You are on page 1of 11

C H A P T E R

2
The Biosynthesis of Cannabinoids
F. Degenhardt, F. Stehle, O. Kayser
Laboratory of Technical Biochemistry, Department of Biochemical and Chemical Engineering,
TU Dortmund University, Dortmund, Germany

SUMMARY POINTS • Two cannabigerol-like compounds were detected in


• This chapter focuses on the pathway which leads the aerial parts of Helichrysum umbraculigerum Less., a
to the biosynthesis of phytocannabinoids in plant common in the eastern parts of South Africa.
C. sativa L. • N-alkyl amides (cannabinomimetics), found in the
medicinal plants Echinaceae angustifolia and Echinaceae
• CBGA is the central precursor of
purpurea (purple cornflower), are known to interact
phytocannabinoid biosynthesis in Cannabis.
with the CB2 receptor.
• CBGAS, only three enzymes—THCAS, CBDAS,
and CBCAS—are involved in the biosynthesis of
phytocannabinoids in Cannabis plants. LIST OF ABBREVIATIONS
• Sequences of CBDAS and THCAS are known.
• The carboxyl group in CBGA seems to be AAE Acyl-activating enzyme
essential for the enzymatic reactions catalyzed by BBE Berberine bridge enzyme
CBDAS, CBCAS, and THCAS. CBC Cannabichromene
CBCA Cannabichromenic acid
• The diversity of more than 60 cannabinoids is the
CBCAS Cannabichromenic acid synthase
result of nonenzymatic modifications.
CBCVA Cannabichrovarinic acid
• Propyl cannabinoids occur by the prenylation of CBD Cannabidiol
divarinic acid (DA) with geranyl diphosphate CBDA Cannabidiolic acid
(GPP). CBDAS Cannabidiolic acid synthase
CBDV Cannabidivarin
CBDVA Cannabidivarinic acid
KEY FA C T S OF CBG Cannabigerol
PH YT OCA NNAB INOIDS —B ES IDE S CBGA Cannabigerolic acid (3-geranyl olivetolate)
C.   SAT I VA CBGAS Cannabigerolic acid synthase
• Phytocannabinoids are plant-derived natural compounds CBGVA Cannabigerovarinic acid
that act as ligands to cannabinoid receptors (CB1 and CB2) CBN Cannabinol
or share chemical similarity with cannabinoids. CBNRA Cannabinerolic acid (cis-CBGA)
• C. sativa L. is intensively investigated for the presence CHS Chalcone synthase
of phytocannabinoids. To date, only a few plants are CsAAE1 C. sativa hexanoyl-CoA synthetase 1
discovered that contain phytocannabinoids other than CsAAE3 C. sativa hexanoyl-CoA synthetase 2
the ones known from Cannabis. CsHCS1 C. sativa hexanoyl-CoA synthetase 1
• The New Zealand liverwort Radula marginata and CsHCS2 C. sativa hexanoyl-CoA synthetase 2
Japanese liverwort Radula perrottetii contain perrotteti- DA Divarinic acid
nene, a naturally occurring bibenzyl cannabinoid. DABB Dimeric α + β barrel
DMAPP Dimethylallyl diphosphate
Handbook of Cannabis and Related Pathologies. http://dx.doi.org/10.1016/B978-0-12-800756-3.00002-8
Copyright © 2017 Elsevier Inc. All rights reserved.
13
14 2.  The Biosynthesis of Cannabinoids

DOXP 1-Deoxy-d-xylulose-5-phosphate for the biosynthesis of cannabinoids, the terpenopheno-


GOT Geranylpyrophosphate:olivetolate lic constituents that show psychoactive effects. But since
geranyltransferase other plants also have secondary metabolites that inter-
GPP Geranyl diphosphate act with the human cannabinoid receptors, a new defini-
HTAL Hexanoyltriacetic acid lactone tion had to be made. Hence, phytocannabinoids are now
IPP Isopentenyl diphosphate defined as any plant-derived natural compound that
MEP 2C-methyl-d-erythritol-4-phosphate can act as a ligand to human cannabinoid receptors (CB1
MVA Mevalonate and CB2) or share chemical similarity with cannabinoids
NPP Neryl diphosphate (Gertsch, Pertwee, & Di Marzo,  2010). Interestingly, all
OA Olivetolic acid parts of the Cannabis plant, with the exception of seeds,
OAC Olivetolic acid cyclase can contain cannabinoids, but they mainly accumulate
OLS Olivetol synthase in the glandular trichomes of female flowers (Gagne
PKS Polyketide synthase et al., 2012; van Bakel et al., 2011).
SNP Single nucleotide polymorphism The following chapter focuses on the pathway that
STS Stilbene synthase leads to the enzymatic biosynthesis of cannabinoids. For
THC Tetrahydrocannabinol a long time, it was postulated that the key intermedi-
THCA Tetrahydrocannabinolic acid ate is cannabidiol (CBD) or cannabidiolic acid (CBDA),
THCAS Tetrahydrocannabinolic acid synthase both resulting from a condensation of a monoterpene,
THCV Tetrahydrocannabivarin and olivetol or olivetolic acid (OA), respectively. In 1964,
THCVA Tetrahydrocannabivarinic acid Gaoni and Mechoulam postulated cannabigerol (CBG)
as the key intermediate, the condensation product of ge-
ranyl diphosphate (GPP), and olivetol or OA. Based on
INTRODUCTION this, they concluded that the cannabinoids CBD, tetrahy-
drocannabinol (THC) and cannabinol (CBN) are all de-
Cannabis sativa L. (hemp) is one of the oldest do- rived from CBG, and just differ in the way of cyclization
mestic plants in the history of mankind, and has been (Gaoni & Mechoulam, 1964). Finally, incorporation stud-
cultivated for at least 10,000  years (Schultes, Klein, ies with 13C-labeled glucose have shown that GPP and
Plowman, & Lockwood,  1974). Together with Humulus OA are indeed the precursors for formation of cannabig-
lupulus (hop), C. sativa belongs to the small family of erolic acid (CBGA). Thus, the general structure of canna-
Cannabaceae. Cannabis is an annual, usually dioecious, binoids is assembled by two parts: (1) a diphenol (resor-
wind-pollinated herb, with both male and female flow- cin) carrying an alkyl chain (OA); and (2) a monoterpene
ers growing on separate plants. The plant is well known moiety (GPP) (Fig.  2.1). Subsequently, Fellermeier and

FIGURE 2.1  General structure of cannabinoids and their precursors, olivetolic acid, and geranyl diphosphate. Cannabinoids are composed
of two parts: a cyclic monoterpene part (red), and a diphenol (resorcin) part, carrying an alkyl chain (blue). The dibenzopyran-numbering system
is used.

