 4 of 27
Acknowledgements
This report has been drafted under the responsibility of the WHO Secretariat, Department of
Essential Medicines and Health Products, Teams of Innovation, Access and Use and Policy,
Governance and Knowledge. The WHO Secretariat would like to thank the following people
for their contribution in producing this review report: Professor Jason White, Adelaide,
Australia (literature search, review and drafting), Ms. Dilkushi Poovendran, Geneva,
Switzerland (questionnaire analysis and report drafting) and Dr. Stephanie Kershaw, Adelaide,
Australia (review report editing, questionnaire analysis and report drafting).
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 5 of 27
Summary
Cannabidiol (CBD) is one of the naturally occurring cannabinoids found in cannabis plants. It
is a 21-carbon terpenophenolic compound which is formed following decarboxylation from a
cannabidiolic acid precursor, although it can also be produced synthetically.
CBD can be converted to tetrahydrocannabinol (THC) under experimental conditions;
however, this does not appear to occur to any significant effect in patients undergoing CBD
treatment.
In experimental models of abuse liability, CBD appears to have little effect on conditioned
place preference or intracranial self-stimulation. In an animal drug discrimination model CBD
failed to substitute for THC. In humans, CBD exhibits no effects indicative of any abuse or
dependence potential.
CBD has been demonstrated as an effective treatment of epilepsy in several clinical trials, with
one pure CBD product (Epidiolex®) currently in Phase III trials. There is also preliminary
evidence that CBD may be a useful treatment for a number of other medical conditions.
There is unsanctioned medical use of CBD based products with oils, supplements, gums, and
high concentration extracts available online for the treatment of many ailments.
CBD is generally well tolerated with a good safety profile. Reported adverse effects may be as
a result of drug-drug interactions between CBD and patients’ existing medications.
Several countries have modified their national controls to accommodate CBD as a medicinal
product.
To date, there is no evidence of recreational use of CBD or any public health related problems
associated with the use of pure CBD.
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 6 of 27
1. Substance identification
A. International Nonproprietary Name (INN)
Cannabidiol
B. Chemical Abstract Service (CAS) Registry Number
13956-29-1 [1]
C. Other Chemical Names
CBD;
2-[1R-3-methyl-6R-(1-methylethenyl)-2-cyclohexen-1-yl]-5-pentyl-1,3-
benzenediol; [2]
D. Trade Names
Epidiolex® (in development)
Arvisol® (in development)
E. Street Names
No data available
F. Physical Appearance
A crystalline solid [2]
G. WHO Review History
Cannabidiol has not been previously pre-reviewed or critically reviewed by
the WHO Expert Committee on Drug Dependence (ECDD). The current
review is based on the recommendation from the 38th ECDD that pre-review
documentation on cannabis-related substances, including cannabidiol, be
prepared and evaluated at a subsequent committee meeting[3].
2. Chemistry
A. Chemical Name
IUPAC Name: 2-[(6R)-3-methyl-6-prop-1-en-2-ylcyclohex-2-en-1-
yl]-5-pentylbenzene-1,3-diol
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 7 of 27
B. Chemical Structure
Molecular Formula: C21H30O2
Molecular Weight: 314.469 g/mol
C. Stereoisomers
Cannabidiol (CBD) is normally taken to refer to the naturally occurring (-)-
enantiomer. (+) CBD has been synthesised [4], but has received little attention.
(+) CBD has been shown to have modest affinity at CB1 and CB2 receptors
unlike (-) CBD ((+)-CBD Ki= 0.84 µM at CB1), whereas both compounds
inhibited anandamide hydrolysis and were agonists at the vanilloid type 1
(VR1) receptor at which capsaicin acts. [5] The (+)-CBD isomer was more
active than the (-)-CBD-isomer as an anticonvulsant agent in a mouse seizure
model. [6] However, to date, there is no substantive evidence as to whether
(+)-CBD is likely to cause THC-like psychoactive effects.
D. Methods and Ease of Illicit Manufacturing
Synthesis of CBD in vitro:
Synthetic routes are available for the production of CBD, but some of the
published methods yield only small amounts of CBD. The two most efficient
routes are:
1) The condensation of (+)-e-mentha-diene-l-01 with olivetol in the presence
of weak acids (oxalic, picric or maleic acid). The isomer obtained in this
reaction may be converted to CBD with BF3-etherate by a retro-FriedelCrafts
reaction, followed by recombination. However, with this reagent the
reaction proceeds further causing cyclisation of CBD to delta-1-THC and
iso-THC [7]
2) A one step reaction for CBD synthesis utilizes boron trifluoride (BF3)-
etherate on alumina as condensing reagent in the reaction of (+)-e-menthadiene-l-01
with olivetol on a 0.8mmol scale (refer to Figure 1). This results
in CBD as the major product, with 55% yield as chromatographically pure
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 8 of 27
oil or 41% yield as crystalline material. On a 100mmol scale, the yields
were 46% as an oil, and 37% as crystalline material. [8]
Figure 1: Synthesis of CBD with boron trifluoride (BF3)-etherate taken
from Mechoulam et al 2002 [9]
Synthesis of CBD in plants:
Cannabis cultivars range from those grown to produce cannabis for
recreational purposes to those produced in order to use hemp fibre derived
from the stems of the plant. In cultivars utilized for recreational purposes, the
quantity of THC exceeds that of CBD in the dried female inflorescences used
for smoking and oral administration. Hemp cultivars produce substantially less
THC and higher levels of CBD. [10] Unsanctioned production of cannabis
cultivars with high CBD levels does occur for purposes of medical treatment
rather than recreational use (refer to Section 13).
In plants, THC and CBD are derived from their acidic precursors Δ9

tetrahydrocannabinolic acid (THCA) and cannabidiolic acid (CBDA) (refer to
Figure 2). THCA and CBDA are both derived from cannabigerolic acid
(CBGA). The final step differs, with THCA synthase and CBDA synthase
producing THCA or CBDA, respectively, from CBGA. Subsequent
decarboxylation of THCA and CBDA via light exposure, heating, or aging,
results in THC or CBD.[10-12]
CBD (major product)
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 9 of 27
Figure 2: Biogenesis of THC and CBD adapted from Taura et al. (2007)
THCA synthase and CBDA synthase catalyze oxidative cyclization of the
monoterpene moiety of CBGA to form THCA and CBDA, respectively. THC
and CBD are generated from THCA and CBDA by non‐enzymatic
decarboxylation. [11]
In addition to genetic characteristics, cultivated plants are influenced by
environmental conditions and production technology during their life cycle. A
study evaluating the effects of ambient temperature and humidity, soil
temperature and precipitation on the content of THC and CBD in industrial
hemp noted that these agroclimatic conditions have differing effects on THC
and CBD. For example, CBD content is positively affected by soil temperature
and ambient temperature, but negatively influenced by precipitation [13]
E. Chemical Properties
Melting point: 62-63°C
Solubility: approx. 23.6 mg/mL in DSMO and ethanol [14]
F. Identification and Analysis
There are a number of published methods for the analytical detection of CBD
in various biological samples. For example,
▪ spectrophotometric determination [15];
▪ liquid chromatography–tandem mass spectrometry (LC–MS/MS)
detection of CBD in whole blood [16, 17] samples;
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 10 of 27
▪ high performance (HP) LC-MS/MS methods for CBD detection in hair
[18], urine [19] and plasma [20] samples;
▪ gas chromatography mass spectrometry (GC-MS) detection of CBD in
hair [21, 22], oral [23] and plasma [24] samples;
▪ 2-dimensional-GC-MS methods for detection in oral fluid [25], plasma
[26] and post mortem blood samples [27].
