ChapterPDF Available

Abstract and Figures

Alkaloids constitute a very chemically diverse group of plant natural products. These molecules, of which about 12,000 have been identified, share the common presence of a nitrogen atom in a heterocyclic ring. Because of their toxicity, many alkaloids are used for human health benefit. This review presents recent progresses in the elucidation of the biosynthetic pathways of four major classes of alkaloids—monoterpene indole alkaloids (MIA), benzylisoquinoline alkaloids (BIA), tropane and nicotine alkaloids (TNA) and purine alkaloids (PA)—and in the understanding of the regulatory processes involved. Three main approaches are discussed, namely the complex spatial organisation of some of these biosynthetic pathways, the crystallisation and modelling of some of the biosynthetic enzymes, and the transcription factor regulatory networks involved in alkaloid biosynthesis. Some applications of these discoveries for metabolic engineering strategies are also presented.
Content may be subject to copyright.
Chapter 8
Biosynthesis and Regulation of Alkaloids
G. Guirimand, V. Courdavault, B. St-Pierre, and V. Burlat
8.1 Introduction
Plant secondary metabolites, also commonly named plant natural products, nowadays
encompassed about 100,000 chemically identified, low molecular weight compounds.
These molecules are commonly synthesised in a plant-, organ- and even cell-specific
manner. Alkaloids constitute a very chemically diverse group of secondary metabo-
lites with an estimated 12,000 different molecules sharing as a unique common
feature the presence of a nitrogen atom within a heterocyclic ring. Many alkaloids
are toxic, in agreement with some function in plant defence. This toxicity has long
been understood by humans, and some alkaloids are very well known for their human
health benefit, leading to the development of many recent research projects towards
understanding the architecture and the regulation of their long and complex biosyn-
thetic pathways. The purpose of this chapter is to give an overview of some of the
major advances obtained recently with the most prominent model species, including
the architecture and the spatial organisation of biosynthetic pathways, the crystal-
lisation and modelling of alkaloid biosynthetic enzymes and the transcription factor
regulatory networks of alkaloid biosynthesis. The ultimate goal of such research is the
application of the discoveries to develop metabolic engineering strategies to over-
come the usually very low yield of production for these biomolecules. In the mean-
time, some peculiarities of plant physiology have been elucidated in these highly
G. Guirimand, V. Courdavault, and B. St-Pierre
Universite
´Franc¸ois Rabelais de Tours, EA 2106 “Biomole
´cules et Biotechnologies Ve
´ge
´tales”,
IFR 135, “Imagerie Fonctionnelle”, 37200, Tours, France
e-mail: vincent.courdavault@univ-tours.fr
V. Burlat
Surfaces Cellulaires et Signalisation chez les Ve
´ge
´taux, UMR 5546 CNRS - UPS - Universite
´de
Toulouse, Po
ˆle de Biotechnologie Ve
´ge
´tale, 24 chemin de Borde-Rouge, BP 42617 Auzeville,
31326, Castanet-Tolosan, France
Universite
´Franc¸ois Rabelais de Tours, EA 2106 “Biomole
´cules et Biotechnologies Ve
´ge
´tales”,
IFR 135, “Imagerie Fonctionnelle”, 37200, Tours, France
e-mail: burlat@scsv.ups-tlse.fr
EC. Pua and M.R. Davey (eds.),
Plant Developmental Biology – Biotechnological Perspectives: Volume 2,
DOI 10.1007/978-3-642-04670-4_8, #Springer-Verlag Berlin Heidelberg 2010
139
specialised plant species, illustrating the fundamental interest of studying such sec-
ondary metabolic pathways beyond their primary interest for industrial application.
8.2 Chemical Diversity and Biosynthesis
Alkaloids are classified in several families that present totally different biosynthetic
pathways. Four major families, for which the biosynthesis and the regulation are
more particularly studied, are discussed in this chapter, these being monoterpene
indole alkaloids (MIA), benzylisoquinoline alkaloids (BIA), tropane and nicotine
alkaloids (TNA) and purine alkaloids (PA; Fig. 8.1, Table 8.1). Despite their
chemical diversity, alkaloids share the fact that they originate commonly from
primary metabolites such as amino acids or bases (Fig. 8.1). Except for Nicotiana
tabacum, no genome sequencing project exists for the major alkaloid-producing
plants. Therefore, most of the enzymatic steps have been identified using classical
biochemical and molecular biology studies. The ongoing elucidation of some
biosynthetic pathways illustrates the recruitment of enzymes belonging to recurrent
multigene families such as cytochrome P450 monooxygenases, acetyl transferases
or methyltransferases. Recent expressed sequence tags (EST) transcriptomic pro-
jects focusing on MIA-producing species (Murata et al. 2006,2008; Rischer et al.
2006; Shukla et al. 2006), BIA-producing species (Morishige et al. 2002; Carlson
et al. 2006; Ziegler et al. 2006; Kato et al. 2007; Zulak et al. 2007) and TNA-
producing species (Li et al. 2006) have been helpful for the identification of some
missing enzymatic steps, and should accelerate this process in the near future
(Table 8.2). Metabolic profiling studies of the major alkaloid-producing species
have also been performed recently (Dra
¨ger 2002; Yamazaki et al. 2003; Schmidt
et al. 2007; Hagel et al. 2008), sometimes in association with transcriptomic studies
(Rischer et al. 2006; Ziegler et al. 2006; Zulak et al. 2007; Hagel and Facchini
2008). Together, these approaches now allow tremendous progress in understand-
ing these complex alkaloid biosynthetic pathways.
The purpose of this section is more to give an overlook and a link to references,
including major recent reviews on the different biosynthetic pathways, than to give
an exhaustive detailed description of all these complicated biochemical processes.
8.2.1 Biosynthesis of Monoterpene Indole Alkaloids (MIA)
The approximately 2,000 MIA chemical structures described so far are widespread
in a large number of plant species (Ziegler and Facchini 2008). Some of these
molecules are of interest to human health, such as the anticancer drugs vinblastine
and vincristine and the antihypertensive drug ajmalicine specifically produced in
Catharanthus roseus, the anti-arrythmic ajmaline produced in Rauvolfia serpen-
tina, or the anticancer compound camptothecin produced mostly in Camptotheca
acuminata (Fig. 8.1, Table 8.1). These molecules are part of the large array of MIA
140 G. Guirimand et al.
OPO3--
OH
OH
H3C
OH
NH2
N
H
COOH
N
N
H
CO2CH3
CO2CH3
HN
NO-COCH3
HO
H3CH
H
MeO
OH
NH2
COOH
HO
O
O
O
O
N+
CH3
HO
O
HO
HNCH3
H
NH2
N
H
NH
COOH
H2N
O
HCH2OH
O
N
CH3
O
N
N
CH3
H
O
HOH2C
OH OH
N
N
N
H
O
O
N
N
HN
HN
N
O
O
CH3
CH3
CH3
N
N
N
N
O
O
CH3
H3C
MEP argininetryptophan tyrosine
xanthosine
theobromine
caffeinescopolamine nicotinesanguinarine morphinevinblastine
MIA BIA TNA PA
Fig. 8.1 Chemical structures of precursors and examples of alkaloids from the four families where the major advances in the elucidation and regulation of the
biosynthetic pathways have been obtained. MIA, monoterpene indole alkaloids; BIA, benzylisoquinoline alkaloids; TNA, tropane and nicotine alkaloids; PA, purine
alkaloids; MEP, 2-C-methyl-D-erythritol 4-phosphate
8 Biosynthesis and Regulation of Alkaloids 141
Table 8.1 Sources, chemical diversity and biological activity of four major families of alkaloids
Alkaloid family Family (plant
species)
Active compounds Pharmacological activity Target
Monoterpene indole alkaloids
(MIA)
Apocynaceae
Catharanthus roseus Vinblastine,
vincristine
Anticancer Tubulin
Ajmalicine Antihypertensive aAdrenergic receptor
Rauvolfia serpentina Ajmaline Anti-arrythmic Na
+
channels
Nyssaceae
Camptotheca
acuminata
Camptothecin Anticancer DNA topoisomerase I
Benzylisoquinoline alkaloids
(BIA)
Papaveraceae
Papaver somniferum Codeine Antitussive, analgesic Nicotinic acetylcholine (nACh) receptor
Morphine Analgesic, narcotic m3 Opioid receptor
Papaverine Spasmolytic, vasodilators c-AMP phosphodiesterase
Eschscholzia
californica
Sanguinarine Antibacterial, proapoptotic FtsZ (bacterial cytokinesis), mitoch.
pathway
Ranunculaceae
Thalictrum flavum Berberine Antibacterial, antimicrobial DNA
Coptis japonica Berberine,
sanguinarine
Tropane and nicotine alkaloids
(TNA)
Solanaceae
Hyoscyamus niger Hyoscyamine Anticholinergic, narcotic,
myorelaxant
Muscarinic receptor
Datura stramonium Scopolamine Anticholinergic, narcotic,
myorelaxant
Muscarinic receptor
Atropa belladonna Hyoscyamine Anticholinergic, narcotic,
myorelaxant
Muscarinic receptor
Nicotiana tabacum Nicotine Neurostimulant, insecticide nACh receptor
Purine alkaloids (PA) Coffeeae
Coffea arabica Caffeine Central nervous system
stimulant
Adenosine A
1
&A
2A
receptors;