I.  Setting the scene, botanical, general and international aspects


Cannabinoid precursor biosynthesis 15
coworkers postulated CBGA as the central cannabinoid via the lipoxygenase pathway (Marks et al., 2009; Stout,
precursor (Fellermeier, Eisenreich, Bacher, & Zenk, 2001; Boubakir, Ambrose, Purves, & Page, 2012). Nevertheless,
Fellermeier & Zenk, 1998). Interestingly, free OA has nev- further studies are necessary to clarify the origin of the
er been detected in Cannabis plant material until now. hexanol moiety.
It is worthy to note that, although more than 60 can- Hexanoyl-CoA is a medium-chain fatty acyl-CoA that
nabinoids are known, only three enzymes, besides can- can be detected in high amounts in Cannabis flowers
nabigerolic acid synthase (CBGAS), namely tetrahydro- (Stout et al., 2012). It is synthesized by an acyl-activating
cannabinolic acid synthase (THCAS), cannabidiolic acid enzyme (AAE) called hexanoyl-CoA synthetase (Marks
synthase (CBDAS), and cannabichromenic acid synthase et  al.,  2009; Page & Stout,  2013). AAEs can use short,
(CBCAS), are involved in cannabinoid biosynthesis. The medium, long as well as very long-chain fatty acids as
resulting acidic cannabinoids are the most abundant ones carboxylic acid substrates. Two novel enzymes were
accumulating in Cannabis. The neutral and psychoactive identified, C. sativa hexanoyl-CoA synthetase 1 (CsHCS1
forms are the results of nonenzymatic decarboxylation or CsAAE1) and C. sativa hexanoyl-CoA synthetase 2
during storage, heat or sunlight; explaining the heating (CsHCS2 or CsAAE3) that are capable of producing
of plant material (ie, smoking or baking), during Canna- hexanoyl-CoA using hexanoate and CoA as substrates.
bis consumption (Fischedick, Hazekamp, Erkelens, Choi, Based on transcript levels, CsHCS1 seems to be tri-
& Verpoorte, 2010; Taura et al., 2007a). Thus, the broad chome-specific. Although CsHCS2 exhibits lower tran-
diversity of the different cannabinoids is mainly due to script levels, in comparison to CsHCS1, it is abundant in
nonenzymatic transformation or degradation of both all tissues. The gene of CsHCS1 consists of a 2163-nucle-
acidic and neutral cannabinoids by the effects of light otide open reading frame, and encodes a 720-amino acid
(UV irradiation) and auto-oxidation (Crombie, Ponsford, polypeptide chain. The gene of CsHCS2 is composed of
Shani, Yagnitinsky, & Mechoulam,  1968; Razdan, a 1632-nucleotide open reading frame, and encodes a
Puttick, Zitko, & Handrick, 1972). It is still unclear if all 543-amino acid polypeptide chain. Both CsHCSs gener-
these forms are present in living plants as natural or ar- ally require divalent cations for activity. This was shown
tefacts, due to storage and sample preparation (ElSohly by adding Mg2+, Mn2+, and Co2+ to the enzyme assays.
& Slade, 2005). Thus, CsHCS1 preferentially accepts Mg2+, and CsHCS2
Co2+. The highest enzyme activity was detected at 40°C
and pH 9 for both enzymes. Furthermore, both enzymes
can be inhibited by high concentrations of CoA (Page &
CANNABINOID PRECURSOR
Stout, 2013; Stout et al., 2012).
BIOSYNTHESIS
Taken together, the published data suggest that
CsHCS1 is the enzyme involved in the biosynthesis of
Polyketide Pathway Toward Olivetolic Acid
cannabinoids: (1) it is the most abundant AAE in tri-
The origin of hexanoate in trichomes has not been chomes; (2) it is highly specific for short-chain fatty acyl-
elucidated so far. Suzuki, Kurano, Esumi, Yamaguchi, and CoA, particularly hexanoate (KM value in the nM range);
Doi (2003) showed that the side-chain moiety of alkyl- and (3) it is localized in the cytosol, as suggested for the
resorcinols is formed by fatty acid units, but it remains olivetol synthase (see later). In contrast, CsHCS2 is lo-
unclear if the moiety is the result of biosynthesis or deg- calized in the peroxisomes and accepts a broad range of
radation of fatty acids. Studies regarding the incorpora- substrates, while showing a KM value for hexanoate in
tion of 13C-labels into cannabinoids indicate that hexano- the mM range (Page & Stout, 2013; Stout et al., 2012).
ate is synthesized from acetyl-CoA as a starter unit, and The alkylresorcinol moiety of cannabinoids is de-
five molecules of malonyl-CoA. These building blocks rived from OA, the product of polyketide synthases
are precursors of the fatty acid biosynthesis (Fellermeier (PKSs) that catalyze the aldol condensation of hexanoyl-
et al., 2001). CoA with three molecules of malonyl-CoA (Fellermeier
Based on this, two pathways are feasibly possible, et  al.,  2001; Raharjo, Chang, Choi, Peltenburg-Looman,
after analysis of a cDNA/EST library generated from & Verpoorte, 2004) (Fig. 2.2). The second precursor mal-
female flowers (glands) of C. sativa. First, the hexanoyl onyl-CoA is predominantly derived from acetyl-CoA
residue could be obtained by an early termination of the by carboxylation. The ATP-dependent reaction is cata-
fatty acid biosynthesis. Subsequently, the hexanoyl moi- lyzed by an acetyl-CoA carboxylase (EC 6.4.1.2). The en-
ety of the resulting hexanoyl-ACP would be cleaved by a zyme utilizes the first step in the fatty acid biosynthesis
thioesterase or transferred to CoA by an ACP-CoA trans- (Chen, Kim, Weng, & Browse, 2011; Konishi, Shinohara,
acylase. Finally, acyl-CoA synthetase would catalyze the Yamada, & Sasaki, 1996). Taura et al. (2009) discovered
conversion of the obtained n-hexanol to hexanoyl-CoA a plant type III PKS in flowers and rapidly expanding
(Marks et  al.,  2009). Second, n-hexanol could be pro- leaves of C. sativa. The gene of olivetol synthase (OLS)
duced by the breakdown of C18 unsaturated fatty acids encodes a 385-amino acid polypeptide chain that does