3. Ease of Convertibility Into Controlled Substances
There is some evidence that CBD can be converted to tetrahydrocannabinol (THC), a
Schedule 1 substance under the United Nations Convention on Psychotropic
Substances 1971. Two main methods have been reported and there has been some
investigation into whether this occurs spontaneously in vivo.
Conversion in the laboratory
Under experimental conditions, it has been demonstrated that heating CBD in
solutions of some acids catalyses cyclizations within the CBD molecule resulting in
delta-9-THC [28]. Gaoni and Mechoulam have published several papers regarding
methods of converting CBD to other cannabinoids including THC, however the yields
vary and purity is unclear. [9]
A version of this method has been reported on a drug user forum. It suggests
dissolving CBD in sulphuric acid/acetic acid and leaving it for anywhere from 3 hours
to 3 days to obtain delta-9-THC and delta-8-THC. After 3 hours, the author states that
CBD has been converted into 52% delta-9-THC and 2% delta-8-THC [31].
A patent (US 2004/0143126 A1) on the conversion of CBD to delta-9-THC details a
method involving the addition of BF3Et2O (50 µl), under nitrogen atmosphere, to an
ice cold solution of CBD (300 mg) in dry methylene chloride (15 ml). The solution is
stirred at 0° C for 1 hour, followed by the addition of saturated aqueous solution of
NaHCO3 (2 ml) until the red colour fades. The organic layer is removed, washed with
water, dried over MgSO4 and evaporated. The composition of the oil obtained
(determined by HPLC) is: trans-delta8-isoTHC 27%, delta-9-THC 66.7%. The oil is
then chromatographed on silica gel column (20 g) and eluted with petroleum ether
followed by graded mixtures, up to 2:98 of ether in petroleum ether. The first fraction
eluted was the delta8-isoTHC (30 mg, 9.5%) followed by a mixture of delta8-iso THC
and delta-9-THC (100 mg). The last compound to be eluted was the delta-9-THC (172
mg, 57%). The purity of delta-9-THC (as determined by HPLC) is 98.7%. [29]
Spontaneous conversion
It has been proposed that the conversion of CBD to delta-9-THC in the presence of
acid could occur in the human gut. Such conversion could be of importance if CBD is
administered orally. Two in vitro studies have used simulated gastric fluid to
demonstrate the potential for this conversion. The first reported the formation of
analytically confirmed delta-9-THC and delta-8-THC when CBD was exposed to
simulated gastric fluid without enzymes at 37ºC. The authors concluded that that the
acidic environment during normal gastrointestinal transit could expose orally CBD-
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 11 of 27
treated patients to levels of THC and other psychoactive cannabinoids that may
exceed the threshold for a physiological response. [30] The second in vitro study also
reported the formation of delta-9-THC along with other cannabinoid products in
artificial gastric juice without pepsin. The conversion rate of CBD to THC was only
2.9%. [31]
The predictive value of these in vitro studies for humans administering cannabidiol
orally has been questioned as simulated gastric fluid does not exactly replicate
physiological conditions in the stomach. [32] Furthermore, spontaneous conversion of
CBD to delta-9-THC has not been demonstrated in humans undergoing CBD
treatment. For example, in a six week clinical study in Huntington’s disease patients
who were administered CBD 700 mg/day, the CBD average plasma concentration
range was 5.9-11.2 ng/mL with no delta-9-THC detected. [33]
In humans, THC effects are characterised by impairment of psychomotor and
cognitive performance, and a range of physical effects including increased heart rate
and dry mouth. In general, clinical studies have reported that even high doses of oral
CBD do not cause the those effects that are characteristic for THC and for cannabis
rich in THC.[34] For example, in a study of healthy volunteers administered 200mg
oral CBD, CBD did not produce any impairments of motor or psychomotor
performance.[35] A number of other studies involving high doses of CBD were
recently summarized by Grotenhermen et al.[34]; they concluded that high doses of
oral CBD consistently fail to demonstrate significant effects or demonstrate effects
opposite to those of THC.
While it has been suggested that further large-scale human studies are needed to
explore the gastric conversion and potential THC-like side effects following oral CBD
administration [36], it is very unlikely that oral cannabidiol will be shown to result in
THC concentrations sufficient to induce any meaningful effects.
4. General Pharmacology
A. Routes of administration and dosage
Currently there are no approved marketed pure CBD medicinal products,
although two are in development (refer to Section 11).
In clinical trials and research studies, CBD is generally administered orally as
either a capsule, or dissolved in an oil solution (e.g. olive or sesame oil). It can
also be administered through sublingual or intranasal routes. A wide range of
oral doses have been reported in the literature, with most from 100-
800mg/day. [37]
B. Pharmacokinetics
Oral delivery of an oil-based capsule formulation of CBD has been assessed in
humans. Probably due to its poor aqueous solubility, the absorption of CBD
from the gastrointestinal tract is erratic, and the resulting pharmacokinetic
profile is variable. Bioavailability from oral delivery was estimated to be 6%
due to significant first-pass metabolism.[38] In healthy male volunteers, the
mean±SD whole blood levels of CBD at 1, 2 and 3 hours after administration
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 12 of 27
of 600mg oral CBD were reported to be 0.36 (0.64) ng/mL, 1.62 (2.98) ng/mL
and 3.4 (6.42) ng/mL, respectively. [39] Aerosolized CBD has been reported
to yield rapid peak plasma concentrations in 5–10 minutes and higher
bioavailability than oral administration.
CBD is rapidly distributed into the tissues with a high volume of distribution
of ~32L/kg. Like THC, CBD may preferentially accumulate in adipose tissues
due to its high lipophilicity. [37, 40]
CBD is extensively metabolised in the liver. The primary route is
hydroxylation to 7-OH-CBD which is then metabolised further resulting in a
number of metabolites that are excreted in faeces and urine.[38] A study in
human liver microsomes (HLMs) demonstrated that CBD was metabolized by
pooled HLMs to eight monohydroxylated metabolites (6α-OH-, 6β-OH-,7-
OH-, 1”-OH-, 2”-OH-, 3”-OH-, 4”-OH-, and 5”-OH-CBDs). Among these
metabolites, 6α-OH-, 6β-OH-, 7-OH-, and 4”-OH-CBDs were the major ones.