phosphodiesterase
Theaceae
142 G. Guirimand et al.
Camellia sinensis Caffeine,
theophylline
Central nervous system
stimulant
Adenosine A
1
&A
2A
receptors;
phosphodiesterase
Byttnerioideae
Theobroma cacao Theobromine,
caffeine
Central nervous system
stimulant
Adenosine A
1
&A
2A
receptors;
phosphodiesterase
8 Biosynthesis and Regulation of Alkaloids 143
Table 8.2 Molecular resources for alkaloid biosynthesis
Family (plant species) NCBI hints
(nt seq)
NCBI hints
(EST)
Alkaloid biosynthetic enzymes (GenBank access numbers, full enzyme name and
abbreviation)
Alkaloid family: monoterpene indole alkaloids (MIA)
Apocynaceae
Catharanthus roseus 20,820 19,899 25 [CAA09804(1-deoxyxylulose 5-phosphate synthase: dxs); ABI35993(dxs2);
AAF65154(1-deoxy-D-xylulose-5-phosphate reductoisomerase: dxr); ACI16377 (4-
diphosphocytidyl-methylerythritol 2-phosphate synthase: cms); ABI35992(4-
diphosphocytidyl-2-C-methyl-D-erythritol kinase: cmk); AAF65155(2C-methyl-D-
erythritol 2,4-cyclodiphosphate synthase: mecs); AAO24774(4-hydroxy-3-methyl-2-
butenyl diphosphate synthase: hds); ABI30631(1-hydroxy-2-methyl-butenyl 4-
diphosphate reductase: hdr); ABW98669(isopentenyl pyrophosphate:dimethylallyl
pyrophosphate isomerase: idi); ACC77966(geranyl pyrophosphate synthase: gpps);
CAC80883(CYP76B6=geraniol 10-hydroxylase: g10h); Q05001(NADPH
cytochrome P450 reductase: cpr); AAQ55962(10-hydroxygeraniol oxidoreductase:
10hgo); ABW38009(loganic acid methyltransferase: lamt); AAA33106
(CYP72A1=secologanin synthase: sls); CAC29060(anthranilate synthase asubunit:
asa); AAA33109(tryptophan decarboxylase: tdc); CAA43936(strictosidine synthase:
str); AAF28800(strictosidine b-D-glucosidase: sgd); AAO13736(minovincinine 19-
hydroxy-O-acetyltransferase: mat); CAB56503(CYP71D12=tabersonine 16-
hydroxylase: t16h); ABR20103(16-hydroxytabersonine O-methyltransferase:
16omt); O04847(desacetoxyvindoline-4-hydroxylase: d4h); AAC99311
(deacetylvindoline 4-O-acetyltransferase: dat); CAJ84723 (peroxidase 1: prx1)]
Rauvolfia serpentina 17 - 7 [CAA44208(strictosidine synthase: str1); CAC83098(strictosidine b-D-glucosidase:
sgd1); AAF22288(polyneuridine aldehyde esterase: pnae); Q70PR7(vinorine
synthase: vs); AAF03675(raucaffricine b-D-glucosidase: rg); AAW88320
(acetylajmalan acetylesterase: aae); AAX11684(perakine reductase: pr)]
Nyssaceae
Camptotheca acuminata 31 - 10 [ABC86579(1-deoxy-D-xylulose-5-phosphate reductoisomerase: dxr); O48964
(isopentenyl pyrophosphate:dimethylallyl pyrophosphate isomerase: idi1); O48965
(idi2); AAV64030(5-enolpyruvylshikimate 3-phosphate synthase: epsps);
AAU84988(anthranilate synthase asubunit: asa1); AAU84989(asa2); AAB97526
(tryptophan synthase bsubunit: tsb); AAB39708(tryptophan decarboxylase: tdc1);
AAB39709(tdc2); AAQ20892(10-hydroxygeraniol oxidoreductase: 10hgo)]
144 G. Guirimand et al.
Alkaloid family: benzylisoquinoline alkaloids (BIA)
Papaveraceae
Papaver somniferum 20,458 20,340 16 [P54768(tyrosine/DOPA decarboxylase: tydc1); P54770(tydc3); AAX56303(S-
norcoclaurine synthase: ncs1); AAX56304.(ncs2); AAQ01669(norcoclaurine 6-O-
methyltransferase: 6omt); AAP45316(coclaurine N-methyltransferase: cnmt);
AAF61400(CYP80B1 renamed CYP80B3=N-methylcoclaurine 30-hydroxylase);
AAP45313(30-hydroxy-N-methylcoclaurine 40-O-methyltransferase: 40omt1);
AAP45314(40omt2); AAQ01668(reticuline 7-O-methyltransferase: 7omt); P93479
(berberine bridge enzyme: bbe); AAY79177(tetrahydroprotoberberine-cis-N-
methyltransferase: tnmt); ABR14720(salutaridine synthase: ss); ABC47654
(salutaridine reductase: salr); Q94FT4(salutaridinol 7-O-acetyltransferase: salat);
AAF13738(codeinone reductase: cor1)]
Eschscholzia californica 9,161 9,083 5 [P30986(berberine bridge enzyme: bbe1); BAE79723(reticuline-7-O-
methyltransferase: 7omt); O64899(CYP80B1 renamed CYP80B3=N-
methylcoclaurine 30-hydroxylase); AAC39454(CYP82B1=N-methylcoclaurine 30-
hydroxylase); O64900(CYP80B2=N-methylcoclaurine 30-hydroxylase)]
Ranunculaceae
Thalictrum flavum
10 - 9 [AAG60665(tyrosine/dopa decarboxylase: tydc1); AAR22502(norcoclaurine
synthase: ncs); AAU20765(norcoclaurine 6-O-methyltransferase: 6omt); AAU20766
(coclaurine N-methyltransferase: cnmt); AAU20767(CYP80B3=N-
methylcoclaurine 30-hydroxylase); AAU20768(30-hydroxy-N-methylcoclaurine 40-
O-methyltransferase: 40omt); AAU20769(berberine bridge enzyme: bbe);
AAU20770(scoulerine 9-O-methyltransferase: somt); AAU20771
(CYP719A1=canadine synthase)]
Coptis japonica 135 37 9 [BAF45337(norcoclaurine synthase: ncs); Q9LEL6(norcoclaurine 6-O-
methyltransferase: 6omt); Q948P7(coclaurine N-methyltransferase: cnmt); Q9LEL5
(30-hydroxy-N-methylcoclaurine 40-O-methyltransferase: 40omt); Q39522(scoulerine
9-O-methyltransferase: somt); BAB68769(CYP719A=methylenedioxy bridge-
forming enzyme); Q8H9A8(columbamine O-methyltransferase: coomt); BAB12433
(CYP80B2=N-methylcoclaurine-30-hydroxylase); BAF80448
(CYP80G2=corytuberine synthase)]
(continued)
8 Biosynthesis and Regulation of Alkaloids 145
Table 8.2 (continued)
Family (plant species) NCBI hints
(nt seq)
NCBI hints
(EST)
Alkaloid biosynthetic enzymes (GenBank access numbers, full enzyme name and
abbreviation)
Alkaloid family: tropane and nicotine alkaloids (TNA)
Solanaceae
Hyoscyamus niger 23 - 5 [BAA82263(putrescine N-methyltransferase: pmt); BAA85844(tropinone reductase:
tr1); P50164(tr2); ABD39696(CYP80F1=littorine mutase/monooxygenase); P24397
(hyoscyamine 6-b-hydroxylase: h6h)]
Atropa belladonna 58 - 3 [BAA82264(putrescine N-methyltransferase: pmt1); BAA82262(pmt2); BAA78340
(hyoscyamine 6-b-hydroxylase: h6h)]
Datura stramonium 31 - 6 [P50134(ornithine decarboxylase: odc); CAB64599(arginine decarboxylase: adc1);
CAE47481(putrescine N-methyltransferase: pmt); P50162(tropinone reductase: tr1);
P50163(tr2); P50165(trh)]
Nicotiana tabacum 1,673,039 240,440 4 [AAQ14852(ornithine decarboxylase: odc); AAF14881(putrescine N-
methyltransferase: pmt); ABI93948(methylputrescine oxidase: mpo1); DQ131886
(CYP82E4v1=nicotine N-demethylase)]
Coffeeae
Coffea arabica 6,079 1,577 10 [BAB39215(xanthosine methyltransferase: xmt1); BAC75665(xmt2); BAB39216(7-
methylxanthine N-methyltransferase: mxmt1); BAC75664(mxmt2); BAC75663(3,7-
dimethylxanthine N-methyltransferase: dxmt1); BAC43756(theobromine synthase:
cts1); BAC43757(cts2); BAC43755(7-methylxanthosine synthase: xrs1); BAC43760
(caffeine synthase: ccs1); BAC43761(ctcs7)]
Camellia sinensis 3,732 3,288 1 [ABP98983(caffeine synthase: tcs)]
Byttnerioideae
Theobroma cacao 7,469 6,790 1 [BAE79730(caffeine synthase: bcs1)]
146 G. Guirimand et al.
that a single plant species is able to produce. For example, there are more than 130
MIA in C. roseus (van der Heijden et al. 2004). The elucidation of MIA biosyn-
thetic pathways in the three mentioned species has undergone major recent progress
with the identification of 42 clones corresponding to 31 enzymatic steps (Table 8.2).
In C. roseus alone, 27 enzymatic steps have been studied (25 cDNA clones and two
additional enzymatic activities with no assigned clone). In these species, the path-
ways share a common origin with strictosidine synthase (STR), catalysing the
condensation of the indole precursor tryptamine with the terpenoid precursor
secologanin to form the first MIA, strictosidine. The upstream biosynthesis of the
indole precursor derived from the shikimate pathway via tryptophan, and of the
terpenoid precursor originating from the methyl erythritol phosphate (MEP) path-
way, is also shared within these plant species. Strictosidine b-glucosidase (SGD),
catalysing the deglucosylation of strictosidine, is the last common enzyme for the
biosynthesis of 2,000 MIA, since the resulting aglycon is the starting point for many
different species-specific lateral MIA pathways, with the observed possibility for a
given species to harbour more than one of these pathways (e.g. C. roseus, reviewed
in van der Heijden et al. 2004). Many enzymatic steps are yet to be discovered.
More details on these pathways and the identified enzymatic steps are available in
recent reviews (Lorence and Nessler 2004; van der Heijden et al. 2004; Sto
¨ckigt and
Panjikar 2007; Mahroug et al. 2007; Ziegler and Facchini 2008).
8.2.2 Biosynthesis of Benzylisoquinoline Alkaloids (BIA)
BIA constitute a diverse class of more than 2,500 compounds with, for some of
them, potent pharmacological properties and socio-economic importance. In opium
poppy (Papaver somniferum) alone, more than 80 alkaloids have been identified.
BIA have also been widely studied in other members of the Papaveraceae, such as
Eschscholzia californica, or members of the Ranunculaceae such as Thalictrum
flavum and Coptis japonica (Table 8.1). The biosynthesis of BIA starts with the
condensation of two tyrosine derivatives to produce (S)-norcoclaurine. Four char-
acterised enzymatic steps are necessary to produce (S)-reticuline, the central
precursor of the five major BIA subpathways, leading to palmatine, berberine,
sanguinarine, laudanine and codeine/morphine respectively. Overall, regardless of
the plant model species, a total of 39 clones and 19 enzymatic steps have been
characterised (Table 8.2), making the BIA biosynthetic pathway one of the best
characterised plant natural product complex pathways (reviewed in Ziegler and
Facchini 2008).
8.2.3 Biosynthesis of Tropane and Nicotine Alkaloids (TNA)
TNA are widely used in medicine as nonselective muscarinic antagonists affecting
peripheral and central nervous systems (Table 8.1). This alkaloid family is found
8 Biosynthesis and Regulation of Alkaloids 147
mainly in Solanaceae species, such as Hyoscyamus niger,Datura stramonium,
Atropa belladonna or Nicotiana tabacum, and accounts for more than 200 different
compounds (Dra
¨ger 2002,2006; Oksman-Caldentey 2007). TNA are amongst the
most studied alkaloids and their pharmacological effects are well documented.
However, their biosynthetic pathways are still only partially understood. TNA are
derived from the amino acids ornithine and arginine (Fig. 8.1), and the early
biosynthetic steps leading to N-methylputrescine formation have been elucidated
in several species (Table 8.2). The oxidative deamination of N-methylputrescine
leads to N-methylpyrrolium cation, which constitutes a branching point towards
tropane alkaloids and nicotine alkaloids respectively. The final steps of both path-
ways are partially characterised, but molecular information concerning the central
steps is still missing. Overall, seven enzymatic steps of the tropane alkaloid
biosynthetic pathway and four enzymatic steps of the nicotinic alkaloid biosyn-
thetic pathway have been characterised. More details on these pathways and on the
identified enzymatic steps are available in recent reviews (Dra
¨ger 2006; Oksman-
Caldentey 2007).