I.  Setting the scene, botanical, general and international aspects


16 2.  The Biosynthesis of Cannabinoids

TABLE 2.1 Enzymes Involved In Cannabinoid Biosynthesis in C. sativa L


Enzyme Accession no.a EC no. References

Olivetol synthase OLS AB164375 2.3.1.206 Taura et al. (2009)

Olivetolic acid cyclase OAC AFN42527.1 4.4.1.26 Gagne et al. (2012)

Cannabigerolic acid synthase CBGAS US2012/0144523 2.5.1.102 Fellermeier and Zenk (1998);
A1b Page and Boubakir (2012)

Cannabichromenic acid synthase CBCAS 1.3.3.- Morimoto et al. (1998)

Cannabidiolic acid CBDAS AB292682 1.21.3.8 Taura et al. (2007a)


synthase

Tetrahydrocannabinolic acid synthase THCAS AB057805 1.21.3.7 Sirikantaramas et al. (2004)


The table lists the enzymes and the corresponding GenBank accession numbers involved in biosynthesis of C. sativa phytocannabinoids.
a
GenBank.
b
Patent number.

not contain a signal peptide (Table 2.1). The OLS protein Finally, using both OLS and OAC with hexanoyl-CoA
has a theoretical molecular mass of 43 kDa, as confirmed and malonyl-CoA in one assay, the formation of OA,
by SDS-PAGE analysis. However, size-exclusion chro- pentyldiacetic acid (triketide pyrone), and hexanoyltri-
matography experiments revealed a molecular mass of acetic acid lactone (HTAL; tetraketide pyrone) could be
about 89 kDa, indicating a homodimeric enzyme (Gagne demonstrated (Page & Gagne,  2013) (Fig.  2.2). It is as-
et al., 2012; Taura et al., 2009). OLS (PKS-1) was prelimi- sumed that OLS catalyzes the formation of an interme-
narily crystallized by Taguchi et al. (2008) and the struc- diate that is subsequently converted into OA by OAC
ture was finally published by Yang et al. (2016). It is of (Gagne et al., 2012; Taguchi et al., 2008).
interest that the enzyme does not produce OA, but olive-
tol, triketide pyrone, and tetraketide pyrone. Analysis of
Biosynthesis of Geranyl Diphosphate
the amino acid sequence displayed a high similarity with
those of Medicago sativa chalcone synthase (CHS), and The monoterpene moiety of cannabinoids (Fig. 2.2) is
other plant PKSs (60–70%). Additionally, the catalytic tri- derived from GPP. Its precursors, isopentenyl diphos-
ade residues of CHS (Cys164-His303-Asn336) are conserved phate (IPP), and dimethylallyl diphosphate (DMAPP),
(Taura et al., 2009). Since CHSs catalyze intramolecular are predominantly (>98%) biosynthesized via the
C6 → C1 Claisen condensations, Raharjo, and coworkers 2C-methyl-d-erythritol-4-phosphate (MEP) pathway
were the first to suggest in 2004 (Raharjo et al., 2004) that [also termed as nonmevalonate pathway or 1-deoxy-d-
OLS could be a stilbene synthase (STS). These enzymes xylulose-5-phosphate (DOXP) pathway] (Fellermeier
catalyze C2 → C7 aldol condensations, followed by a et al., 2001). These results are supported by Marks et al.
decarboxylation step. Additionally, studies by Austin, (2009). They isolated RNA from the glands of a tetra-
Bowman, Ferrer, Schröder, & Noel (2004) showed that hydrocannabinolic acid (THCA)-producing Cannabis
the cyclization reaction can be changed from a Claisen- strain and generated a cDNA library. After sequencing,
type (CHS) to an aldol-type (STS) by substitution of a they were able to identify all but one enzyme involved
few amino acids in CHS (= aldol switch). in the MEP pathway. Additionally, Stout et  al. (2012)
Nevertheless, since OLS alone is not capable to form found high expression of MEP pathway genes in Can-
OA, another enzyme/PKS might be involved in the nabis flowers. Furthermore, in higher plants the MEP
biosynthesis. The missing enzyme should catalyze a C2 pathway, mainly involved in secondary metabolism,
→ C7 intramolecular aldol condensation upon which is localized in plastids (described in detail elsewhere,
the carboxylate moiety is preserved. This is important for example, Eisenreich, Bacher, Arigoni, & Rohdich,
since CBGAS does not accept olivetol as a prenyl do- (2004), or Hunter (2007), whereas the mevalonate (MVA)
nor (Fellermeier & Zenk, 1998). Gagne et al. (2012) iso- pathway, predominantly contributing to primary me-
lated a gene encoding a 101-amino acid polypeptide tabolism, is localized in the cytosol. The compartmental
chain. This small protein (12 kDa) shows similarities to separation between these two pathways is not absolute.
a polyketide cyclase that belongs to the dimeric α + β The metabolites of both pathways can be transported bi-
barrel (DABB)-type protein family. Furthermore, the directionally across the plastid membranes (Eisenreich
identified gene exhibits high expression levels in glan- et al., 2004).
dular trichomes. Together, this made the polyketide Subsequently, the head-to-tail condensation of IPP
cyclase a promising candidate for the missing olivetolic and DMAPP to form GPP is catalyzed by geranyl di-
acid cyclase (OAC). phosphate synthase (Fig. 2.2) (Burke et al., 1999).