Seven recombinant human CYP enzymes were identified as capable of
metabolising CBD: CYP1A1, CYP1A2, CYP2C9, CYP2C19, CYP2D6,
CYP3A4, and CYP3A5. The two main isoforms involved are CYP3A4 and
CYP2C19. [41]
In a number of studies, CBD has been shown to inhibit CYP isozymes in vitro,
but it is not clear that this occurs at concentrations achieved with doses used
clinically.
C. Pharmacodynamics
There are two main cannabinoid (CB) receptors, CB1 which is primarily
located in the central nervous system with some expression in peripheral
tissues and CB2 receptors, which can be found in the periphery on cells with
immune function and in the gastrointestinal tract and at low densities in the
central nervous system.
CBD does not appear to act directly at CB1 receptors, with a number of studies
reporting that there is no measurable response in binding assays. In studies
examining potential agonist effects at CB1 receptors, most find no effect, with
one report of a weak agonist and one of a weak antagonist effect, each at high
concentrations (>10uM). CBD also shows low affinity at CB2 receptors. [42]
Across a range of measures in humans and animals, CBD had been shown to
have very different effects from those of THC. In mice, CBD failed to produce
the behavioral characteristics (e.g. suppression of locomotor activity,
hypothermia, antinociception) associated with CB1 activation, whereas THC
generated all of the effects which occur when CB1 is activated. [43, 44]
Neuroimaging studies in humans and animals have shown that CBD has
effects which are generally opposite to those of THC.[45] In contrast to THC,
CBD has no effect on heart rate or blood pressure under normal conditions, but
in animal models of stress it reduces heart rate and blood pressure.[46] Other
differences between THC and CBD are discussed below.
Some studies have shown that CBD may reduce or antagonize some of the
effects of THC. The mechanism for this is unclear, with some suggesting that
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 13 of 27
it may be a weak CB1 antagonist. Recent evidence suggests that it may be a
negative allosteric modulator of the CB1 receptor, thereby acting as a noncompetitive
antagonist of the actions of THC and other CB1 agonists.[42, 47]
CBD may also interact with the endocannabinoid system through indirect
mechanisms such as enhanced action of the endogenous cannabinoid ligand
anandamide. This results from blockade of anandamide reuptake and the
inhibition of its enzymatic degradation. [5, 9, 41]
CBD has been shown to modulate several non-endocannabinoid signaling
systems. It is not clear which, if any, of these mechanisms are responsible for
any of CBD’s potential clinical or other effects. Some of these mechanism
include [48]:
• Inhibition of adenosine uptake, possibly resulting in indirect
agonist activity at adenosine receptors.
• Enhanced activity at the 5-HT1a receptor.
• Enhanced activity at glycine receptor subtypes
• Blockade of the orphan G-protein-coupled receptor GPR55
5. Toxicology
The potential toxic effects of CBD have been extensively reviewed [49] with a recent
update of the literature. [50] In general, CBD has been found to have relatively low
toxicity, although not all potential effects have been explored. The following are some
of the relevant findings to date from in vitro and animal studies:
• CBD affects growth of tumoral cell lines, but has no effect in most nontumour
cells. However, a pro-apoptotic effect has been observed in
lymphocytes.
• It has no effect on embryonic development (limited research)
• Evidence on potential hormonal changes is mixed, with some evidence of
possible effects and other studies suggesting no effect, depending on the
method used and the particular hormone
• It has no effect on a wide range of physiological and biochemical parameters
or significant effects on animal behaviour unless extremely high doses are
administered (eg, in excess of 150 mg/kg iv as an acute dose or in excess of
30 mg/kg orally daily for 90 days in monkeys)
• Effects on the immune system are unclear; there is evidence of immune
suppression at higher concentrations, but immune stimulation may occur at
lower concentrations.
• There is potential for CBD to be associated with drug interactions through
inhibition of some cytochrome P450 enzymes, but it is not yet clear whether
these effects occur at physiological concentrations.
6. Adverse Reactions in Humans
As noted above, CBD does not produce the effects that are typically seen with
cannabinoids such as THC. It also failed to produce significant effects in a human
study of abuse potential discussed below.[39] Across a number of controlled and open
label trials CBD of the potential therapeutic effects of CBD it is generally well
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 14 of 27
tolerated, with a good safety profile. [37, 50] Clinical trials involving use of CBD for
treatment of epilepsy will be discussed in Section 9: Therapeutic Applications.
7. Dependence Potential
A. Animal Studies
Male mice were injected i.p once a day for 14 days with either CBD (0.1, 1, or
3mg/kg) or delta-9-THC (1, 3, or 10mg/kg). Tolerance to the effects of THC
was observed, however no tolerance to CBD at any of the dosages was
observed. [51] No studies of the physical dependence potential of CBD in
animals were identified.
B. Human Studies
Controlled, human studies regarding the potential physical dependence effects
(e.g. withdrawal and tolerance) of cannabidiol have not been reported
8. Abuse Potential
A. Animal Studies
In male Sprague-Dawley rats, administration of low dose (5 mg/kg) CBD did
not change the threshold frequency required for intracranial self-stimulation
(ICSS). However, high dose (10 mg/kg and 20 mg/kg) CBD resulted in an
elevation of the threshold suggestive of diminished reward activity. This effect
is opposite to that of drugs of abuse such as cocaine, methamphetamine and
opioids which lower the threshold.[52]
Increased dopamine release in cells of the mesolimbic ventral tegmental area –
nucleus accumbens pathway is a common effect characteristic of almost all
drugs of abuse. While THC has been shown to increase the firing rate of these
cells, cannabidiol had no effect. [53]
It appears that CBD given alone has little effect on conditioned place
preference (CPP). For example, Long-Evans rats treated with 10 mg/kg CBD
showed neither CPP nor CPA.[54] However, rats treated with increasing doses
of CBD and THC (1, 3, and 10 mg/kg) exhibited a trend towards CPP not seen
in those given THC alone. [55] The authors attributed this to a
pharmacokinetic interaction leading to higher THC concentrations rather than
a change in receptor action.
CBD appears not to exhibit THC-like discriminative stimulus effects. For
example, in rats trained to discriminate THC from vehicle, CBD did not
substitute for THC at any dose tested [54]. CBD also failed to substitute for
THC in pigeons trained to discriminate THC from vehicle. [56]
B. Human Studies
While the number of studies is limited, the evidence from well controlled
human experimental research indicates that CBD is not associated with abuse
potential.
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 15 of 27
Single dose administration of cannabidiol has been evaluated in healthy
volunteers using a variety of tests of abuse potential as well as physiological
effects in a randomised double blind placebo controlled trial.[39] An orally
administered dose of 600mg of CBD did not differ from placebo on the scales
of the Addiction Research Centre Inventory, a 16 item Visual Analogue Mood
Scale, subjective level of intoxication or psychotic symptoms. In contrast,
THC (10mg oral) administration was associated with subjective intoxication
and euphoria as well as changes in ARCI scales reflecting sedation and
hallucinogenic activity. THC also increased psychotic symptoms and anxiety.
While THC increased heart rate, CBD had no physiological effects.