8.2.4 Biosynthesis of Purine Alkaloids (PA)
PA are natural products derived from purine nucleotides (Fig. 8.1; Ashihara et al.
2008). The main PA are caffeine and theobromine that affect the central nervous
system as neurostimulants and are synthesised by several plants, including Coffea
arabica,Camellia sinensis or Theobroma cacao (Table 8.1). The initial precursor
of PA is xanthosine, which is supplied by at least four different pathways. The main
caffeine biosynthetic pathway has four enzymatic steps, comprising three charac-
terised S-adenosylmethionine-dependant N-methylation reactions and one unchar-
acterised nucleosidase reaction (Table 8.2). However, according to structural
studies of N-methyltransferases involved in caffeine biosynthesis, it has been
suggested that the ribose hydrolysis could be performed by xanthosine 7 N-methyl-
transferase (McCarthy and McCarthy 2007). The detail of this pathway has been
reviewed recently (Ashihara et al. 2008).
8.3 Spatial Organisation of Alkaloid Biosynthesis
The spatial organisation of alkaloid biosynthesis has been recently investigated
extensively using in situ hybridisation and immunocytochemistry methods
(reviewed in De Luca and St-Pierre 2000; Kutchan 2005; Mahroug et al. 2007;
Ziegler and Facchini 2008). An astonishing complexity has been uncovered
showing multicellular organisations as a recurrent common feature. These types
of organisation implicate the necessity of intercellular translocation processes. In
C. roseus, a series of publications showed the sequential involvement of internal
phloem associated parenchyma (IPAP), epidermis and laticifers-idioblasts during
148 G. Guirimand et al.
a
c
ef
d
b
Fig. 8.2 Examples of multicellular organisation of alkaloid biosynthesis implicating the necessity
of intercellular translocation events. Six models of spatial organisation of alkaloid biosynthesis in
higher plants are presented in af.Inac, the greenorangered traffic light signal colour code
8 Biosynthesis and Regulation of Alkaloids 149
MIA biosynthesis in aerial organs (Fig. 8.2a; St-Pierre et al. 1999; Irmler et al.
2000; Burlat et al. 2004; Oudin et al. 2007). The IPAP cells harbour the expression
of genes involved in early steps of monoterpenoid biosynthesis, i.e. four MEP
pathway genes and geraniol 10-hydroxylase (G10H, CYP76B6) encoding the first
committed enzyme in monoterpenoid biosynthesis (Burlat et al. 2004; Oudin et al.
2007; Guirimand et al. 2009). The intermediate steps leading to the synthesis of the
two MIA precursors, tryptamine and secologanin, and to their subsequent conden-
sation to form the first MIA strictosidine, occur within the epidermis (St-Pierre et al.
1999; Irmler et al. 2000). Finally, the last two steps in the biosynthesis of vindoline,
one of the monomeric MIA precursors of the dimeric MIA vinblastine, are localised
to specialised laticifer-idioblast cells (St-Pierre et al. 1999). Recently, these results
were elegantly completed by an RT-PCR analysis of laser capture microdissected
C. roseus leaf cells (Murata and De Luca 2005) and by an EST analysis study of
epidermis-enriched fractions obtained using an original carborundum abrasion
technique (Levac et al. 2008; Murata et al. 2008). Altogether these results suggest
that, to ensure a continuity in the metabolic flux along the MIA pathway, it is
necessary to consider the translocation of an unknown monoterpenoid intermediate
from IPAP to epidermis, and the translocation of an unknown MIA intermediate
from epidermis to laticifer-idioblast cells (Fig. 8.2a). The identification of these
shuttling intermediates will necessitate the localisation of two subsequent enzy-
matic steps within two different cell types. Similarly, in the roots of several
solanaceous species, the shuttle of intermediate TNA between the pericycle and
the endodermis is suggested by the specific localisation of early and late enzymatic
steps (putrescine N-methyltransferase and hyoscyamine 6-hydroxylase respec-
tively) to the pericycle, and the specific expression of an intermediate enzymatic
step (tropinone reductase 1) within the neighbouring endodermis (Fig. 8.2b;
Fig. 8.2 (continued) illustrates (respectively) early, intermediate and late biosynthetic steps
sequentially occurring in different cell types along a given pathway. These three models implicate
the intercellular translocation of metabolic intermediates. In df, the yellow, magenta and blue
colour code shows the common localisation of all the mRNAs (yellow), enzymes (magenta) and
alkaloids (blue) from a given pathway in different cell types. The Facchini model (d) implicates
the translocation of enzymes from companion cells to sieve elements and the translocation of
alkaloids from sieve elements to laticifers. The models in dfillustrate that a same pathway can
be localised to different cell types according to the plant species and/or to the organ considered.
IPAP, internal phloem associated parenchyma; Ep, epidermis; Lat, laticifers; Pal Id, palisade
idioblasts; Sp Id, spongy parenchyma idioblasts; Pal Par, palisadic parenchyma; Sp par, spongy
parenchyma; ICP, internal conducting phloem; ECP, external conducting phloem; Xy, xylem; Pi,
pith; Per, pericycle; End, endodermis; Co, cortex; CC, companion cells; SE, sieve elements; Co
Par/Id, cortex parenchyma/Idioblasts; PAP, phloem associated parenchyma; PD, protoderm;
SAM, shoot apical meristem; VB, vascular bundles; RAM, root apical meristem; Dev End,
developing endodermis; Mat End, mature endodermis. These models were drawn from results
taken from the reports by aSt-Pierre et al. (1999), Irmler et al. (2000), Burlat et al. (2004) and
Oudin et al. (2007); bHashimoto et al. (1991), Nakajima and Hashimoto (1999) and Suzuki et al.
(1999a,b); cBock et al. (2002) and Weid et al. (2004); dBird et al. (2003) and Samanani et al.
(2005); e,fSamanani et al. (2005)
<
150 G. Guirimand et al.
Hashimoto et al. 1991; Nakajima and Hashimoto 1999; Suzuki et al. 1999a,b). In
this case, given the proximity of conducting tissues, it is hypothesised that the
precursor of TNA (arginine) could come from the phloem, and that final TNA
products, such as scopolamine, could flow to the aerial parts of the plant through the
xylem (Fig. 8.2b; Nakajima and Hashimoto 1999). Four models of the spatial
organisation of BIA synthesis in opium poppy and in a species (T. flavum)of
Ranunculaceae also illustrate the complexity of these compartmentations. In
opium poppy, two different models have been proposed (Fig. 8.2c, Kutchan
2005; Fig. 8.2d, Facchini and St-Pierre 2005). In one model, a rationale similar to
that described for MIA and TNA biosynthesis is proposed, since a cell-specific
separation occurs between early and late steps of different branches of the BIA
pathway. According to immunolocalisation studies performed on five biosynthetic
enzymes, this model proposed that BIA synthesis commonly starts in the phloem
parenchyma with the synthesis of (S)-reticuline. Different situations are then
observed for three BIA subpathways leading to laudanine, codeine/morphine and
scoulerine respectively. The synthesis of laudanine appears to occur within the
same phloem parenchyma cells, whereas the biosynthesis of scoulerine is localised
to leaf idioblasts and root cortex cells. Finally, an intermediate step in the codeine
pathway is also localised in phloem parenchyma, whereas the final step occurs in
laticifers (Fig. 8.2c; Bock et al. 2002; Weid et al. 2004; reviewed in Kutchan 2005).
However, Facchini and St-Pierre (2005) proposed a totally different type of com-
partmentation in the same species with a so-called tale of three cell types, implicat-
ing the transcription of seven BIA biosynthesis genes within companion cells,
the immunolocalisation of the corresponding enzymes to the sieve elements, and
the accumulation of BIA in neighbouring laticifers (Fig. 8.2d; Bird et al. 2003;
Samanani et al. 2006; reviewed in Ziegler and Facchini 2008). The discrepancy
between both models has been attributed tentatively to differences in cultivars and/
or in developmental stages. The last example of BIA synthesis in T. flavum
illustrates that several steps of a pathway may be compartmentalised in a different
manner within different plant species (compare P. somniferum, Fig. 8.2c, d and
T. flavum, Fig. 8.2e, f) and even within different organs (rhizomes versus roots) of
the same plant species (Fig. 8.2e, f; Samanani et al. 2005).
Alkaloid biosynthesis also involves subcellular compartmentation of biosyn-
thetic enzymes, as is the case for most natural product biosynthetic pathways.
Beside cytosol, organelles such as ER (either lumen or membranes), plastids (either
stroma or thylakoids), mitochondria and vacuoles have been implicated in various
alkaloid biosynthetic pathways (reviewed in Facchini and St-Pierre 2005; Mahroug
et al. 2007; Ziegler and Facchini 2008). These results are based on in silico analysis
of biosynthetic enzyme sequences and also on more formal experimental evidence.
Part of these experiments corresponded to density gradient analysis, and a few
examples of direct localisation of enzymes by immunogold (McKnight et al. 1991;
Bock et al. 2002; Alcantara et al. 2005; Samanani et al. 2006; Oudin et al. 2007;
Guirimand et al. 2009) and by GFP-fusion image analyses (Bird and Facchini 2001;
Costa et al. 2008; Guirimand et al. 2009) have been published. This emphasizes that
a systematic (re)evaluation of the subcellular localisation of all the enzymes
8 Biosynthesis and Regulation of Alkaloids 151
available in a given alkaloid pathway constitutes a future challenge that should
enable the drawing of more complete spatial compartmentation models that inte-
grate both cellular and subcellular levels.
Together, these results suggest the recurrent necessity of transmembrane and
intercellular translocation processes during alkaloid biosynthesis. Members of the
ATP binding cassette (ABC) transporter superfamily have been demonstrated to be
able to recruit various alkaloids such as vinblastine, hyoscyamine, scopolamine or
berberine (Kolaczkowski et al. 1996; Sakai et al. 2002; Goossens et al. 2003; Shitan
et al. 2003). The possibility that alkaloid intermediates and/or enzymes flow
through the symplasm should also be considered, even though no experimental
proofs are available. Finally, the organisation of clusters of biosynthetic enzymes in
metabolic channels has also been purported in BIA synthesis (Samanani et al.
2006).
8.3.1 Crystallisation and Three-Dimensional Structure
of Alkaloid Biosynthetic Enzymes
The molecular architecture of some alkaloid biosynthetic enzymes has recently
been investigated, providing details on their catalytic mechanism and opening new
perspectives on the production of alkaloid derivatives with improved properties
using enzyme engineering. The TNA biosynthetic pathway provides the first two
examples of elucidation of three-dimensional architecture of plant alkaloid biosyn-
thetic enzymes by determination of the crystal structures of two tropinone reduc-
tases from D. stramonium (Nakajima et al. 1998). Similarly, elucidation of the
structure of two N-methyltransferases from the caffeine (PA) biosynthetic pathway
of Coffea canephora led to the identification of critical residues for substrate
selectivity and catalysis (McCarthy and McCarthy 2007). The MIA biosynthetic
pathway shows the broadest characterisation of alkaloid biosynthetic enzyme
structure. In R. serpentina, in addition to the preliminary X-ray analysis performed
on crystals of raucaffricine glucosidase and perakine reductase (Rosenthal et al.