I.  Setting the scene, botanical, general and international aspects


Cannabinoid precursor biosynthesis 17

FIGURE 2.2  Biosynthesis of cannabigerolic acid (CBGA). The biosynthesis of the central intermediate CBGA is colored in dark green. The
minor products CBNRA and CBGVA are shaded in light green. The precursor pathways are highlighted in light blue (GPP) and blue (OA). MEP,
2C-methyl-d-erythritol-4-phosphate; DOXP, 1-deoxy-d-xylulose-5-phosphate; MVA, mevalonate (Burke, Wildung, & Croteau,  1999; de Meijer
et al., 2009; Fellermeier & Zenk, 1998; Page & Gagne, 2013; Taura et al., 2009).

Cannabigerolic Acid Biosynthesis by mass spectrometry (MS) measurements as CBGA and


Cannabigerolic acid synthase (CBGAS) or geranylpy its cis-isomer cannabinerolic acid (CBNRA; Fellermeier
rophosphate:olivetolate geranyltransferase (GOT) pre- and Zenk, 1998, used CBNA instead of CBNRA). The en-
dominantly catalyzes the C-prenylation of OA by GPP zyme activity was found to be Mg2+-dependent. CBGAS
to form CBGA (Fig.  2.2). CBGA is presumed to be the seems to be specific for OA as a prenyl acceptor, but also
central precursor for cannabinoid biosynthesis, since dif- accepts different prenyl donors like GPP and, to a lesser
ferent cyclization of the prenyl moiety leads to THCA degree, neryl diphosphate (NPP) (Fig. 2.2). The produc-
or its isomers cannabichromenic acid (CBCA) and CBDA tion ratio of CBGA/CBNRA changes from 2:1 to 1:1
(Page & Boubakir,  2012; Sirikantaramas, Morimoto, & when NPP is used as a prenyl donor instead of GPP.
Shoyama, 2007). However, the aromatic prenyltransferase CBGAS
Fellermeier and Zenk (1998) detected the enzyme in seems to be a soluble enzyme, but Fellermeier and Zenk
crude homogenates of rapidly expanding young leaves (1998) could not completely exclude a membrane-bound
of C. sativa. This part of the plant contains the later en- activity. Besides, two soluble hop prenyltransferases,
zymes of the THCA biosynthetic pathway (Morimoto, involved in the biosynthesis of hop bitter acids, are de-
Komatsu, Taura, & Shoyama, 1997; Taura et al., 1995a). scribed by Zuurbier, Fung, Scheffer, & Verpoorte (1998).
There are indications that CBGAS, like other prenyl- Nevertheless, these are the only descriptions of soluble
transferases, is a membrane-bound prenyltransferase plant C-prenylating enzymes; until now, it was not pos-
(Yamamoto, Kimata, Senda, & Inoue,  1997). However, sible to get the sequence information or to isolate the cor-
Fellermeier and Zenk (1998) could not detect any enzyme responding genes or enzymes.
activity in particulate fractions, but in the soluble fraction Contradictorily, all known sequences of plant aro-
of the crude extract. Two major products were identified matic prenyltransferases belong to membrane-bound

I.  Setting the scene, botanical, general and international aspects


18 2.  The Biosynthesis of Cannabinoids

enzymes (Yamamoto et  al.,  1997; Yamamoto, Senda, & cations, whereas the highest enzyme activity was ob-
Inoue,  2000; Zhao, Inoue, Kouno, & Yamamoto,  2003). tained by using Mg2+ (Page & Boubakir, 2012).
This is in accordance with the second report dealing with
the CBGAS (Page & Boubakir, 2012). They published a
sequence of CBGAS that was mainly expressed in glan- CANNABINOID PATHWAY
dular trichomes of female flowers and young leaves
of Cannabis plants. The gene encodes a 395-amino acid CBGA, the central precursor of cannabinoid biosyn-
polypeptide chain showing a membrane-bound type of thesis, is converted by three enzymes (Fig. 2.3): CBDAS,
prenyltransferases. They were able to express the recom- CBCAS, and THCAS. They predominantly use CBGA
binant CBGAS in Sf9 insect as well as in Saccharomyces as substrate, and catalyze the stereoselective, oxida-
cerevisiae cells, and verified the CBGAS activity in the tive cyclization of the monoterpene moiety of CBGA to
microsomal fractions. Using MS measurements, CBGA CBDA, CBCA, or THCA, respectively. The THCAS and
(3-geranyl olivetolate; comparison with CBGA stan- CBDAS reactions are oxygen-dependent, producing hy-
dard) was identified as the major product, and 5-geranyl drogen peroxide proportional to either CBDA or THCA
olivetolate (identification only by LC-MS analysis) as the (Sirikantaramas et  al.,  2004; Taura et  al., 2007b). Re-
minor product. Furthermore, Page and Boubakir (2012) markably, the CBCAS reaction is oxygen independent,
showed that CBGAS is specific only to GPP as a prenyl and can be inhibited by hydrogen peroxide. Thus, the
donor, and approves OA, olivetol, phlorisovalerophe- enzyme seems not to be an oxygenase or a peroxidase
none, naringenin, and resveratrol as prenyl acceptor. Ad- (Morimoto, Komatsu, Taura, & Shoyama,  1998). Fur-
ditionally, the enzyme reaction is dependent on divalent thermore, all three enzymes also convert CBNRA, the

FIGURE 2.3  Biosynthesis of cannabinoids. The enzymatically catalyzed reactions are highlighted in dark green. All nonenzyme-dependent
modifications reactions are colored in light green. Biosynthesis of C3-cannabinoids starting from cannabigerovarinic acid (CBGVA) is carried out
by the same enzymes and for better clarity not shown (Crombie et al., 1968; de Meijer, 2011; Morimoto et al., 1998; Shoyama, Fujita, Yamauchi, &
Nishioka, 1968; Shoyama, Oku, Yamauchi, & Nishioka, 1972; Taura et al., 1995a; 1996).