A randomized, double-blind, within-subject laboratory study was undertaken
to assess the influence of CBD (0, 200, 400, 800mg, p.o.) pre-treatment on the
effects of inactive (0.01% THC) and active (5.30–5.80% THC) smoked
cannabis. Healthy cannabis smokers (n=31) completed eight outpatient
sessions with CBD administered 90min prior to cannabis administration.
Under placebo CBD conditions, active cannabis was self-administered by
significantly more participants and produced significant, time-dependent
increases in subjective ratings and heart rate relative to inactive cannabis.
CBD alone produced no significant psychoactive, cardiovascular or other
effects. Cannabis self-administration, subjective effects, and cannabis ratings
did not vary as a function of CBD dose relative to placebo capsules. These
findings suggest that oral CBD does not reduce the reinforcing, physiological,
or positive subjective effects of smoked cannabis.[57]
The authors of the study then undertook a second analysis of this data to
examine the abuse liability profile of oral cannabidiol in comparison to oral
placebo and active smoked cannabis. The results of this analysis demonstrated
that CBD was placebo-like on all measures (including visual analogue scales,
psychomotor performance such as the digit symbol substitution task, heart rate
and blood pressure) compared to active cannabis, which produced abuserelated
subjective effects as well as a range of other effects. [58]
9. Therapeutic Applications and Extent of Therapeutic Use and
Epidemiology of Medical Use
Epilepsy
The clinical use of CBD is most advanced in the treatment of epilepsy. In clinical
trials, CBD has been demonstrated as an effective treatment for at least some forms of
epilepsy, with one pure CBD product (Epidiolex®) currently in Phase III trials.
The use of CBD for this purpose is based on a number of studies in animals dating
back to the 1970s. [59] These studies demonstrated the anti-seizure activity of
cannabidiol in a number of animal models. Based on this research, cannabidiol has
been trialled in patients with epilepsy.
In a very early small-scale double-blind placebo controlled trial, patients received
either 200 mg CBD daily (4 patients) or placebo (5 patients) for a 3-month period, in
addition to their habitual medication. In the CBD group, two patients had no seizures
for the entire 3-month period, one partially improved, and the fourth had no
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 16 of 27
improvement. No improvements were observed in the placebo group and no toxic
effects were reported for either group. This study has a number of limitations,
including the small sample size, unclear design as to blinding, and lack of definition
of partial improvement. [60]
In another study, 15 patients with “secondarily generalized epilepsy with temporal
focus,” were randomly divided into two groups. In a double-blind procedure, each
patient received 200-300 mg daily of CBD or placebo for up to four and a half months
in combination with their existing prescribed antiepileptic medications (which were
no longer effective in the control of their symptoms). CBD was tolerated in all
patients, with no signs of toxicity or serious side effects. Of the eight participants in
the CBD treatment group, four were reported to be almost free of seizure episodes
throughout the trial, whereas three others showed partial clinical improvement. CBD
was ineffective in one patient. In comparison, the clinical condition of seven placebo
patients remained unchanged with one patient showing improvement.[61]
There have also been some negative reports regarding the effectiveness of CBD. In a
trial reported in 1986, a dose of CBD of 200–300 mg/day for a month resulted in no
significant differences between the treatment and placebo groups. [62] Similarly, a 6-
month double blind study administering CBD 100 mg 3 times each day did not result
in any changes in seizure frequency or improvement in cognition or behaviour. [63]
The results of two trials examining the effects of CBD in patients with severe,
intractable, childhood-onset, treatment-resistant epilepsy have been reported. The first
was an open label study of 214 patients (aged 1–30 years) who were receiving stable
doses of antiepileptic drugs before study entry. Patients were given oral cannabidiol,
initially at 2–5 mg/kg per day, and then titrated until intolerance or to a maximum
dose of 25 mg/kg or 50 mg/kg per day, dependent on study site. The primary measure
was the percentage change in the frequency of seizures. In the CBD group, the median
monthly frequency of motor seizures reduced from 30·0 at baseline to 15·8 over the
12 week treatment period. The trial was also designed to assess safety, but the absence
of a control group means that the results cannot be used to assess the likelihood of
CBD producing particular effects. Adverse events reported in more than 10% of
patients were somnolence, decreased appetite, diarrhoea, fatigue, and convulsion. Five
(3%) patients discontinued treatment because of an adverse event. Serious adverse
events were reported in 48 (30%) patients, of which 20 (12%) experienced severe
adverse events possibly related to cannabidiol use, the most common of which was
status epilepticus (n=9 [6%]). [64]
The same research group recently reported the results of a controlled trial of CBD
treatment for Dravet syndrome, a complex childhood epilepsy disorder that is
associated with drug-resistant seizures and a high mortality rate. In a double-blind,
placebo-controlled trial, 120 children and young adults with Dravet syndrome were
randomly assigned to receive either cannabidiol oral solution (20 mg per kilogram per
day) or placebo, in addition to standard antiepileptic treatment (a median of 3.0
drugs). The authors reported that cannabidiol decreased the median frequency of
convulsive seizures per month from 12.4 to 5.9, as compared with a decrease from
14.9 to 14.1 with placebo. A small percentage (5%) of patients in the CBD group
became seizure free as compared to zero in the placebo group. Adverse events that
occurred more frequently in the cannabidiol group than in the placebo group included
diarrhoea (31% vs 10%), loss of appetite (28% vs 5%) and somnolence (36% vs
10%). Other adverse effects noticed were vomiting, fatigue, pyrexia and abnormal
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 17 of 27
results on liver-function tests. Adverse effects led to the withdrawal of eight patients
in the cannabidiol group compared with one in the placebo group.[65]
It has been suggested that some of the adverse effects of cannabidiol observed in the
clinical studies may relate to interactions with other antiepileptic drugs. For example,
a recent study evaluated thirteen subjects with refractory epilepsy concomitantly
taking clobazam and CBD. Nine of 13 subjects had a >50% decrease in seizures,
corresponding to a responder rate of 70%. Side effects were reported in 10 (77%) of
the 13 subjects, but were alleviated with clobazam dose reduction. All subjects
tolerated CBD well. [66]
It has been suggested that cannabidiol (as Epidiolex) is likely to be submitted for
regulatory approval by GW Pharmaceuticals for epilepsy treatment in 2017 following
the successful outcomes reported in treatment of Dravet syndrome.
Other indications
There is also evidence that CBD may be a useful treatment for a number of other
medical conditions. However, this research is considerably less advanced than for
treatment of epilepsy. For most indications, there is only pre-clinical evidence, while
for some there is a combination of pre-clinical and limited clinical evidence. The
range of conditions for which CBD has been assessed is diverse, consistent with its
neuroprotective, antiepileptic, hypoxia-ischemia, anxiolytic, antipsychotic, analgesic,
anti-inflammatory, anti-asthmatic, and antitumor properties.[37, 50, 67] The evidence
for CBD’s various therapeutic applications was recently reviewed by Pisanti et al
(2017), refer to Table 1.
Another possible therapeutic application which has been investigated is the use of
CBD to treat drug addiction. A recent systematic review concluded that there were a
limited number of preclinical studies which suggest that CBD may have therapeutic
properties on opioid, cocaine, and psychostimulant addiction, and some preliminary
data suggest that it may be beneficial in cannabis and tobacco addiction in humans.