2006; Ruppert et al. 2006), X-ray crystallography allowed the elucidation of STR
and SGD structures, as well as vinorine synthase located in the ajmaline branch of
MIA biosynthesis (Ma et al. 2005,2006; Barleben et al. 2007). STR represents a
novel six-bladed b-propeller fold protein that catalyses a Pictet-Spengler conden-
sation between secologanin and tryptamine in a highly substrate-specific manner.
The elucidation of the architecture of the substrate binding pockets and of catalytic
residues led to the development of structure-based engineering of STR aiming at a
redesign of the substrate binding pockets (Chen et al. 2006; Bernhardt et al. 2007;
Loris et al. 2007). Several STR variants generated by site-directed mutagenesis
displayed altered substrate specificity and accommodated analogs of both trypt-
amine and secologanin. The three-dimensional structure of SGD appears as a
typical (b/a)
8
barrel fold as encountered in the glucosidase family 1. Using
152 G. Guirimand et al.
site-directed mutagenesis and structural analysis, catalytic residues have been
identified that are involved in the deglucosylation of strictosidine and the confor-
mation of the catalytic pocket (Barleben et al. 2007). Furthermore, steady-state
kinetics of SGD with various strictosidine analogs enabled the identification of the
substrate preference of SGD at two positions of strictosidine, opening new oppor-
tunities for SGD redesign (Yerkes et al. 2008). Finally, crystal structure analysis of
vinorine synthase provides the first example of the three-dimensional organisation
of an enzyme of the BAHD superfamily (Ma et al. 2005). This is of particular
interest, since the vinorine synthase structure could be used for homology-based
modelling of other BAHD enzymes occurring in MIA and BIA biosynthetic path-
ways. Indeed, homology-based modelling has already been applied successfully to
the identification of catalytic residue of polyneuridine aldehyde esterase (PNAE),
an ab-hydrolase superfamily member from the ajmaline biosynthetic pathway of
R. serpentina, using the X-ray crystallographic structure of hydroxynitrile lyase,
another ab-hydrolase superfamily member (Mattern-Dogru et al. 2002). In Papaver
bracteatum, the structure of salutaridine reductase implicated in BIA biosynthesis
was also analysed by homology modelling using the X-ray structure of human
carbonyl reductase 1 (Geissler et al. 2007).
The elucidation of the structure of alkaloid biosynthetic enzymes using X-ray
crystallography and homology modelling, as well as the identification of key
residues for reaction catalysis and substrate specificity, open new enzyme redesign
perspectives for the production of large libraries of alkaloid analogs of potential
interest.
8.3.2 Transcription Factor Regulatory Networks of Alkaloid
Biosynthesis
Transcription factors (TF) are thought to play a key role in regulating fluxes through
alkaloid pathways by controlling the levels of pathway gene expression, transpor-
ters, and the differentiation of specialised cellular structures where the alkaloids are
synthesised and accumulate. TF usually exert their control by modulating the
expression level of multiple pathway genes. In several species, alkaloid accumula-
tion is preceded by the coordinated induction of several pathway genes, a result of
their regulation by specific TF. Methyl jasmonate (MeJA), one of the major internal
signal molecule in plant defence response, is known to induce a number of
secondary metabolisms, including nicotine (Shoji et al. 2008), MIA (van der Fits
and Memelink 2000) and BIA (Fa
¨rber et al. 2003). For instance, the majority of
MIA pathway genes tested are induced by MeJA in C. roseus cell cultures (van der
Fits and Memelink 2000; Oudin et al. 2007). Detailed analysis of the STR gene
promoter led to the identification of several TF involved in elicitor and jasmonate
responses. The jasmonate- and elicitor-responsive element (JERE) was found to
interact with two TF, octadecanoid-responsive Catharanthus AP2-domain proteins
8 Biosynthesis and Regulation of Alkaloids 153
(ORCAs; Menke et al. 1999; van der Fits and Memelink 2000). The expression of
the ORCA2 and ORCA3 genes themselves is rapidly induced by MeJA (Menke
et al. 1999; van der Fits and Memelink 2001). In nicotine biosynthesis, MeJA
response is controlled by homologs of the Arabidopsis MeJA-signal transduction
pathway (Chini et al. 2007; Thines et al. 2007; Katsir et al. 2008), namely the
NtCOI1 subunit of E3-ubiquitin ligase complex, the NtJAZ repressor family (Shoji
et al. 2008), as well as by the transcription factor NtORC1, an homolog of ORCA3,
and NtJAP1 (De Sutter et al. 2005). NtORC1 and NtJAP1 were shown to up-
regulate the promoter of the PMT gene encoding the first committed step in nicotine
biosynthesis (De Sutter et al. 2005). In BIA biosynthesis, a WRKY protein was
isolated recently as a transcriptional regulator of the berberine pathway in
C. japonica (Kato et al. 2007). Silencing of CjWRKY1 down-regulated the expres-
sion of several berberine pathway genes, whereas over-expression of CjWRKY1
up-regulated the berberine pathway but did not regulate primary metabolism genes,
including the tyrosine biosynthetic genes. Interestingly, over-expression of an
Arabidopsis WRKY TF was found to enhance BIA concentration up to 30-fold in
California poppy cell cultures (Apuya et al. 2008). This TF was isolated in a study
that involved screening of regulatory factors isolated from Arabidopsis to identify
those that modulate the expression of genes encoding enzymes involved in the
biosynthesis of BIA in P. somniferum and E. californica. In opium poppy, the over-
expression of selected regulatory factors was found to increase the concentrations
of codeinone reductase, 30-hydroxy-N-methylcoclaurine 40-O-methyltransferase
and norcoclaurine 6-O-methyltransferase transcripts 10- to 100-fold, and to
enhance BIA production up to tenfold (Apuya et al. 2008).
8.3.3 Metabolic Engineering of Alkaloid Biosynthesis
Progress in the knowledge of alkaloid pathways at the gene level allows several
attempts to improve alkaloid factories. Various objectives have been considered to
improve the economic value of medicinal plants or cell cultures, including yield
improvement of active or total alkaloids by enhancing flux in the pathway from
primary precursor, by overcoming the rate-limiting step, and by reducing flux to
competing pathways. Metabolic capacity to produce alkaloids has also been intro-
duced into new species by gene transfer or, in contrast, blocked to eliminate toxic or
undesired alkaloids.
Significant yield improvement has been achieved in opium poppy recently by
deregulating either an early or a late metabolic step. In an elite commercial line of
P. somniferum, over-expression of the CYP80B3 gene encoding the early-step
N-methylcoclaurine 30-hydroxylase resulted in an increase in morphine reaching
450% under greenhouse conditions (Frick et al. 2007). Over-expression of the
penultimate step, codeinone reductase, also resulted in an increase in total alkaloids
and morphine of up to 20% under field conditions (Larkin et al. 2007). In MIA
biosynthesis, attempts to deregulate the indole branch have been disappointing,
154 G. Guirimand et al.
probably owing to the dual origin of this class of alkaloids and the complexity of the
regulatory networks. Over-expression of tryptophan decarboxylase (TDC) in
C. roseus cell culture increased the concentration of only tryptamine (Whitmer
et al. 2002), while over-expression of both TDC and STR resulted in increased
tryptamine and MIA (Canel et al. 1998; Geerlings et al. 1999). Similarly, in
C. roseus hairy roots, inducible expression of both a feedback insensitive anthrani-
late synthase from Arabidopsis and TDC led to a marked increase in tryptophan and
tryptamine pools, but only modest change in alkaloid pools (Hong et al. 2006). The
terpenoid branch has been shown to be limiting in these transgenic lines by
precursor feeding experiments (Whitmer et al. 2002; Peebles et al. 2006), suggest-
ing that further improvement would need to target this part of the pathway.
Genetic manipulation of the tropane alkaloid pathway is a good example to
improve the metabolic profile of a medicinal plant. Compared to the more valuable
scopolamine, hyoscyamine usually accumulates to greater concentrations in vari-
ous solanaceous plants (e.g. Atropa,Datura,Hyoscyamus). Improved conversion of
the precursor hyoscyamine to scopolamine has been achieved by over-expression of
hyoscyamine 6-hydroxylase (Yun et al. 1992; Hashimoto et al. 1993). Recently,
biotransformation of hyoscyamine into scopolamine has been reported in transgenic
tobacco cell cultures by over-expressing the H6H gene from Hyoscyamus muticus
(Moyano et al. 2007). In some cases, incomplete knowledge of metabolic pathways
may be overcome by using a more conventional approach. For instance, chemical
mutagenesis of seeds has resulted in the isolation of a P. somniferum mutant, top1,
which accumulates thebaine, the starting material for several analgesic semi-
synthetic derivatives (Milgate et al. 2004).
Blocking an undesired production of alkaloid may be achieved by silencing an
essential enzymatic step early in the pathway. For instance, engineering of decaf-
feinated coffee plants is in progress in Coffea canophora, by silencing 7-methyl-
xanthine methyltransferase (Ashihara et al. 2006). Due to the high conservation of
the amino acid sequence, the two other N-methyltransferases were also down-
regulated. Conversely, new metabolic capacity for caffeine production was intro-
duced into tobacco by expressing the three N-methyltransferase genes. The
transgenic plants accumulated caffeine, which apparently deters insect pests and
enhances resistance to bacterial and viral infections (Uefuji et al. 2005; Kim and
Sano 2008). In tobacco, RNAi-induced suppression of nicotine demethylase activ-
ity was shown to reduce the levels of N0-nitrosonornicotine, a carcinogen in cured
tobacco leaves (Lewis et al. 2008).
8.4 Conclusions
Progress in metabolomic and transcriptomic analyses as well as system biology
approaches are likely to yield detailed analysis of bottlenecks in metabolic path-
ways and more rational strategies to improve alkaloid biosynthesis. The importance
of cellular and subcellular compartmentation, including metabolic channels, has
8 Biosynthesis and Regulation of Alkaloids 155
been so far overlooked in improving alkaloid yield. The targeting of enzymes to
proper compartments is likely to be an important factor. Engineering at the enzyme
structural level by site-directed mutagenesis will also likely be an important avenue
to create a new level of metabolic diversity by generating new semi-synthetic
alkaloid structures from synthetic precursors and modified enzymes with broad
substrate specificity.