I.  Setting the scene, botanical, general and international aspects


Cannabinoid pathway 19
cis-isomer of CBGA, with a lower specificity (Morimoto acid substitutions. These substitutions seem to be the rea-
et  al.,  1998; Shoyama et  al.,  2012; Sirikantaramas, son for decreased THCAS activity in “fiber-type” strains
Morimoto, & Shoyama, 2007; Taura et al., 1995b; Taura, (Kojoma, Seki, Yoshida, & Muranaka,  2006). Minise-
Morimoto, & Shoyama, 1996). Since no enzymatic activ- quencing of samples from both types of Cannabis plants
ity was detectable using the neutral cannabinoid CBG, showed three different single nucleotide polymorphism
the carboxyl group in CBGA seems to be essential for the (SNP) genotypes. “Fiber-type” plants are homozygous
enzymatic reactions catalyzed by THCAS, CBDAS, and for the inactive THCAS form. “Drug-type” plants are
CBCAS (Morimoto et al., 1997; Taura et al., 1995a; Taura either homozygous or heterozygous for the active form
et al., 1996). of THCAS. It seems that only a single copy of the gene
Besides, it was postulated that THCA is biosyn- encoding the active THCAS form is necessary for the
thesized and stored in the storage cavity of the glan- biosynthesis of THCA (Rotherham & Harbison, 2011).
dular ­trichomes of Cannabis plants (Sirikantaramas,
­Morimoto, & Shoyama, 2007).
Cannabidiolic Acid Synthase
The gene of the wildtype CBDAS encodes a 544-amino
Tetrahydrocannabinolic Acid Synthase acid polypeptide (Table  2.1). According to Taura et  al.
The THCAS gene encodes a 545-amino acid polypep- (2007b), processed CBDAS consists of 517 amino ac-
tide chain (Table 2.1). According to Sirikantaramas et al. ids following cleavage of the 28 amino acid long signal
(2004), a 28 amino acid long signal peptide is cleaved peptide. The mature CBDAS has a theoretical molecu-
in the processed THCAS, leading to a protein of 517 lar mass of 59 kDa. An actual mass of about 74 kDa was
amino acids. The mature THCAS has a theoretical mo- detected by SDS-PAGE that is possibly caused by post-
lecular mass of 59 kDa. An actual mass of about 75 kDa translational glycosylation of seven Asn residues (Taura
was detected using SDS-PAGE (Taura et  al., 1995b). et al., 1996; Taura et al., 2007b). CBDAS is a monomeric
This could be explained by posttranslational modifica- enzyme with a pH optimum of 5.0 (Taura et  al.,  1996).
tions, since eight possible Asn glycosylation sites were A comparison between THCAS and CBDAS revealed a
confirmed (Sirikantaramas et  al.,  2004). Furthermore, sequence similarity of 84% (Taura et al., 2007b) (Figs. 2.4
deglycosylated THCAS indeed showed a molecular and 2.5). Like THCAS, CBDAS is a flavinated enzyme in
mass of 59 kDa, and remained fully active (Taura et al., which His114 and Cys176 are most likely the FAD-binding
2007b). THCAS is a monomeric enzyme with the high- sites. Since CBDAS exhibits structural and biochemical
est activity between pH 5.5 and pH 6.0 (Taura et  al., properties related to those of THCAS, it is probable that
1995b). Sequence comparison identified similarities the reaction mechanism of CBDAS is similar to that of
to the berberine bridge enzyme (BBE) of Eschscholzia THCAS (Taura et al., 2007b).
californica (Shoyama et  al.,  2012). BBE belongs to the
family of oxidoreductases and has a covalently bound
Cannabichromenic Acid Synthase (CBCAS)
FAD (Kutchan & Dittrich, 1995). The THCAS amino acid
sequence revealed a flavinylation consensus sequence The sequence of CBCAS is still unknown. Morimoto
(Arg110-Ser-Gly-Gly-His114) in which His114 is probably et al. (1998) purified the enzyme to apparent homogene-
the FAD-binding site (Sirikantaramas et al., 2004). This ity, but its sequence is not yet available in public databas-
could be confirmed by X-ray crystallography (PDB ID: es. According to van Bakel et al. (2011), possible CBCAS
3VTE) at a resolution of 2.75Å. The results show that candidates are currently analyzed biochemically.
the enzyme is composed of two domains and one FAD However, CBCAS was isolated and partially purified
binding pocket present in between. Besides His114, a from young leaves of C. sativa (Morimoto et  al.,  1997;
second residue, Cys176, could be identified to be cova- 1998). In contrast to CBDAS and THCAS, CBCAS seems
lently bound to the FAD (Shoyama et al., 2012). Based to be homodimer with a determined native molecular
on X-ray structure data and mutational analysis of mass of 136  kDa and a maximum activity at pH 6.5. A
THCAS, a possible catalytic reaction mechanism of molecular mass of 71 kDa was estimated for the mono-
THCAS was proposed by Shoyama et al. (2012), assign- mers using SDS-PAGE. According to kinetic data, CBCAS
ing a central role to Tyr484 in the catalytic mechanism has a higher affinity for CBGA than THCAS and CBCAS
(Fig. 2.5). Nevertheless, since the crystal structure was (Morimoto et al., 1998). CBCA and its neutral form CBC
published without substrate analoga, further studies are both racemic. Studies of Morimoto et al. (1997) sug-
are necessary to verify the suggested mechanism. gested that both enantiomers of CBCA are formed by
Cannabis plants can be divided into two groups: a CBCAS catalyzed reaction with a molar ratio of 5:1.
“fiber-type” and “drug-type” plants (see dictionary). But it is still unknown which of the two isomers is the
Alignment of THCAS coding sequences from “fiber- major product (Gaoni & Mechoulam,  1971; Morimoto
type” and “drug-type” plants showed 37 major amino et al., 1997; Taura et al., 2007a).

I.  Setting the scene, botanical, general and international aspects


20 2.  The Biosynthesis of Cannabinoids

FIGURE 2.4  Alignment of amino acid sequences of THCA synthase and CBDA synthase. The alignment was performed with CLUSTAL W
using the BLOSUM 62 matrix (Henikoff & Henikoff, 1992; Larkin et al., 2007). The signal peptide cleavage sites are indicated by a triangle. Second-
ary elements (α-alpha-helices; β-beta-sheets; TT-turns; η-310 helix) are shown for the THCAS. Fully conserved residues are shaded in black. The
sequences show an overall identity of 84%. The figure was made with ESPript 3.0 (Robert & Gouet, 2014). THCAS, tetrahydrocannabinolic acid
synthase; CBDAS, cannabidiolic acid synthase.