However, considerably more research is required to evaluate CBD as a potential
treatment. [68]
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 18 of 27
Table 1. Overview of diseases for which CBD may have therapeutic benefits taken from
Pisanti et al (2017) [69]
Disease Effects
Alzheimer’s disease Antinflammatory, antioxidant, antiapoptotic in in vitro and in vivo models
of Aβ-evoked neuroinflammatory and neurodegenerative responses.
Parkinson’s disease Attenuation of the dopaminergic impairment in vivo; neuroprotection;
improvement of psychiatric rating and reduction of agitation, nightmare
and aggressive behaviour in patients.
Multiple sclerosis Improved signs of EAE in mice, antinflammatory and immunomodulatory
properties.
Huntington’s
disease
Neuroprotective and antioxidant in mice transgenic models; no significant
clinically important differences in patients.
Hypoxia-ischemia
injury
Short term neuroprotective effects; inhibition of excitotoxicity, oxidative
stress and inflammation in vitro and in rodent models.
Pain Analgesic effect in patients with neuropathic pain resistant to other
treatments.
Psychosis Attenuation of the behavioural and glial changes in animal models of
schizophrenia; anti-psychotic properties on ketamine-induced symptoms
Anxiety Reduction of muscular tension, restlessness, fatigue, problems in
concentration, improvement of social interactions in rodent models of
anxiety and stress; reduced social anxiety in patients.
Depression Anti-depressant effect in genetic rodent model of depression.
Cancer Antiproliferative and anti-invasive actions in a large range of cancer types;
induction of autophagy-mediated cancer cell death; chemopreventive
effects.
Nausea Suppression of nausea and conditioned gaping in rats
Inflammatory
diseases
Antinflammatory properties in several in vitro and in vivo models;
inhibition of inflammatory cytokines and pathways.
Rheumatoid
arthritis
Inhibition of TNF-α in an animal model
Infection Activity against methicillin-resistant Staphylococcus aureus
Inflammatory
bowel and Crohn’s
diseases
Inhibition of macrophage recruitment and TNF-α secretion in vivo and ex
vivo; reduction in disease activity index in Crohn’s patients.
Cardiovascular
diseases
Reduced infarct size through anti-oxidant and anti-inflammatory
properties in vitro and in vivo.
Diabetic
complications
Attenuation of fibrosis and myocardial dysfunction
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 19 of 27
10. Listing on the WHO Model List of Essential Medicines
Cannabidiol is not listed on the WHO Model List of Essential Medicines (20th List) or
the WHO Model List of Essential Medicines for Children (6th List).[70]
11. Marketing Authorizations (as a Medicinal Product)
CBD is present in nabiximols (Sativex®) which is marketed by GW Pharmaceuticals
in a number of countries. [71] As nabiximols also contains an equal amount of THC,
it will be covered in a separate ECDD review.
There are no currently authorized pure CBD products. However, there are several in
development including Epidiolex® and Arvisol®.
Epidiolex® is a liquid formulation of pure plant-derived CBD. It is produced by GW
Pharmaceuticals and has shown positive results in phase 3 trials for Dravet and
Lennox-Gastaut syndromes which are both treatment resistant seizure disorders. The
published results related to this therapeutic application are covered in Section 9:
Therapeutic Applications. [72], [73]
Arvisol® is an oral tablet containing pure CBD. It has been developed by Echo
Pharmaceuticals in the Netherlands and is intended to be registered in the treatment of
disorders such as schizophrenia and epilepsy. Arvisol® is still undergoing Phase I
clinical trials and is not yet available as a medicinal product. [74]
In 2015, the US Food and Drug Administration (FDA) granted GW Pharmaceuticals
Fast Track designation for intravenous CBD to treat Neonatal Hypoxic-Ischemic
Encephalopathy (NHIE).[75] The European Commission also granted orphan
designation (EU/3/15/1520) for cannabidiol to be used in the treatment of perinatal
asphyxia.[76] NHIE and Perinatal Asphyxia are forms of acute or sub-acute brain
injury due to asphyxia caused during the birth process and resulting from deprivation
of oxygen during birth (hypoxia). Currently there are no other treatments available for
these conditions, but there is evidence of the effectiveness of cannabidiol in animal
models. [77]
12. Industrial Use
Pure CBD has no legitimate industrial uses.
13. Non-Medical Use, Abuse and Dependence
At present, there are no case reports of abuse or dependence relating to the use of pure
CBD. There are also no published statistics on non-medical use of pure CBD.
There is unsanctioned medical use of CBD based products. These are produced from
high CBD content plants and distributed in a variety of forms, including oils and
capsules. These products are sold online as unapproved treatments for a variety of
disorders including epilepsy, cancer, AIDS/HIV, anxiety, arthritis, pain, and posttraumatic
stress disorder (PTSD). Additionally, CBD is being used in skin and beauty
products such as shampoos and skin creams.[78, 79] Also see Annex 1: Report on
WHO questionnaire for review of psychoactive substances.
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 20 of 27
14. Nature and Magnitude of Public Health Problems Related to Misuse,
Abuse and Dependence
At present no public health problems (e.g. driving under the influence of drugs cases,
comorbidities) have been associated with the use of pure CBD.
Also refer to Annex 1: Report on WHO questionnaire for review of psychoactive
substances.
15. Licit Production, Consumption and International Trade
Licit production of CBD for medical purposes is described in Section 11. Also refer to
Annex 1: Report on WHO questionnaire for review of psychoactive substances.
16. Illicit Manufacture and Traffic and Related Information
There are no published statistics (e.g. country data on seizures of illicit CBD)
currently available. Refer to Annex 1: Report on WHO questionnaire for review of
psychoactive substances.
17. Current International Controls and Their Impact
Cannabidiol is not listed in the schedules of the 1961, 1971 or 1988 United Nations
International Drug Control Conventions.[80]
However, cannabidiol is being produced for pharmaceutical purposes as an extract of
cannabis by GW Pharmaceuticals. Cannabidiol that is produced as an extract of
cannabis is currently included in Schedule I of the 1961 Convention.
18. Current and Past National Controls
United Kingdom: A statement was issued by the Medicines and Healthcare products
Regulatory Agency (MHRA) in 2016 that products containing CBD used for medical
purposes are considered as a medicine subject to standard licensing requirements. [81]
United States: CBD is one of many cannabinoids present in cannabis, and as such is in
schedule I of the Controlled Substances Act. However in December 2015, the FDA
eased the regulatory requirements to allow researchers to conduct CBD trials. The
Drug Enforcement Agency (DEA) stated that these modifications are intended to
streamline the research process regarding CBD’s possible medicinal value and help
foster ongoing scientific studies. [82]
Canada: CBD is specifically listed in ‘Cannabis, its preparations and derivatives’ as a
controlled substance listed in Schedule II Controlled Drugs and Substances Act.