References
Alcantara J, Bird DA, Franceschi VR, Facchini PJ (2005) Sanguinarine biosynthesis is associated
with the endoplasmic reticulum in cultured opium poppy cells after elicitor treatment. Plant
Physiol 138:173–183
Apuya NR, Park J-H, Zhang L, Ahyow M, Davidow P, Van Fleet J, Rarang JC, Hippley M,
Johnson TW, Yoo H-D, Trieu A, Krueger S, Wu C-y, Lu Y-p, Flavell RB, Bobzin SC (2008)
Enhancement of alkaloid production in opium and California poppy by transactivation using
heterologous regulatory factors. Plant Biotechnol J 6:160–175
Ashihara H, Zheng XQ, Katahira R, Morimoto M, Ogita S, Sano H (2006) Caffeine biosynthesis
and adenine metabolism in transgenic Coffea canephora plants with reduced expression of
N-methyltransferase genes. Phytochemistry 67:882–886
Ashihara H, Sano H, Crozier A (2008) Caffeine and related purine alkaloids: biosynthesis,
catabolism, function and genetic engineering. Phytochemistry 69:841–856
Barleben L, Panjikar S, Ruppert M, Koepke J, Sto
¨ckigt J (2007) Molecular architecture of
strictosidine glucosidase: the gateway to the biosynthesis of the monoterpenoid indole alkaloid
family. Plant Cell 19:2886–2897
Bernhardt P, McCoy E, O’Connor SE (2007) Rapid identification of enzyme variants for reengi-
neered alkaloid biosynthesis in periwinkle. Chem Biol 14:888–897
Bird DA, Facchini PJ (2001) Berberine bridge enzyme, a key branch-point enzyme in benzyliso-
quinoline alkaloid biosynthesis, contains a vacuolar sorting determinant. Planta 213:888–897
Bird DA, Franceschi VR, Facchini P (2003) A tale of three cell types: alkaloid biosynthesis is
localized to sieve elements in opium poppy. Plant Cell 15:2626–2635
Bock A, Wanner G, Zenk MH (2002) Immunocytological localization of two enzymes involved in
berberine biosynthesis. Planta 216:57–63
Burlat V, Oudin A, Courtois M, Rideau M, St-Pierre B (2004) Co-expression of three MEP
pathway genes and geraniol 10-hydroxylase in internal phloem parenchyma of Catharanthus
roseus implicates multicellular translocation of intermediates during the biosynthesis of
monoterpene indole alkaloids and isoprenoid-derived primary metabolites. Plant J 38:131–141
Canel C, Lopes-Cardoso I, Whitmer S, van der Fits L, Pasquali G, van der Heijden R, Hoge JHC,
Verpoorte R (1998) Effects of over-expression of strictosidine synthase and tryptophan
decarboxylase on alkaloid production by cell cultures of Catharanthus roseus. Planta
205:414–419
Carlson JE, Leebens-Mack JH, Wall PK, Zahn LM, Mueller LA, Landherr LL, Hu Y, Ilut DC,
Arrington JM, Choirean S, Becker A, Field D, Tanksley SD, Ma H, dePamphilis CW (2006)
EST database for early flower development in California poppy (Eschscholzia californica
Cham., Papaveraceae) tags over 6,000 genes from a basal eudicot. Plant Mol Biol 62:351–369
Chen S, Galan MC, Coltharp C, O’Connor SE (2006) Redesign of a central enzyme in alkaloid
biosynthesis. Chem Biol 13:1137–1141
Chini A, Fonseca S, Ferna
´ndez G, Adie B, Chico JM, Lorenzo O, Garcı
´a-Casado G, Lo
´pez-
Vidriero I, Lozano FM, Ponce MR, Micol JL, Solano R (2007) The JAZ family of repressors is
the missing link in jasmonate signalling. Nature 448:666–671
156 G. Guirimand et al.
Costa MM, Hilliou F, Duarte P, Pereira LG, Almeida I, Leech M, Memelink J, Barcelo
´AR,
Sottomayor M (2008) Molecular cloning and characterization of a vacuolar class III peroxidase
involved in the metabolism of anticancer alkaloids in Catharanthus roseus. Plant Physiol
146:403–417
De Luca V, St-Pierre B (2000) The cell and developmental biology of alkaloid biosynthesis.
Trends Plant Sci 5:168–173
De Sutter V, Vanderhaeghen R, Tilleman S, Lammertyn F, Vanhoutte I, Karimi M, Inze
´D,
Goossens A, Hilson P (2005) Exploration of jasmonate signalling via automated and standar-
dized transient expression assays in tobacco cells. Plant J 44:1065–1076
Dra
¨ger B (2002) Analysis of tropane and related alkaloids. J Chromatogr A 978:1–35
Dra
¨ger B (2006) Tropinone reductases, enzymes at the branch point of tropane alkaloid metabo-
lism. Phytochemistry 67:327–337
Facchini PJ, St-Pierre B (2005) Synthesis and trafficking of alkaloid biosynthetic enzymes. Curr
Opin Plant Biol 8:657–666
Fa
¨rber K, Schumann B, Miersch O, Roos W (2003) Selective desensitization of jasmonate and pH-
dependent signaling in the induction of benzophenanthridine biosynthesis in cells of
Eschscholzia californica. Phytochemistry 62:491–500
Frick S, Kramell R, Kutchan TM (2007) Metabolic engineering with a morphine biosynthetic P450
in opium poppy surpasses breeding. Metab Eng 9:169–176
Geerlings A, Hallard D, Caballero AM, Cardoso IL, van der Heijden R, Verpoorte R (1999)
Alkaloid production by a Cinchona officinalis “Ledgeriana” hairy root culture containing
constitutive expression constructs of tryptophan decarboxylase and strictosidine synthase
cDNAs from Catharanthus roseus. Plant Cell Rep 19:191–196
Geissler R, Brandt W, Ziegler J (2007) Molecular modeling and site-directed mutagenesis reveal
the benzylisoquinoline binding site of the short-chain dehydrogenase/reductase salutaridine
reductase. Plant Physiol 143:1493–1503
Goossens A, Ha
¨kkinen ST, Laakso I, Oksman-Caldentey KM, Inze
´D (2003) Secretion of
secondary metabolites by ATP-binding cassette transporters in plant cell suspension cultures.
Plant Physiol 131:1161–1164
Guirimand G, Burlat V, Oudin A, Lanoue A, St-Pierre B, Courdavault V (2009) Optimization of
the transient transformation of Catharanthus roseus cells by particle bombardment and its
application to the subcellular localization of hydroxymethylbutenyl 4-diphosphate synthase
and geraniol 10-hydroxylase. Plant Cell Rep 28:1215–1234
Hagel JM, Facchini PJ (2008) Plant metabolomics: analytical platforms and integration with
functional genomics. Phytochem Rev 7:479–497
Hagel JM, Weljie AM, Vogel HJ, Facchini PJ (2008) Quantitative 1H nuclear magnetic resonance
metabolite profiling as a functional genomics platform to investigate alkaloid biosynthesis in
opium poppy. Plant Physiol 147:1805–1821
Hashimoto T, Hayashi A, Amano Y, Kohno J, Iwanari H, Usuda S, Yamada Y (1991) Hyoscya-
mine 6b-hydroxylase, an enzyme involved in tropane alkaloid biosynthesis, is localized at the
pericycle of the root. J Biol Chem 266:4648–4653
Hashimoto T, Yun DJ, Yamada Y (1993) Production of tropane alkaloids in genetically engineered
root cultures. Phytochemistry 32:713–718
Hong SB, Peebles CAM, Shanks JV, San KY, Gibson SI (2006) Expression of the Arabidopsis
feedback-insensitive anthranilate synthase holoenzyme and tryptophan decarboxylase genes in
Catharanthus roseus hairy roots. J Biotechnol 122:28–38
Irmler S, Schro
¨der G, St-Pierre B, Crouch NP, Hotze M, Schmidt J, Strack D, Matern U, Schro
¨der J
(2000) Indole alkaloid biosynthesis in Catharanthus roseus: new enzyme activities and
identification of cytochrome P450 CYP72A1 as secologanin synthase. Plant J 24:797–804
Kato N, Dubouzet E, Kokabu Y, Yoshida S, Taniguchi Y, Dubouzet JG, Yazaki K, Sato F (2007)
Identification of a WRKY protein as a transcriptional regulator of benzylisoquinoline alkaloid
biosynthesis in Coptis japonica. Plant Cell Physiol 48:8–18
8 Biosynthesis and Regulation of Alkaloids 157
Katsir L, Schilmiller AL, Staswick PE, He SY, Howe GA (2008) COI1 is a critical component of a
receptor for jasmonate and the bacterial virulence factor coronatine. Proc Natl Acad Sci USA
105:7100–7105
Kim Y-S, Sano H (2008) Pathogen resistance in transgenic tobacco plants producing caffeine.
Phytochemistry 69:882–888
Kolaczkowski M, van der Rest M, Cybularz-Kolaczkowska A, Soumillion J-P, Konings WN,
Goffeau A (1996) Anticancer drugs, ionophoric peptides, and steroids as substrates of the yeast
multidrug transporter Pdr5p. J Biol Chem 271:31543–31548
Kutchan TM (2005) A role for intra- and intercellular translocation in natural product biosynthesis.
Curr Opin Plant Biol 8:292–300
Larkin PJ, Miller JAC, Allen RS, Chitty JA, Gerlach WL, Frick S, Kutchan TM, Fist AJ (2007)
Increasing morphinan alkaloid production by over-expressing codeinone reductase in trans-
genic Papaver somniferum. Plant Biotechnol J 5:26–37
Levac D, Murata J, Kim WS, De Luca V (2008) Application of carborundum abrasion for
investigating the leaf epidermis: molecular cloning of Catharanthus roseus 16-hydroxytaber-
sonine-16-O-methyltransferase. Plant J 53:225–236
Lewis RS, Jack AM, Morris JW, Robert VJM, Gavilano LB, Siminszky B, Bush LP, Hayes AJ,
Dewey RE (2008) RNA interference (RNAi)-induced suppression of nicotine demethylase
activity reduces levels of a key carcinogen in cured tobacco leaves. Plant Biotechnol J
6:346–354
Li R, Reed DW, Liu E, Nowak J, Pelcher LE, Page JE, Covello PS (2006) Functional genomic
analysis of alkaloid biosynthesis in Hyoscyamus niger reveals a cytochrome P450 involved in
littorine rearrangement. Chem Biol 13:513–520
Lorence A, Nessler CL (2004) Camptothecin, over four decades of surprising findings. Phyto-
chemistry 65:2735–2749
Loris EA, Panjikar S, Ruppert M, Barleben L, Unger M, Schu
¨bel H, Sto
¨ckigt J (2007) Structure-
based engineering of strictosidine synthase: auxiliary for alkaloid libraries. Chem Biol
14:979–985
Ma X, Koepke J, Panjikar S, Fritzsch G, Sto
¨ckigt J (2005) Crystal structure of vinorine synthase,
the first representative of the BAHD superfamily. J Biol Chem 280:13576–13583
Ma X, Panjikar S, Koepke J, Loris E, Sto
¨ckigt J (2006) The structure of Rauvolfia serpentina
strictosidine synthase is a novel six-bladed beta-propeller fold in plant proteins. Plant Cell
18:907–920
Mahroug S, Burlat V, St-Pierre B (2007) Cellular and sub-cellular organization of the mono-
terpenoid indole alkaloid pathway in Catharanthus roseus. Phytochem Rev 6:363–381
Mattern-Dogru E, Ma X, Hartmann J, Decker H, Sto
¨ckigt J (2002) Potential active-site residues in
polyneuridine aldehyde esterase, a central enzyme of indole alkaloid biosynthesis, by model-
ling and site-directed mutagenesis. Eur J Biochem 269:2889–2896
McCarthy AA, McCarthy JG (2007) The structure of two N-methyltransferases from the caffeine
biosynthetic pathway. Plant Physiol 144:879–889
McKnight TD, Bergey DR, Burnett RJ, Nessler CL (1991) Expression of enzymatically active and
correctly targeted strictosidine synthase in transgenic tobacco plants. Planta 185:148–152
Menke FLH, Champion A, Kijne JW, Memelink J (1999) A novel jasmonate- and elicitor-
responsive element in the periwinkle secondary metabolite biosynthetic gene Str interacts
with a jasmonate- and elicitor-inducible AP2-domain transcription factor, ORCA2. EMBO J
18:4455–4463
Milgate AG, Pogson BJ, Wilson IW, Kutchan TM, Zenk MH, Gerlach WL, Fist AJ, Larkin PJ
(2004) Morphine-pathway block in top1 poppies. Nature 431:413–414
Morishige T, Dubouzet E, Choi KB, Yazaki K, Sato F (2002) Molecular cloning of columbamine
O-methyltransferase from cultured Coptis japonica cells. Eur J Biochem 269:5659–5667
Moyano E, Palazo
´n J, Bonfill M, Osuna L, Cusido
´RM, Oksman-Caldentey K-M, Pin
˜ol MT (2007)
Biotransformation of hyoscyamine into scopolamine in transgenic tobacco cell cultures. J Plant