I.  Setting the scene, botanical, general and international aspects


Mini-dictionary 21

FIGURE 2.5  X-ray structure of the active center of THCAS. The backbone is shown as cartoon diagram (PDB ID: 3VTE). The FAD molecule
(orange) is covalently attached to His114 and Cys176 (yellow). The active site residues are highlighted in green (Shoyama et al., 2012). The close-up of
active center was prepared with PyMOL. THCAS, tetrahydrocannabinolic acid synthase.

Cannabinoids with Propyl Side Chains (THCV) and/or cannabidivarin (CBDV) are usually only
detectable in Cannabis indica (de Zeeuw, 1972).
In contrast to the classic C5-phytocannabinoids,
which contain an n-pentyl side chain, cannabinoids with
an n-propyl side chain are called C3-phytocannabinoids MINI-DICTIONARY
or propyl cannabinoids. The C-prenylation of divarinic
acid (DA) instead of OA by GPP yields in cannabiger- Cannabinoids  Cannabinoids are a group of terpenophenolic
ovarinic acid (CBGVA) (Fig. 2.2) (de Meijer, Hammond, compounds. They show affinities to cannabinoid receptors (CB1,
& Micheler, 2009). The formation of propyl cannabinoids CB2) or are structurally related to tetrahydrocannabinol (THC).
Cannabinoids can be differentiated into phytocannabinoids,
does not occur by shortening the side chain of pentyl can-
synthetic cannabinoids, and endocannabinoids.
nabinoids (Kajima & Piraux, 1982). CBGVA is the central CBGA  Cannabigerolic acid (CBGA) is the central precursor of
branch-point intermediate in the biosynthesis of C3-can- phytocannabinoid biosynthesis. It is nonpsychoactive.
nabinoid acids, like CBGA for pentyl cannabinoid acids. “Drug-type” plants  These are THCA-rich plants. The THCA
The enzymes CBDAS, CBCAS, and THCAS are not se- is converted into psychoactive ∆9-THC by a nonenzymatic
decarboxylation that enables the plants to be labelled as THC-rich.
lective for the length of the alkyl side chain, and can use
“Fiber-type” plants  “Fiber-type” plants are also known as
both as a substrate. The resulting cannabinoids are called “nondrug” plants. These plants have a low (<0.2%) or no THCA
cannabidivarinic acid (CBDVA), cannabichrovarinic acid content, but they contain a high amount of cannabidiolic acid
(CBCVA), and tetrahydrocannabivarinic acid (THCVA). (CBDA).
The diverse amount of 2-carboxylic acids in different Phytocannabinoids  Phytocannabinoids are a unique group
of secondary metabolites (cannabinoids) occurring naturally
Cannabis strains is caused by dissimilar enzyme specifici-
in plants. Other names include natural cannabinoids or herbal
ties at the level of CBGA or CBGA-analogs formation (de cannabinoids (see Key facts).
Meijer et al., 2009; Shoyama, Hirano, & Nishioka, 1984). ∆8-THC  In Cannabis plants, ∆8-tetrahydrocannabinol (∆8-THC)
Relatively high amounts of tetrahydrocannabivarin is detectable in low amounts (<1% of the present ∆9-THC). Like