However, in 2016 Canada’s Access to Cannabis for Medical Purposes Regulations
came into effect. These regulations improve access to cannabis used for medicinal
purposes, including CBD. [83]
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 21 of 27
Australia: In 2015, CBD in preparations for therapeutic use containing 2 per cent or
less of other cannabinoids found in cannabis was placed in Schedule 4 as a
‘Prescription Only Medicine OR Prescription Animal Remedy’. Previous to this it
was captured in Schedule 9 as a prohibited substance.[84]
New Zealand CBD is a controlled drug, however many of the restrictions currently
imposed by the regulations will be removed by the end of 2017. The changes will
mean that CBD products, where the level of other naturally occurring cannabinoids is
less than 2% of the cannabinoid content, will be easier to access for medical use. [85]
Switzerland: CBD is not subject to the Narcotics Act because it does not produce a
psychoactive effect. It is still subject to standard Swiss legislation. [86]
Also refer to Annex 1: Report on WHO questionnaire for review of psychoactive
substances.
19. Other Medical and Scientific Matters Relevant for a Recommendation
on the Scheduling of the Substance
None
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 22 of 27
References
1. NCBI. PubChem Compound Database; CID=26346 August 1 2017]; Available from:
https://pubchem.ncbi.nlm.nih.gov/compound/26346
2. Cayman Chemical. Cannabidiol (DEA Schedule I Regulated Compound). Safety Data
Sheet 2015; Available from: https://www.caymanchem.com/msdss/90080m.pdf.
3. WHO Expert Committee on Drug Dependence : thirty-eighth report. Geneva: World
Health Organization; 2017 (WHO technical report series ; no. 1005). Licence: CC
BY-NC-SA 3.0 IGO.
4. Shah, V.J., Synthesis of cannabidiol stereoisomers and analogs as potential
anticonvulsant agents. The University of Arizona.
5. Bisogno, T., et al., Molecular targets for cannabidiol and its synthetic analogues:
effect on vanilloid VR1 receptors and on the cellular uptake and enzymatic hydrolysis
of anandamide. British Journal of Pharmacology, 2001. 134(4): p. 845-852.
6. Leite, J., et al., Anticonvulsant Effects of the (–) and (+)Isomers of Cannabidiol and
Their Dimethylheptyl Homologs. Vol. 24. 1982. 141-6.
7. Petrzilka, T., W. Haefliger, and C. Sikemeier, Synthese von Haschisch-Inhaltsstoffen.
4. Mitteilung. Helvetica Chimica Acta, 1969. 52(4): p. 1102-1134.
8. Baek, S.-H., M. Srebnik, and R. Mechoulam, Boron triflouride etherate on alimina – a
modified Lewis acid reagent. An improved synthesis of cannabidiol. Vol. 26. 1985.
1083–1086.
9. Mechoulam, R. and L. Hanus, Cannabidiol: an overview of some chemical and
pharmacological aspects. Part I: chemical aspects. Chem Phys Lipids, 2002. 121(1-
2): p. 35-43.
10. Marks, M.D., et al., Identification of candidate genes affecting Δ(9)-
tetrahydrocannabinol biosynthesis in Cannabis sativa. Journal of Experimental
Botany, 2009. 60(13): p. 3715-3726.
11. Taura, F., et al., Cannabidiolic-acid synthase, the chemotype-determining enzyme in
the fiber-type Cannabis sativa. FEBS Letters, 2007. 581(16): p. 2929-2934.
12. Russo, E.B., Cannabidiol Claims and Misconceptions. Trends in pharmacological
sciences, 2017. 38(3): p. 198-201.
13. Sikora, V., et al., Influence of agroclimatic conditions on content of main
cannabinoids in industrial hemp (Cannabis sativa L.). Vol. 43. 2011.
14. ; Available from: https://www.tocris.com/products/minus-cannabidiol_1570#productdetails.
15. Aman, T., A. Rashid, and I. Khokhar, Spectrophotometric Determination of
Cannabidiol. Analytical Letters, 1993. 26(10): p. 2113-2125.
16. Schwope, D.M., K.B. Scheidweiler, and M.A. Huestis, Direct quantification of
cannabinoids and cannabinoid glucuronides in whole blood by liquid
chromatography–tandem mass spectrometry. Analytical and Bioanalytical Chemistry,
2011. 401(4): p. 1273.
17. Sorensen, L.K. and J.B. Hasselstrom, Sensitive Determination of Cannabinoids in
Whole Blood by LC-MS-MS After Rapid Removal of Phospholipids by Filtration. J
Anal Toxicol, 2017. 41(5): p. 382-391.
18. Salomone, A., et al., Simultaneous analysis of several synthetic cannabinoids, THC,
CBD and CBN, in hair by ultra-high performance liquid chromatography tandem
mass spectrometry. Method validation and application to real samples. J Mass
Spectrom, 2012. 47(5): p. 604-10.
19. Wei, B., L. Wang, and B.C. Blount, Analysis of Cannabinoids and Their Metabolites
in Human Urine. Anal Chem, 2015. 87(20): p. 10183-7.
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 23 of 27
20. Aizpurua-Olaizola, O., et al., Simultaneous quantification of major cannabinoids and
metabolites in human urine and plasma by HPLC-MS/MS and enzyme-alkaline
hydrolysis. Drug Test Anal, 2017. 9(4): p. 626-633.
21. Cirimele, V., et al., Testing human hair for Cannabis. III. rapid screening procedure
for the simultaneous identification of delta 9-tetrahydrocannabinol, cannabinol, and
cannabidiol. J Anal Toxicol, 1996. 20(1): p. 13-6.
22. Kim, J.Y., et al., Simultaneous determination of cannabidiol, cannabinol, and delta9-
tetrahydrocannabinol in human hair by gas chromatography-mass spectrometry.
Arch Pharm Res, 2005. 28(9): p. 1086-91.
23. Moore, C., S. Rana, and C. Coulter, Simultaneous identification of 2-carboxytetrahydrocannabinol,
tetrahydrocannabinol, cannabinol and cannabidiol in oral
fluid. J Chromatogr B Analyt Technol Biomed Life Sci, 2007. 852(1-2): p. 459-64.
24. Andrenyak, D.M., et al., Determination of -9-Tetrahydrocannabinol (THC), 11-
hydroxy-THC, 11-nor-9-carboxy-THC and Cannabidiol in Human Plasma using Gas
Chromatography-Tandem Mass Spectrometry. J Anal Toxicol, 2017. 41(4): p. 277-
288.
25. Milman, G., et al., Cannabinoids and Metabolites in Expectorated Oral Fluid
Following Controlled Smoked Cannabis. Clinica Chimica Acta; International Journal
of Clinical Chemistry, 2012. 413(7-8): p. 765-770.
26. Karschner, E.L., et al., Validation of a Two-Dimensional Gas Chromatography Mass
Spectrometry Method for the Simultaneous Quantification of Cannabidiol, Δ(9)-
Tetrahydrocannabinol (THC), 11-Hydroxy-THC and 11-nor-9-Carboxy-THC in
Plasma. Analytical and bioanalytical chemistry, 2010. 397(2): p. 603-611.