Physiol 164:521–524
158 G. Guirimand et al.
Murata J, De Luca V (2005) Localization of tabersonine 16-hydroxylase and 16-OH tabersonine-
16-O-methyltransferase to leaf epidermal cells defines them as a major site of precursor
biosynthesis in the vindoline pathway in Catharanthus roseus. Plant J 44:581–594
Murata J, Bienzle D, Brandle JE, Sensen CW, De Luca V (2006) Expressed sequence tags from
Madagascar periwinkle (Catharanthus roseus). FEBS Lett 580:4501–4507
Murata J, Roepke J, Gordon H, De Luca V (2008) The leaf epidermome of Catharanthus roseus
reveals its biochemical specialization. Plant Cell 20:524–542
Nakajima K, Hashimoto T (1999) Two tropinone reductases, that catalyze opposite stereospecific
reductions in tropane alkaloid biosynthesis, are localized in plant root with different cell-
specific patterns. Plant Cell Physiol 40:1099–1107
Nakajima K, Yamashita A, Akama H, Nakatsu T, Kato H, Hashimoto T, Oda J, Yamada Y (1998)
Crystal structures of two tropinone reductases: different reaction stereospecificities in the same
protein fold. Proc Natl Acad Sci USA 95:4876–4881
Oksman-Caldentey KM (2007) Tropane and nicotine alkaloid biosynthesis-novel approaches
towards biotechnological production of plant-derived pharmaceuticals. Curr Pharm Biotechnol
8:203–210
Oudin A, Mahroug S, Courdavault V, Hervouet N, Zelwer C, Rodrı
´guez-Concepcio
´n M, St-Pierre
B, Burlat V (2007) Spatial distribution and hormonal regulation of gene products from methyl
erythritol phosphate and monoterpene-secoiridoid pathways in Catharanthus roseus. Plant Mol
Biol 65:13–30
Peebles CAM, Hong SB, Gibson SI, Shanks IV, San KY (2006) Effects of terpenoid precursor
feeding on Catharanthus roseus hairy roots over-expressing the alpha or the alpha and beta
subunits of anthranilate synthase. Biotechnol Bioeng 93:534–540
Rischer H, Oresic M, Seppa
¨nen-Laakso T, Katajamaa M, Lammertyn F, Ardiles-Diaz W, Van
Montagu MC, Inze
´D, Oksman-Caldentey KM, Goossens A (2006) Gene-to-metabolite net-
works for terpenoid indole alkaloid biosynthesis in Catharanthus roseus cells. Proc Natl Acad
Sci USA 103:5614–5619
Rosenthal C, Mueller U, Panjikar S, Sun L, Ruppert M, Zhao Y, Sto
¨ckigt J (2006) Expression,
purification, crystallization and preliminary X-ray analysis of perakine reductase, a new
member of the aldo-keto reductase enzyme superfamily from higher plants. Acta Crystallogr
Sect F Struct Biol Cryst Commun 62:1286–1289
Ruppert M, Panjikar S, Barleben L, Sto
¨ckigt J (2006) Heterologous expression, purification,
crystallization and preliminary X-ray analysis of raucaffricine glucosidase, a plant enzyme
specifically involved in Rauvolfia alkaloid biosynthesis. Acta Crystallogr Sect F Struct Biol
Cryst Commun 62:257–260
Sakai K, Shitan N, Sato F, Ueda K, Yazaki K (2002) Characterization of berberine transport into
Coptis japonica cells and the involvement of ABC protein. J Exp Bot 53:1879–1886
Samanani N, Park SU, Facchini PJ (2005) Cell type-specific localization of transcripts encoding
nine consecutive enzymes involved in protoberberine alkaloid biosynthesis. Plant Cell
17:915–926
Samanani N, Alcantara J, Bourgault R, Zulak KG, Facchini PJ (2006) The role of phloem sieve
elements and laticifers in the biosynthesis and accumulation of alkaloids in opium poppy. Plant
J 47:547–563
Schmidt J, Boettcher C, Kuhnt C, Kutchan TM, Zenk MH (2007) Poppy alkaloid profiling by
electrospray tandem mass spectrometry and electrospray FT-ICR mass spectrometry after
[ring-
13
C
6
]-tyramine feeding. Phytochemistry 68:189–202
Shitan N, Bazin I, Dan K, Obata K, Kigawa K, Ueda K, Sato F, Forestier C, Yazaki K (2003)
Involvement of CjMDR1, a plant multidrug-resistance-type ATP-binding cassette protein, in
alkaloid transport in Coptis japonica. Proc Natl Acad Sci USA 100:751–756
Shoji T, Ogawa T, Hashimoto T (2008) Jasmonate-induced nicotine formation in tobacco is
mediated by tobacco COI1 and JAZ genes. Plant Cell Physiol 49:1003–1012
8 Biosynthesis and Regulation of Alkaloids 159
Shukla AK, Shasany AK, Gupta MM, Khanuja SP (2006) Transcriptome analysis in Catharanthus
roseus leaves and roots for comparative terpenoid indole alkaloid profiles. J Exp Bot
57:3921–3932
Sto
¨ckigt J, Panjikar S (2007) Structural biology in plant natural product biosynthesis-architecture
of enzymes from monoterpenoid indole and tropane alkaloid biosynthesis. Nat Prod Rep
24:1382–1400
St-Pierre B, Vazquez-Flota FA, De Luca V (1999) Multicellular compartmentation of Cathar-
anthus roseus alkaloid biosynthesis predicts intercellular translocation of a pathway interme-
diate. Plant Cell 11:887–900
Suzuki K, Yamada Y, Hashimoto T (1999a) Expression of Atropa belladonna putrescine
N-methyltransferase gene in root pericycle. Plant Cell Physiol 40:289–297
Suzuki K, Yun DJ, Chen XY, Yamada Y, Hashimoto T (1999b) An Atropa belladonna hyoscya-
mine 6b-hydroxylase gene is differentially expressed in the root pericycle and anthers. Plant
Mol Biol 40:141–152
Thines B, Katsir L, Melotto M, Niu Y, Mandaokar A, Liu G, Nomura K, He SY, Howe GA,
Browse J (2007) JAZ repressor proteins are targets of the SCF
COI1
complex during jasmonate
signalling. Nature 448:661–665
Uefuji H, Tatsumi Y, Morimoto M, Kaothien-Nakayama P, Ogita S, Sano H (2005) Caffeine
production in tobacco plants by simultaneous expression of three coffee N-methyltransferases
and its potential as a pest repellent. Plant Mol Biol 59:221–227
van der Fits L, Memelink J (2000) ORCA3, a jasmonate-responsive transcriptional regulator of
plant primary and secondary metabolism. Science 289:295–297
van der Fits L, Memelink J (2001) The jasmonate-inducible AP2/ERF-domain transcription factor
ORCA3 activates gene expression via interaction with a jasmonate-responsive promoter
element. Plant J 25:43–53
van der Heijden R, Jacobs DI, Snoeijer W, Hallard D, Verpoorte R (2004) The Catharanthus
alkaloids: pharmacognosy and biotechnology. Curr Med Chem 11:607–628
Weid M, Ziegler J, Kutchan TM (2004) The roles of latex and the vascular bundle in mor-
phine biosynthesis in the opium poppy, Papaver somniferum. Proc Natl Acad Sci USA 101:
13957–13962
Whitmer S, van der Heijden R, Verpoorte R (2002) Effect of precursor feeding on alkaloid
accumulation by a tryptophan decarboxylase over-expressing transgenic cell line T22 of
Catharanthus roseus. J Biotechnol 96:193–203
Yamazaki Y, Urano A, Sudo H, Kitajima M, Takayama H, Yamazaki M, Aimi N, Saito K (2003)
Metabolite profiling of alkaloids and strictosidine synthase activity in camptothecin producing
plants. Phytochemistry 62:461–470
Yerkes N, Wu JX, McCoy E, Galan MC, Chen S, O’Connor SE (2008) Substrate specificity and
diastereoselectivity of strictosidine glucosidase, a key enzyme in monoterpene indole alkaloid
biosynthesis. Bioorg Med Chem Lett 18:3095–3098
Yun D-J, Hashimoto T, Yamada Y (1992) Metabolic engineering of medicinal plants: transgenic
Atropa belladonna with an improved alkaloid composition. Proc Natl Acad Sci USA 89:
11799–11803
Ziegler J, Facchini PJ (2008) Alkaloid biosynthesis: metabolism and trafficking. Annu Rev Plant
Biol 59:735–769
Ziegler J, Voigtla
¨nder S, Schmidt J, Kramell R, Miersch O, Ammer C, Gesell A, Kutchan TM
(2006) Comparative transcript and alkaloid profiling in Papaver species identifies a short chain
dehydrogenase/reductase involved in morphine biosynthesis. Plant J 48:177–192
Zulak KG, Cornish A, Daskalchuk TE, Deyholos MK, Goodenowe DB, Gordon PM, Klassen D,
Pelcher LE, Sensen CW, Facchini PJ (2007) Gene transcript and metabolite profiling of
elicitor-induced opium poppy cell cultures reveals the coordinate regulation of primary and
secondary metabolism. Planta 225:1085–1106
160 G. Guirimand et al.
... 53,54 In short-term conversion attempts, the STRGFP signal occurs in the vacuole. 55 STR is designated as a vacuole through the ER to the Golgi to the vacuole route. SGD is a dichotomous NLS centered on the nucleus and tends to multimerize in this compartment of the cell. ...