I.  Setting the scene, botanical, general and international aspects


22 2.  The Biosynthesis of Cannabinoids

∆9-THC, it is also psychoactive. Maybe ∆8-THC is an artefact of the National Academy of Sciences of the United States of America, 109,
extraction and/or analysis process. The term THC includes a 12811–12816.
combination of ∆8-THC and ∆9-THC. Gaoni, Y., & Mechoulam, R. (1964). Isolation, structure, and partial
∆9-THC  ∆9-Tetrahydrocannabinol (∆9-THC) and ∆1-THC synthesis of an active constituent of hashish. Journal of the American
describe the same compound, differing in the numbering system Chemical Society, 86, 1646–1647.
used (dibenzopyran-numbering and monoterpene-numbering Gaoni, Y., & Mechoulam, R. (1971). The isolation and structure of
system, respectively). ∆9-THC is responsible for the psychoactive ∆1-tetrahydrocannabinol and other neutral cannabinoids from
effects of Cannabis products. It binds to the human cannabinoid hashish. Journal of the American Chemical Society, 93, 217–224.
receptors located in the central and peripheral nervous system. Gertsch, J., Pertwee, R. G., & Di Marzo, V. (2010). Phytocannabinoids
Misleadingly, ∆9-THC is termed as the main component of drug- beyond the Cannabis plant – do they exist? British Journal of Phar-
type Cannabis plants (see THCA). macology, 160, 523–529.
∆9-THCA  ∆9-Tetrahydrocannabinolic acid (THCA), not THC, Henikoff, S., & Henikoff, J. G. (1992). Amino acid substitution matrices
is the main component of drug-type Cannabis plants. It is from protein blocks. Proceedings of the National Academy of Sciences of
nonpsychoactive. THCA is converted into neutral psychoactive ∆9- the United States of America, 89, 10915–10919.
THC by a nonenzymatic decarboxylation during heating or storage. Hunter, W. N. (2007). The non-mevalonate pathway of isoprenoid pre-
cursor biosynthesis. Journal of Biological Chemistry, 282, 21573–21577.
Kajima, M., & Piraux, M. (1982). The biogenesis of cannabinoids in
Acknowledgment Cannabis sativa. Phytochemistry, 21, 67–69.
Kojoma, M., Seki, H., Yoshida, S., & Muranaka, T. (2006). DNA poly-
We gratefully acknowledge Parijat Kusari for critically reading this
morphisms in the tetrahydrocannabinolic acid (THCA) synthase
manuscript.
gene in “drug-type” and “fiber-type” Cannabis sativa L. Forensic
Science International, 159, 132–140.
Konishi, T., Shinohara, K., Yamada, K., & Sasaki, Y. (1996). Acetyl-CoA
References carboxylase in higher plants: Most plants other than gramineae
Austin, M. B., Bowman, M. E., Ferrer, J. L., Schröder, J., & Noel, J. P. have both the prokaryotic and the eukaryotic forms of this enzyme.
(2004). An aldol switch discovered in stilbene synthases mediates Plant and Cell Physiology, 37, 117–122.
cyclization specificity of type III polyketide synthases. Chemistry Kutchan, T. M., & Dittrich, H. (1995). Characterization and mechanism
and Biology, 11, 1179–1194. of the berberine bridge enzyme, a covalently flavinylated oxidase
Burke, C. C., Wildung, M. R., & Croteau, R. (1999). Geranyl diphos- of benzophenanthridine alkaloid biosynthesis in plants. Journal of
phate synthase: Cloning, expression, and characterization of this Biological Chemistry, 270, 24475–24481.
prenyltransferase as a heterodimer. Proceedings of the National Acad- Larkin, M. A., Blackshields, G., Brown, N. P., Chenna, R., McGettigan,
emy of Sciences of the United States of America, 96, 13062–13067. P. A., McWilliam, H., Valentin, F., Wallace, I. M., Wilm, A., Lopez,
Chen, H., Kim, H. U., Weng, H., & Browse, J. (2011). Malonyl-CoA syn- R., Thompson, J. D., Gibson, T. J., & Higgins, D. G. (2007). Clustal W
thetase, encoded by ACYL ACTIVATING ENZYME13, is essential for and Clustal X version 2.0. Bioinformatics, 23, 2947–2948.
growth and development of Arabidopsis. The Plant Cell, 23, 2247–2262. Marks, M. D., Tian, L., Wenger, J. P., Omburo, S. N., Soto-Fuentes, W.,
Crombie, L., Ponsford, R., Shani, A., Yagnitinsky, B., & Mechoulam, R. He, J., Gang, D. R., Weiblen, G. D., & Dixon, R. A. (2009). Identifica-
(1968). Hashish components. Photochemical production of canna- tion of candidate genes affecting ∆9-tetrahydrocannabinol biosynthe-
bicyclol from cannabichromene. Tetrahedron Letters, 9, 5771–5772. sis in Cannabis sativa. Journal of Experimental Botany, 60, 3715–3726.
de Meijer, E. (2011). Cannabis sativa plants rich in cannabichromene and Morimoto, S., Komatsu, K., Taura, F., & Shoyama, Y. (1997). Enzymo-
its acid, extracts thereof and methods of obtaining extracts there- logical evidence for cannabichromenic acid biosynthesis. Journal of
from. U.S. Patent No. 2011/0098348 A1 Natural Products, 60, 854–857.
de Meijer, E., Hammond, K., & Micheler, M. (2009). The inheritance of Morimoto, S., Komatsu, K., Taura, F., & Shoyama, Y. (1998). Purifica-
chemical phenotype in Cannabis sativa L.(III): variation in cannabi- tion and characterization of cannabichromenic acid synthase from
chromene proportion. Euphytica, 165, 293–311. Cannabis sativa. Phytochemistry, 49, 1525–1529.
de Zeeuw, R. A., Wijsbek, J., Breimer, D. D., Vree, T. B., van Ginneken, Page, J.E. & Boubakir, Z. (2012). Aromatic prenyltransferase from
C. A., & van Rossum, J. M. (1972). Cannabinoids with a propyl side Cannabis. U.S. Patent No. 2012/0144523 A1.
chain in Cannabis: occurrence and chromatographic behaviour. Sci- Page, J.E. & Gagne, S. (2013). Genes and proteins for aromatic
ence, 175, 778–779. polyketide synthesis. U.S. Patent No. 2013/0067619 A1.
Eisenreich, W., Bacher, A., Arigoni, D., & Rohdich, F. (2004). Biosyn- Page, J.E. & Stout, J.M. (2013). Genes and proteins for alkanoyl-CoA
thesis of isoprenoids via the non-mevalonate pathway. Cellular and synthesis. Patent No. WO 2013/006953 A1.
Molecular Life Sciences, 61, 1401–1426. Raharjo, T. J., Chang, W. -T., Choi, Y. H., Peltenburg-Looman, A. M. G.,
ElSohly, M., & Slade, D. (2005). Chemical constituents of marijuana: the & Verpoorte, R. (2004). Olivetol as product of a polyketide synthase
complex mixture of natural cannabinoids. Life Sciences, 78, 539–548. in Cannabis sativa L. Plant Science, 166, 381–385.
Fellermeier, M., Eisenreich, W., Bacher, A., & Zenk, M. H. (2001). Bio- Razdan, R. K., Puttick, A. J., Zitko, B. A., & Handrick, G. R. (1972).
synthesis of cannabinoids. Incorporation experiments with 13C- Hashish VI1: Conversion of (-)-∆1(6)-tetrahydrocannabinol to
labeled glucoses. European Journal of Biochemistry, 268, 1596–1604. (-)-∆1(7)-tetrahydrocannabinol. Stability of (-)-∆1 and (-)-∆1(6)-tetrahy-
Fellermeier, M., & Zenk, M. H. (1998). Prenylation of olivetolate by a drocannabinols. Experientia, 28, 121–122.
hemp transferase yields cannabigerolic acid, the precursor of tetra- Robert, X., & Gouet, P. (2014). Deciphering key features in protein
hydrocannabinol. FEBS Letters, 427, 283–285. structures with the new ENDscript server. Nucleic Acids Research,
Fischedick, J. T., Hazekamp, A., Erkelens, T., Choi, Y. H., & Verpoorte, 42, W320–W324.
R. (2010). Metabolic fingerprinting of Cannabis sativa L., cannabi- Rotherham, D., & Harbison, S. A. (2011). Differentiation of drug and
noids and terpenoids for chemotaxonomic and drug standardiza- non-drug Cannabis using a single nucleotide polymorphism (SNP)
tion purposes. Phytochemistry, 71, 2058–2073. assay. Forensic Science International, 207, 193–197.
Gagne, S. J., Stout, J. M., Liu, E., Boubakir, Z., Clark, S. M., & Page, J. E. Schultes, R., Klein, W., Plowman, T., & Lockwood, T. (1974). Cannabis:
(2012). Identification of olivetolic acid cyclase from Cannabis sativa An example of taxonomic neglect. Botanical Museum Leaflets,
reveals a unique catalytic route to plant polyketides. Proceedings of Harvard University, 23, 337–367.