27. Andrews, R. and S. Paterson, A validated method for the analysis of cannabinoids in
post-mortem blood using liquid-liquid extraction and two-dimensional gas
chromatography-mass spectrometry. Forensic Sci Int, 2012. 222(1-3): p. 111-7.
28. Gaoni, Y. and R. Mechoulam, Hashish-VII. The isomerization of cannabidiol to
tetrahydrocannabinols. Vol. 22. 1966. 1481–1488.
29. Webster, G.R., L. Sarna, and R. Mechoulam, Conversion of cbd to delta8-thc and
delta9-thc. 2004, Google Patents.
30. Merrick, J., et al., Identification of psychoactive degradants of cannabidiol in
simulated gastric and physiological fluid. Cannabis and Cannabinoid Research, 2016.
1(1): p. 102-112.
31. Watanabe, K., et al., Conversion of cannabidiol to Δ9-tetrahydrocannabinol and
related cannabinoids in artificial gastric juice, and their pharmacological effects in
mice. Forensic Toxicology, 2007. 25(1): p. 16-21.
32. Nahler, G., et al., A Conversion of Oral Cannabidiol to Delta9-Tetrahydrocannabinol
Seems Not to Occur in Humans. Cannabis and Cannabinoid Research, 2017. 2(1): p.
81-86.
33. Consroe, P., et al., Controlled clinical trial of cannabidiol in Huntington’s disease.
Pharmacology Biochemistry and Behavior, 1991. 40(3): p. 701-708.
34. Grotenhermen, F., E. Russo, and A.W. Zuardi, Even High Doses of Oral Cannabidol
Do Not Cause THC-Like Effects in Humans: Comment on Merrick et al. Cannabis
and Cannabinoid Research 2016;1(1):102–112; DOI: 10.1089/can.2015.0004.
Cannabis and Cannabinoid Research, 2017. 2(1): p. 1-4.
35. Consroe, P., et al., Interaction of cannabidiol and alcohol in humans.
Psychopharmacology, 1979. 66(1): p. 45-50.
36. Bonn-Miller, M.O., S.L. Banks, and T. Sebree, Conversion of Cannabidiol Following
Oral Administration: Authors‘ Response to Grotenhermen et al. DOI: 10.1089/can.
2016.0036. Cannabis and Cannabinoid Research, 2017. 2(1): p. 5-7.
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 24 of 27
37. Fasinu, P.S., et al., Current Status and Prospects for Cannabidiol Preparations as
New Therapeutic Agents. Pharmacotherapy, 2016. 36(7): p. 781-96.
38. Hawksworth, G. and K. McArdle, Metabolism and pharmacokinetics of cannabinoids.
The Medicinal Uses of Cannabis and Cannabinoids. Pharmaceutical Press, London,
2004: p. 205-228.
39. Martin-Santos, R., et al., Acute effects of a single, oral dose of d9-
tetrahydrocannabinol (THC) and cannabidiol (CBD) administration in healthy
volunteers. Curr Pharm Des, 2012. 18(32): p. 4966-79.
40. Ohlsson, A., et al., Single-dose kinetics of deuterium-labelled cannabidiol in man
after smoking and intravenous administration. Biological Mass Spectrometry, 1986.
13(2): p. 77-83.
41. Jiang, R., et al., Identification of cytochrome P450 enzymes responsible for
metabolism of cannabidiol by human liver microsomes. Life Sci, 2011. 89(5-6): p.
165-70.
42. McPartland, J.M., et al., Are cannabidiol and Δ9-tetrahydrocannabivarin negative
modulators of the endocannabinoid system? A systematic review. British Journal of
Pharmacology, 2015. 172(3): p. 737-753.
43. Pertwee, R., The diverse CB1 and CB2 receptor pharmacology of three plant
cannabinoids: Δ9‐tetrahydrocannabinol, cannabidiol and Δ9‐
tetrahydrocannabivarin. British journal of pharmacology, 2008. 153(2): p. 199-215.
44. Long, L.E., et al., A behavioural comparison of acute and chronic Δ9-
tetrahydrocannabinol and cannabidiol in C57BL/6JArc mice. International Journal of
Neuropsychopharmacology, 2010. 13(7): p. 861-876.
45. Batalla, A., et al., Neuroimaging studies of acute effects of THC and CBD in humans
and animals: a systematic review. Current pharmaceutical design, 2014. 20(13): p.
2168-2185.
46. Sultan, S.R., et al., A systematic review and meta-analysis of the haemodynamic
effects of Cannabidiol. Frontiers in pharmacology, 2017. 8.
47. Laprairie, R., et al., Cannabidiol is a negative allosteric modulator of the cannabinoid
CB1 receptor. British journal of pharmacology, 2015. 172(20): p. 4790-4805.
48. Bih, C.I., et al., Molecular targets of cannabidiol in neurological disorders.
Neurotherapeutics, 2015. 12(4): p. 699-730.
49. Machado Bergamaschi, M., et al., Safety and side effects of cannabidiol, a Cannabis
sativa constituent. Current drug safety, 2011. 6(4): p. 237-249.
50. Iffland, K. and F. Grotenhermen, An Update on Safety and Side Effects of
Cannabidiol: A Review of Clinical Data and Relevant Animal Studies. Cannabis and
Cannabinoid Research, 2017. 2(1): p. 139-154.
51. Hayakawa, K., et al., Repeated treatment with cannabidiol but not Δ9-
tetrahydrocannabinol has a neuroprotective effect without the development of
tolerance. Neuropharmacology, 2007. 52(4): p. 1079-1087.
52. Katsidoni, V., I. Anagnostou, and G. Panagis, Cannabidiol inhibits the rewardfacilitating
effect of morphine: Involvement of 5-HT1A receptors in the dorsal raphe
nucleus. Addiction Biology, 2013. 18(2): p. 286-296.
53. French, E.D., K. Dillon, and X. Wu, Cannabinoids excite dopamine neurons in the
ventral tegmentum and substantia nigra. Neuroreport, 1997. 8(3): p. 649-652.
54. Vann, R.E., et al., Divergent effects of cannabidiol on the discriminative stimulus and
place conditioning effects of Δ9-tetrahydrocannabinol. Drug and Alcohol
Dependence, 2008. 94(1-3): p. 191-198.
55. Klein, C., et al., Cannabidiol potentiates Δ 9-tetrahydrocannabinol (THC)
behavioural effects and alters THC pharmacokinetics during acute and chronic
treatment in adolescent rats. Psychopharmacology, 2011. 218(2): p. 443-457.
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 25 of 27
56. Jarbe, T.U.C., B.G. Henriksson, and G.C. Ohlin, Δ9-THC as a discriminative cue in
pigeons: effects of Δ8-THC, CBD, and CBN. Archives Internationales de
Pharmacodynamie et de Therapie, 1977. 228(1): p. 68-72.
57. Haney, M., et al., Oral Cannabidiol does not Alter the Subjective, Reinforcing or
Cardiovascular Effects of Smoked Cannabis. Neuropsychopharmacology: official
publication of the American College of Neuropsychopharmacology, 2016. 41(8): p.