... 56 In transient transformation experiments, STR GFP signals are generated in vacuoles. 55 The STR is directed to the vacuole through the ER to the Golgi to the vacuole pathway. SGD is a dichotomous NLS that targets the nucleus and tends to multimerize in this cell compartment. ...
Article
Full-text available
The medicinal plant C. roseus synthesizes biologically active alkaloids via the terpenoid indole alkaloid (TIAs) biosynthetic pathway. Most of these alkaloids have high therapeutic value, such as vinblastine and vincristine. Plant signaling components, plant hormones, precursors, growth hormones, prenylated proteins, and transcriptomic factors regulate the complex networks of TIA biosynthesis. For many years, researchers have been evaluating the scientific value of the TIA biosynthetic pathway and its potential in commercial applications for market opportunities. Metabolic engineering has revealed the major blocks in metabolic pathways regulated at the molecular level, unknown structures, metabolites, genes, enzyme expression, and regulatory genes. Conceptually, this information is necessary to create transgenic plants and microorganisms for the commercial production of high-value dimer alkaloids, such as vinca alkaloids, vinblastine, and vincristine In this review, we present current knowledge of the regulatory mechanisms of these components in the C. roseus TIA pathway, from genes to metabolites. Keywords: Metabolites,Vinblastine and Vincristine. Plant signalling components, Terpenoid indole alkaloids,transcriptom factors.
... More than 200 alkaloids can be classified as tropane alkaloids, and these are mainly distributed in the angiosperms of different families, such as Proteaceae, Solanaceae, Erythroxylaceae, Convolvulaceae, Brassicaceae, and Euphorbiaceae (Jirschitzka et al. 2012). The most studied are found in the Solanaceae family, mostly in the genera Datura (hyoscyamine and scopolamine), Hyoscyamus (scopolamine), and Atropa (atropine alkaloid) (Ajungla et al. 2009;Wink 2010;Guirimand et al. 2010). Another well-known tropane alkaloid is the cocaine, obtained from Erythroxylum coca (Erythroxylaceae). ...
... There are approximately 2000 identified compounds known as indole alkaloids distributed in different families such as Apocynaceae, Loganiaceae, Rubiaceae, and Nyssaceae, as well as plant species such as Catharanthus roseus (vinblastine and vincristine alkaloids), Rauvolfia serpentina (ajmalicine), Camptotheca acuminata (camptothecin), Passiflora incarnata (harman, harmol, harmine, harmaline), Mitragyna speciosa (mitragynine), Rauwolfia serpentina (reserpine, serpentinine, ajmaline, corynanthine, and yohimbine (Guirimand et al. 2010;Sagi et al. 2016;Hamid et al. 2017). ...
Chapter
Full-text available
Alkaloids are nitrogen-containing natural products found in bacteria, fungi, animals, and plants with complex and diverse structures. The widespread distribution of alkaloids along with their wide array of structures makes their classification often difficult. However, for their study, alkaloids can be classified depending on their chemical structure, biochemical origin, and/or natural origin. Alkaloids can be derived from several biosynthetic pathways, such as the shikimate pathway; the ornithine, lysine, and nicotinic acid pathway; the histidine and purine pathway; and the terpenoid and polyketide pathway. Traditionally, plant alkaloids have played a pivotal role in folk medicines since ancient times as purgatives, antitussives, sedatives, and treatments for a wide variety of ailments. Currently, several alkaloids have served as models for modern drugs, and there are several alkaloids used in pharmacology, such as codeine, brucine, morphine, ephedrine, and quinine. Herein, this work is a comprehensive revision from the Web of Knowledge and Scopus databases on the recent information (2010–2019) regarding plant-derived alkaloids, their structural classification and bioactive properties.
... Their high biological activities also confer to MIAs a prominent place in human pharmacopoeia with premium anticancer agents vinblastine and vincristine from the Madagascar periwinkle (Catharanthus roseus), the antimalarial quinine produced in the cinchona bark (Cinchona officinalis), the antiarrhythmic ajmaline synthesized by the devil pepper (Rauwolfia serpentina) or the nootropic vincamine accumulated in the dwarf periwinkle (Vinca minor) (O'Connor and Maresh 2006). Over the last decades, numerous studies established that MIAs are produced in planta thanks to complex routes encompassing up to 30-40 enzymatic steps in the aforementioned plants (Guirimand et al. 2010b, Courdavault et al. 2014, Caputi et al. 2018. These pathways involve enzymes from several protein families including P450, glucosyltransferases, βglucosidases, dioxygenases, BAHD acyltransferases, decarboxylases, short-and medium-chain dehydrogenases/reductases, 'Pictet-Spenglerases' or O-methyltransferases and N-methyltransferases (NMTs) as non-exhaustive examples (Dugé de Bernonville et al. 2015a). ...
Article
Many plant species from the Apocynaceae, Loganiaceae and Rubiaceae families evolved a specialized metabolism leading to the synthesis of a broad palette of monoterpene indole alkaloids (MIAs). These compounds are believed to constitute a cornerstone of the plant chemical arsenal but above all several MIAs display pharmacological properties that have been exploited for decades by humans to treat various diseases. It is established that MIAs are produced in planta due to complex biosynthetic pathways engaging a multitude of specialized enzymes but also a complex tissue and subcellular organization. In this context, N-methyltransferases (NMTs) represent an important family of enzymes indispensable for MIA biosynthesis but their characterization has always remained challenging. In particular, little is known about the subcellular localization of NMTs in MIA-producing plants. Here, we performed an extensive analysis on the subcellular localization of NMTs from four distinct medicinal plants but also experimentally validated that two putative NMTs from Catharanthus roseus exhibit N-methyltransferase activity. Apart from providing unprecedented data regarding the targeting of these enzymes in planta, our results point out an additional layer of complexity to the subcellular organization of the MIA biosynthetic pathway by introducing tonoplast and peroxisome as new actors of the final steps of MIA biosynthesis.
... Plants produce an amazing diversity of low molecular weight natural products that exhibit wide range of structural diversity and pharmacological activities [1,2]. The chemical synthesis of these bioactive compounds is tedious due to their complexity frameworks. ...
Article
Flavonoids are naturally occurring compounds widely distributed in plants, which have hypoglycemic potential. Amylase inhibitors belong to flavonoids that block amylases which are involved in hydrolyzing the carbohydrates to yield low molecular weight dextrin and sugars. Inhibition of α-amylase along with α-glucosidase can significantly reduce the post-prandial increase of blood glucose and can be an important strategy in the control of blood glucose level in the type-2 diabetic patients. In this background, 21 samples including leaf, stem and bark of 10 plant species collected from different geographical regions of Jammu and Kashmir, India were evaluated for presence of α-amylase inhibitors. To quantify inhibition rates, acarbose was used as control. Methanolic extracts of Cedrus deodara leaves showed 97.31% amylase inhibition followed by Hippophae rhamnoides (75.43%) and Platanus orientalis (25.71%) as compared to control. Further, effect of pH and thermal stability of amylase inhibitors was analyzed. However, at pH 7 and 9, the amylase inhibition activity remains unaltered. Furthermore, activity of inhibitors was evaluated at different temperatures. Heat treated extracts of Cedrus deodara and Hippophae rhamnoides showed thermal stability of their activity. In the present study, rich sources of amylase inhibitors were identified. Characterization of key active metabolites from these species would be useful in management of postprandial hyperglycemia. Inhibition of α-amylase activity is only one possibility and cost-effective way to lower postprandial blood glucose levels.
... TDC activity was maximum in the epidermal cells,while DAT activity was detected only in the whole leaf whereas NMT activity was found in the whole leaf extract. The subcellular localization of pathway enzymes involved with vindoline synthesis and their dimerization reactions with catharanthine to form vinblastine and vincristine are fairly well understood now van der Fits et al. 2001;Murata et al. 2006Murata et al. , 2008O'Connor and Maresh 2006;Rischer et al. 2006;Loyola-Vargas et al. 2007;Stockigt and Panjikar 2007;Roepke et al. 2010;Shukla et al. 2010;Guirimand et al. 2010;Qu et al. 2015). The readers are advised to refer to the elegant review published by Pan et al. (2015) to have a detailed comprehensive idea about the land mark studies carried out in the field of TIAs (MIAs) pathway elucidation and regulation in C. roseus. ...
... Furthermore, vacuoles can be part of complex systems of spatial segregation between phytoanticipins and their activating enzymes such as in the well-known mustard oil bomb system relying on the vacuolar sequestration of myrosinases [19]. Interestingly, this type of organization has also been evolved for other plant specialized metabolites and notably monoterpene indole alkaloids (MIAs) that are synthetized in the Apocynaceae family whose Catharanthus roseus (Madagascar periwinkle) constitutes one of the prominent members [20,21]. C. roseus synthesizes more than 130 distinct MIAs that are a part of the chemical arsenal deployed by this plant to face pathogens and pest attacks [22,23]. ...
Article
Full-text available
Meta-analysis is a powerful method used to discern reliable genomic regions associated with quantitative traits. The objective of this study was to use a meta-QTL (MQTL) analysis method to identify the most reliable and accurate QTL controlling morphological traits and terpenoid indole alkaloids (TIAs); serpentine (S), vinblastine (VB), ajmalicine (A) and vincristine (VC) in Catharanthus roseus. A quantitative trait loci (QTL) database containing 40 QTLs retrieved from two individual mapping studies associated with 13 phytochemical and morphological traits. The results of projection of 40 putative QTLs on to the consensus genetic map showed that 18 MQTLs had a 25.88% reduction in confidence interval (CI) on C. roseus chromosomes. Co-localisation results suggested that leaves TAL and serpentine in roots (SR) with 50% and 60% frequencies were the most common QTLs associated with vindoline (V) and catharanthine (C). The MQTL.LG1.5, MQTL.LG2.2, MQTL.LG1.1, MQTL.LG1.2, MQTL.LG1.7 and MQTL.LG3.3 were the most important MQTLs and the most reliable chromosome regions for refining genetic control of the tested traits in C. roseus. The consensus map provided information for the identification of the QTL-dense genomic regions that will assist in improving the efficiency of genomic and marker-assisted selection (MAS) in pharmacological researches.
Article
Discovery of new natural products, especially those with high biological activities and application values, is of great research significance. However, conventional methods based on the cultivation of microbial mono-cultures can hardly satisfy the increasing need of novel natural product generation. Recently, the development of co-cultures composed of different species has emerged as an effective approach for mining novel natural products. Inspired by microbial communities in nature, these co-culture systems create favorable environmental conditions to promote interactions between co-culture members for activating the natural product biosynthesis that is hard to induce otherwise. A large variety of novel natural products have been identified using this robust approach. This review summarizes the recent achievements of using cross-species co-cultures for natural products discovery and discusses the existing challenges and future directions.