I.  Setting the scene, botanical, general and international aspects


REFERENCES 23
Shoyama, Y., Fujita, T., Yamauchi, T., & Nishioka, I. (1968). Cannabi- Taura, F., Morimoto, S., & Shoyama, Y. (1995a). Cannabinerolic acid, a
chromenic acid, a genuine substance of cannabichromene. Chemical cannabinoid from Cannabis sativa. Phytochemistry, 39, 457–458.
and Pharmaceutical Bulletin, 16, 1157–1158. Taura, F., Morimoto, S., & Shoyama, Y. (1995b). First direct evidence
Shoyama, Y., Hirano, H., & Nishioka, I. (1984). Biosynthesis of propyl for the mechanism of ∆1-tetrahydrocannabinolic acid biosynthesis.
cannabinoid acid and its biosynthetic relationship with pentyl and Journal of the American Chemical Society, 117, 9766–9767.
methyl cannabinoid acids. Phytochemistry, 23, 1909–1912. Taura, F., Morimoto, S., & Shoyama, Y. (1996). Purification and char-
Shoyama, Y., Oku, R., Yamauchi, T., & Nishioka, I. (1972). Cannabis. acterization of cannabidiolic acid synthase from Cannabis sativa L.
VI. Cannabicyclolic acid. Chemical and Pharmaceutical Bulletin, 20, Journal of Biological Chemistry, 271, 17411–17416.
1927–1930. Taura, F., Sirikantaramas, S., Shoyama, Y., Shoyama, Y., & Morimoto,
Shoyama, Y., Tamada, T., Kurihara, K., Takeuchi, A., Taura, F., Arai, S., S. (2007a). Phytocannabinoids in Cannabis sativa: recent studies on
Blaber, M., Shoyama, Y., Morimoto, S., & Kuroki, R. (2012). Struc- biosynthetic enzymes. Chemistry and Biodiversity, 4, 1649–1663.
ture and function of ∆1-tetrahydrocannabinolic acid (THCA) syn- Taura, F., Sirikantaramas, S., Shoyama, Y., Yoshikai, K., Shoyama, Y., &
thase, the enzyme controlling the psychoactivity of Cannabis sativa. Morimoto, S. (2007b). Cannabidiolic-acid synthase, the chemotype-
Journal of Molecular Biology, 423, 96–105. determining enzyme in the fiber-type Cannabis sativa. FEBS Letters,
Sirikantaramas, S., Morimoto, S., Shoyama, Y., Ishikawa, Y., Wada, Y., 581, 2929–2934.
& Taura, F. (2004). The gene controlling marijuana psychoactivity: Taura, F., Tanaka, S., Taguchi, C., Fukamizu, T., Tanaka, H., Shoyama,
molecular cloning and heterologous expression of ∆1-tetrahydro- Y., & Morimoto, S. (2009). Characterization of olivetol synthase, a
cannabinolic acid synthase from Cannabis sativa L. Journal of Biologi- polyketide synthase putatively involved in cannabinoid biosyn-
cal Chemistry, 279, 39767–39774. thetic pathway. FEBS Letters, 583, 2061–2066.
Sirikantaramas, S. T. S., Morimoto, S., & Shoyama, Y. (2007). Recent van Bakel, H., Stout, J., Cote, A., Tallon, C., Sharpe, A., Hughes, T., &
advances in Cannabis sativa research: biosynthetic studies and its Page, J. (2011). The draft genome and transcriptome of Cannabis sa-
potential in biotechnology. Current Pharmaceutical Biotechnology, 8, tiva. Genome Biology, 12, R102.
237–243. Yamamoto, H., Kimata, J., Senda, M., & Inoue, K. (1997). Dimethylallyl
Stout, J. M., Boubakir, Z., Ambrose, S. J., Purves, R. W., & Page, J. E. diphosphate: Kaempferol 8-dimethylallyl transferase in Epimedium
(2012). The hexanoyl-CoA precursor for cannabinoid biosynthesis is diphyllum cell suspension cultures. Phytochemistry, 44, 23–28.
formed by an acyl-activating enzyme in Cannabis sativa trichomes. Yamamoto, H., Senda, M., & Inoue, K. (2000). Flavanone 8-dimethylal-
The Plant Journal, 71, 353–365. lyltransferase in Sophora flavescens cell suspension cultures. Phyto-
Suzuki, Y., Kurano, M., Esumi, Y., Yamaguchi, I., & Doi, Y. (2003). chemistry, 54, 649–655.
Biosynthesis of 5-alkylresorcinol in rice: Incorporation of a puta- Yang, X., Matsui, T., Kodama, T., Mori, T., Zhou, X., Taura, F., Noguchi,
tive fatty acid unit in the 5-alkylresorcinol carbon chain. Bioorganic H., Abe, I., & Morita, H. (2016). Structural basis for olivetolic acid
Chemistry, 31, 437–452. formation by a polyketide cyclase from Cannabis sativa. FEBS Jour-
Taguchi, C., Taura, F., Tamada, T., Shoyama, Y., Tanaka, H., Kuroki, R., nal, 283, 1088–1106.
& Morimoto, S. (2008). Crystallization and preliminary X-ray dif- Zhao, P., Inoue, K., Kouno, I., & Yamamoto, H. (2003). Characterization
fraction studies of polyketide synthase-1 (PKS-1) from Cannabis of leachianone G 2”-dimethylallyltransferase, a novel prenyl side-
sativa. Acta Crystallographica. Section F, Structural biology and crystal- chain elongation enzyme for the formation of the lavandulyl group
lization communications, 64, 217–220. of sophoraflavanone G in Sophora flavescens Ait. cell suspension cul-
Taura, F., Dono, E., Sirikantaramas, S., Yoshimura, K., Shoyama, Y., & tures. Plant Physiology, 133, 1306–1313.
Morimoto, S. (2007). Production of ∆1-tetrahydrocannabinolic acid Zuurbier, K. W. M., Fung, S. -Y., Scheffer, J. J. C., & Verpoorte, R. (1998).
by the biosynthetic enzyme secreted from transgenic Pichia pastoris. In-vitro prenylation of aromatic intermediates in the biosynthesis
Biochemical and Biophysical Research Communications, 361, 675–680. of bitter acids in Humulus lupulus. Phytochemistry, 49, 2315–2322.

I.  Setting the scene, botanical, general and international aspects

You might also like