1974-1982.
58. Babalonis, S., et al., Oral cannabidiol does not produce a signal for abuse liability in
frequent marijuana smokers. Drug and alcohol dependence, 2017. 172: p. 9-13.
59. Do Val-da Silva, R.A., et al., Protective effects of cannabidiol against seizures and
neuronal death in a rat model of mesial temporal lobe epilepsy. Frontiers in
Pharmacology, 2017. 8.
60. Mechoulam, R. and E. Carlini, Toward drugs derived from cannabis.
Naturwissenschaften, 1978. 65(4): p. 174-179.
61. Cunha, J.M., et al., Chronic administration of cannabidiol to healthy volunteers and
epileptic patients. Pharmacology, 1980. 21(3): p. 175-185.
62. Ames, F. and S. Cridland, Anticonvulsant effect of cannabidiol. South African
medical journal= Suid-Afrikaanse tydskrif vir geneeskunde, 1986. 69(1): p. 14-14.
63. Trumbly, B. Double-blind clinical study of cannabidiol as a secondary
anticonvulsant. in Presented at Marijuana’90 Int. Conf. on Cannabis and
Cannabinoids, Kolympari (Crete). 1990.
64. Devinsky, O., et al., Cannabidiol in patients with treatment-resistant epilepsy: an
open-label interventional trial. The Lancet Neurology, 2016. 15(3): p. 270-278.
65. Devinsky, O., et al., Trial of Cannabidiol for Drug-Resistant Seizures in the Dravet
Syndrome. New England Journal of Medicine, 2017. 376(21): p. 2011-2020.
66. Geffrey, A.L., et al., Drug–drug interaction between clobazam and cannabidiol in
children with refractory epilepsy. Epilepsia, 2015. 56(8): p. 1246-1251.
67. Devinsky, O., et al., Cannabidiol: Pharmacology and potential therapeutic role in
epilepsy and other neuropsychiatric disorders. Epilepsia, 2014. 55(6): p. 791-802.
68. Prud’homme, M., R. Cata, and D. Jutras-Aswad, Cannabidiol as an intervention for
addictive behaviors: a systematic review of the evidence. Substance abuse: research
and treatment, 2015. 9: p. 33.
69. Pisanti, S., et al., Cannabidiol: State of the art and new challenges for therapeutic
applications. Pharmacol Ther, 2017. 175: p. 133-150.
70. World Health Organisation. WHO Model Lists of Essential Medicines. March 2017
21 August 2017]; Available from:
http://www.who.int/medicines/publications/essentialmedicines/en/.
71. GW pharmaceuticals. Sativex. 2016 10 August 2017]; Available from:
https://www.gwpharm.com/products-pipeline/sativex
72. GW pharmaceuticals. GW Pharmaceuticals Announces Positive Phase 3 Pivotal
Study Results for Epidiolex (cannabidiol). March 14 2016 10 August 2017];
Available from: https://www.gwpharm.com/about-us/news/gw-pharmaceuticalsannounces-positive-phase-3-pivotal-study-results-epidiolex.
73. GW pharmaceuticals. GW Pharmaceuticals Announces Second Positive Phase 3
Pivotal Trial for Epidiolex (cannabidiol) in the Treatment of Lennox-Gastaut
Syndrome. September 26 2016 10 August 2017]; Available from:
https://www.gwpharm.com/about-us/news/gw-pharmaceuticals-announces-secondpositive-phase-3-pivotal-trial-epidiolex.
74. Echo Pharmaceuticals B.V. Improved uptake of cannabinoid based medicine.
Available from: http://www.echo-pharma.com/en/about-us/news/improved-uptake-ofcannabinoid-based-medicine
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD) (link werkt niet meer)
Page 26 of 27
75. GW pharmaceuticals. GW Pharmaceuticals Receives FDA Fast Track and EMA
Orphan Designations for Intravenous Cannabidiol in the Treatment of Neonatal
Hypoxic-Ischemic Encephalopathy (NHIE). 6 August 2015 11 August 2017];
Available from: https://www.gwpharm.com/about-us/news/gw-pharmaceuticalsreceives-fda-fast-track-and-ema-orphan-designations-intravenous.
76. European Medicines Agency. EU/3/15/1520 orphan designation for cannabidiol for
the treatment of perinatal asphyxia. 28 July 2015 10 August 2017]; Available from:
http://www.ema.europa.eu/ema/index.jsp?curl=pages/medicines/human/orphans/2015
/08/human_orphan_001612.jsp&mid=WC0b01ac058001d12b.
77. Mohammed, N., et al., Neuroprotective Effects of Cannabidiol in Hypoxic Ischemic
Insult. The Therapeutic Window in Newborn Mice. CNS & Neurological Disorders –
Drug Targets- CNS & Neurological Disorders), 2017. 16(1): p. 102-108.
78. Medical Marijuana Inc. What is cannabidiol? 11 October 2016 20 August 2018];
Available from: http://www.medicalmarijuanainc.com/what-is-cannabidiol/.
79. Canabidol™ The Best Selling CBD Supplement in Europe. CBD cannabis oil. 20
August 2017]; Available from: https://canabidol.com/.
80. United Nations Office on Drugs and Crime. International Drug Control Conventions.
[cited 21 August 2017; Available from:
https://www.unodc.org/unodc/en/commissions/CND/conventions.html.
81. Medicines and Healthcare products Regulatory Agency. 1 August 2017]; Available
from: https://www.gov.uk/government/news/mhra-statement-on-products-containing-cannabidiol-cbd

82. United States Department of Justice Drug Enforcement Administration. 1 August
2017]; https://www.dea.gov/press-releases/2015/12/23/dea-eases-requirements-fda-approved-clinical-trials-cannabidiol
83. Goverment of Canada Justice Laws Website. 1 August 2017]; Available from:
http://laws-lois.justice.gc.ca/eng/acts/c-38.8/FullText.html.
84. Australian Goverment Department of Health Therapeutic Goods Administration. 1
August 2017]; Available from: https://www.tga.gov.au/book/part-final-decisionsmatters-referred-expert-advisory-committee-2.
85. New Zealand Goverment Ministry of Health. 1 August 2017]; Available from: https://www.health.govt.nz/our-work/regulation-health-and-disability-system/medicinal-cannabis-agency/medicinal-cannabis-agency-information-industry/medicinal-cannabis-agency-working-medicinal-cannabis/medicinal-cannabis-agency-cannabidiol-cbd-products

86. Swiss Agency for Therapeutic Products. 1 August 2017]; Available from:
https://www.swissmedic.ch/swissmedic/en/home/news/mitteilungen/products-containing-cannabidiol–cbd—overview.html
39th ECDD (2017) Agenda item 5.2 Cannabidiol (CBD)
Page 27 of 27
Annex 1: Report on WHO Questionnaire for Review of
Psychoactive Substances for the 39th ECDD: Evaluation of
Cannabidiol
Please refer to separate Annex 1 document published on ECDD website