Article
Full-text available
The Madagascar periwinkle (Catharanthus roseus) synthesizes the highly valuable monoterpene indole alkaloids (MIAs) through a long metabolic route initiated by the 2C-methyl-D-erythritol 4-phosphate (MEP) pathway. In leaves, a complex compartmentation of the MIA biosynthetic pathway occurs at both the cellular and subcellular levels, notably for some gene products of the MEP pathway. To get a complete overview of the pathway organization, we cloned four genes encoding missing enzymes involved in the MEP pathway before conducting a systematic analysis of transcript distribution and protein subcellular localization. RNA in situ hybridization revealed that all MEP pathway genes were coordinately and mainly expressed in internal phloem-associated parenchyma of young leaves, reinforcing the role of this tissue in MIA biosynthesis. At the subcellular level, transient cell transformation and expression of fluorescent protein fusions showed that all MEP pathway enzymes were targeted to plastids. Surprisingly, two isoforms of 1-deoxy-D-xylulose 5-phosphate synthase and 1-deoxy-D-xylulose 5-phosphate reductoisomerase initially exhibited an artifactual aggregated pattern of localization due to high protein accumulation. Immunogold combined with transmission electron microscopy, transient transformations performed with a low amount of transforming DNA and fusion/deletion experiments established that both enzymes were rather diffuse in stroma and stromules of plastids as also observed for the last six enzymes of the pathway. Taken together, these results provide new insights into a potential role of stromules in enhancing MIA precursor exchange with other cell compartments to favor metabolic fluxes towards the MIA biosynthesis.
Chapter
Agricultural ethanol has not been an optional fuel for petrol due to a big difference in prices. Nowadays purposefulness of developing alternative fuel production processes, including bioethanol production, is supported by both economic and environmental concerns. The role of secondary metabolites in development of plant growth and architect for enhancing the biofuel production. Besides now as days the main focus are on gene and gene products as they construct the main apparatus for the studying the plant genome. The sequencing technology and reduced cost of sequencing are accelerating the ability to discover new natural product pathways by identifying the underlying genes. In this chapter various plant products and substances are documented right from the classical biofules to gene products vis a vis gene clusters.
Article
Full-text available
The Madagascar periwinkle (Catharanthus roseus) monoterpene indole alkaloids (MIAs), either as (hetero)dimers (vinblastine, vincristine) or as monomers (ajmalicine) are valuable pharmaceuticals with antitumoral and hypotensive properties. The biosynthetic pathway of these secondary metabolites involves several subpathways (so-called indole, methylerythritol phosphate, secoiridoid-monoterpenoids, and the MIA pathway itself). So far, 16 cDNA sequences with direct involvement in the pathway have been obtained and some additional enzymatic activities have been characterised. Metabolomic, transcriptomic and proteomic studies have revealed that this pathway presents a high degree of spatial organisation at the organ, cellular and sub-cellular levels. This paper reviews the evidence on the spatial organisation of the MIA pathway at various levels of resolution in terms of the following topics. (1) The organ-specific distribution of MIAs, related genes, related enzymes and related enzymatic activities, mainly focussing on repartition in roots versus aerial organs. (2) The cell-specific occurrence of MIA-related gene expression identified by in situ hybridisation and corresponding protein accumulation identified by immunolocalisation, with particular emphasis on the well-characterised multi-cellular compartmentation in aerial organs. In situ hybridisation revealed that the nine transcripts studied could be classified into three populations according to their position in the pathway (early-, intermediate- and late-step genes) and to their cell-specific expression patterns (internal phloem parenchyma, epidermis, laticifers/idioblasts). (3) The still poorly characterised sub-cellular localisation of MIA-related enzymes. Finally, these results are compared with the situation in other families producing other types of alkaloids, and future prospects are discussed.
Article
Full-text available
 Cinchona officinalis 'Ledgeriana', former called Cinchona ledgeriana, hairy roots were initiated containing constitutive-expression constructs of cDNAs encoding the enzymes tryptophan decarboxylase (TDC) and strictosidine synthase (STR) from Catharanthus roseus, two key enzymes in terpenoid indole and quinoline alkaloid biosynthesis. The successful integration of these genes and the reporter gene gus-int was demonstrated using Southern blotting and the polymerase chain reaction. The products of TDC and STR, tryptamine and strictosidine, were found in high amounts, 1200 and 1950 μg g–1 dry weight, respectively. Quinine and quinidine levels were found to rise up to 500 and 1000 μg g–1 dry weight, respectively. The results show that genetic engineering with multiple genes is well possible in hairy roots of C. officinalis. However, 1 year after analyzing the hairy roots for the first time, they had completely lost their capacity to accumulate alkaloids.
Article
Cultured Coptis japonica cells are able to take up berberine, a benzylisoquinoline alkaloid, from the medium and transport it exclusively into the vacuoles. Uptake activity depends on the growth phase of the cultured cells whereas the culture medium had no effect on uptake. Treatment with several inhibitors suggested that berberine uptake depended on the ATP level. Some inhibitors of P‐glycoprotein, an ABC transporter involved in multiple drug resistance in cancer cells, strongly inhibited berberine uptake, whereas a specific inhibitor for glutathione biosynthesis and vacuolar ATPase, bafilomycin A1, had little effect. Vanadate‐induced ATP trap experiments to detect ABC proteins expressed in C. japonica cells showed that three membrane proteins of between 120 and 150 kDa were photolabelled with 8‐azido‐[α‐32P] ATP. Two revealed the same photoaffinity‐labelling pattern as P‐glycoprotein, and the interaction of these proteins with berberine was also demonstrated. These results suggest that ABC proteins of the MDR‐type are involved in the uptake of berberine from the medium.
Article
Strictosidine, a precursor to over 1000 indole alkaloids including the anti-tumor drugs vinblastine, vincristine, and camptothecin, is produced by the condensation of tryptamine and secologanin. Strictosidine synthase, the enzyme responsible for this condensation, is the first committed step in the indole-alkaloid pathway. We have introduced a modified cDNA encoding Strictosidine synthase from Catharanthus roseus (L.) Don. (McKnight et al. 1990, Nucl. Acids Res. 18, 4939) driven by the CaMV 35S promoter into tobacco (Nicotiana tabacum L.). Transgenic tobacco plants expressing this construct had from 3 to 22 times greater strictosidinesynthase activity than C. roseus plants. Ultrastructural immunolocalization demonstrated that strictosidine synthase is a vacuolar protein in C. roseus and is correctly targeted to the vacuole in transgenic tobacco. Immunoblot analysis of strictosidine synthase showed that two distinct forms of the enzyme were produced in transgenic tobacco plants but that only a single form was made in C. roseus. This observation indicates that the second form of the protein is not simply a result of overexpression in tobacco, but may reflect differences in protein processing between tobacco and C. roseus.
Article
ChemInform is a weekly Abstracting Service, delivering concise information at a glance that was extracted from about 200 leading journals. To access a ChemInform Abstract of an article which was published elsewhere, please select a “Full Text” option. The original article is trackable via the “References” option.
Article
The molecular characterization of CYP72A1 from Catharanthus roseus (Madagascar periwinkle) was described nearly a decade ago, but the enzyme function remained unknown. We now show by in situ hybridization and immunohistochemistry that the expression in immature leaves is epidermis-specific. It thus follows the pattern previously established for early enzymes in the pathway to indole alkaloids, suggesting that CYP72A1 may be involved in their biosynthesis. The early reactions in that pathway, i.e. from geraniol to strictosidine, contain several candidates for P450 activities. We investigated in this work two reactions, the conversion of 7-deoxyloganin to loganin (deoxyloganin 7-hydroxylase, DL7H) and the oxidative ring cleavage converting loganin into secologanin (secologanin synthase, SLS). The action of DL7H has not been demonstrated in vitro previously, and SLS has only recently been identified as P450 activity in one other plant. We show for the first time that both enzyme activities are present in microsomes from C. roseus cell cultures. We then tested whether CYP72A1 expressed in E. coli as a translational fusion with the C. roseus P450 reductase (P450Red) has one or both of these activities. The results show that CYP72A1 converts loganin into secologanin.
Article
The molecular characterization of CYP72A1 from Catharanthus roseus (Madagascar periwinkle) was described nearly a decade ago, but the enzyme function remained unknown. We now show by in situ hybridization and immunohistochemistry that the expression in immature leaves is epidermis-specific. It thus follows the pattern previously established for early enzymes in the pathway to indole alkaloids, suggesting that CYP72A1 may be involved in their biosynthesis. The early reactions in that pathway, i.e. from geraniol to strictosidine, contain several candidates for P450 activities. We investigated in this work two reactions, the conversion of 7-deoxyloganin to loganin (deoxyloganin 7-hydroxylase, DL7H) and the oxidative ring cleavage converting loganin into secologanin (secologanin synthase, SLS). The action of DL7H has not been demonstrated in vitro previously, and SLS has only recently been identified as P450 activity in one other plant. We show for the first time that both enzyme activities are present in microsomes from C. roseus cell cultures. We then tested whether CYP72A1 expressed in E. coli as a translational fusion with the C. roseus P450 reductase (P450Red) has one or both of these activities. The results show that CYP72A1 converts loganin into secologanin.
Article
As the final downstream product of the genome, the plant metabolome is a highly complex, dynamic assortment of primary and secondary compounds. Although technological platforms to study genomes, transcriptomes and even proteomes are presently available, methods to pursue genuine metabolomics have not yet been developed due to the extensive chemical diversity of plant primary and secondary metabolites. No single analytical method can accurately survey the entire metabolome. However, recent technical, chemometric and bioinformatic advances promise to enhance our global understanding of plant metabolism. Separation-based mass spectrometry (MS) approaches, such as gas (GC) or liquid chromatography (LC)-MS, are relatively inexpensive, highly sensitive and provide excellent identifying capacity. However, Fourier transform-ion cyclotron resonance (FT-ICR)-MS is better suited for rapid, high-throughput applications and is currently the most sensitive method available. Unlike MS-based analyses, nuclear magnetic resonance (NMR) spectroscopy provides a large amount of information regarding molecular structure, and novel software innovations have facilitated the unequivocal identification and absolute quantification of compounds within composite samples. Due to the size and complexity of metabolomics datasets, numerous chemometric methods are used to extract and display systematic variation. Coupled with pattern recognition techniques and plant-specific metabolite databases, broad-scope metabolite analyses have emerged as functional genomics tools for novel gene discovery and functional characterization. In this review, key metabolomics technologies are compared and the applications of FT-ICR-MS and NMR to the study of benzylisoquinoline alkaloid metabolism in opium poppy are discussed.
Article
Hyoscyamine 6β-hydroxylase catalyses the oxidative reactions in the biosynthetic pathway leading from hyoscyamine to scopolamine. The hydroxylase gene of Hyoscyamus niger was placed under the control of the cauliflower mosaic virus 35S promoter and introduced to hyoscyamine-rich Atropa belladonna by a binary vector system using Agrobacterium rhizogenes. The presence of the transgene in kanamycin-resistant hairy roots was confirmed by the analysis using polymerase chain reaction. The engineered belladonna hairy roots showed increased amounts and enzyme activities of the hydroxylase, and contained up to five-fold higher concentrations of scopolamine than wild-type hairy roots. The contents of 6β-hydroxyhyoscyamine also increased in the transformed roots. Such genetically engineered hairy roots should be useful for enhancing scopolamine productivity in in vitro root culture systems.