ArticlePDF AvailableLiterature Review

Updates on artemisinin: an insight to mode of actions and strategies for enhanced global production

Authors:

Abstract and Figures

Application of traditional Chinese drug, artemisinin, originally derived from Artemisia annua L., in malaria therapy has now been globally accepted. Artemisinin and its derivatives, with their established safety records, form the first line of malaria treatment via artemisinin combination therapies (ACTs). In addition to its antimalarial effects, artemisinin has recently been evaluated in terms of its antitumour, antibacterial, antiviral, antileishmanial, antischistosomiatic, herbicidal and other properties. However, low levels of artemisinin in plants have emerged various conventional, transgenic and nontransgenic approaches for enhanced production of the drug. According to WHO (2014), approximately 3.2 billion people are at risk of this disease. However, unfortunately, artemisinin availability is still facing its short supply. To fulfil artemisinin's global demand, no single method alone is reliable, and there is a need to collectively use conventional and advanced approaches for its higher production. Further, it is the unique structure of artemisinin that makes it a potential drug not only against malaria but to other diseases as well. Execution of its action through multiple mechanisms is probably the reason behind its wide spectrum of action. Unfortunately, due to clues for developing artemisinin resistance in malaria parasites, it has become desirable to explore all possible modes of action of artemisinin so that new generation antimalarial drugs can be developed in future. The present review provides a comprehensive updates on artemisinin modes of action and strategies for enhanced artemisinin production at global level.
Content may be subject to copyright.
REVIEW ARTICLE
Updates on artemisinin: an insight to mode of actions
and strategies for enhanced global production
Neha Pandey
1
&Shashi Pandey-Rai
1
Received: 19 January 2015 /Accepted: 16 March 2015
#Springer-Verlag Wien 2015
Abstract Application of traditional Chinese drug,
artemisinin, originally derived from Artemisia annua L., in
malaria therapy has now been globally accepted.
Artemisinin and its derivatives, with their established safety
records, form the first line of malaria treatment via artemisinin
combination therapies (ACTs). In addition to its antimalarial
effects, artemisinin has recently been evaluated in terms of its
antitumour, antibacterial, antiviral, antileishmanial,
antischistosomiatic, herbicidal and other properties.
However, low levels of artemisinin in plants have emerged
various conventional, transgenic and nontransgenic ap-
proaches for enhanced production of the drug. According to
WHO (2014), approximately 3.2 billion people are at risk of
this disease. However, unfortunately, artemisinin availability
is still facing its short supply. To fulfil artemisinins global
demand, no single method alone is reliable, and there is a need
to collectively use conventional and advanced approaches for
its higher production. Further, it is the unique structure of
artemisinin that makes it a potential drug not only against
malaria but to other diseases as well. Execution of its action
through multiple mechanisms is probably the reason behind
its wide spectrum of action. Unfortunately, due to clues for
developing artemisinin resistance in malaria parasites, it has
become desirable to explore all possible modes of action of
artemisinin so that new generation antimalarial drugs can be
developed in future. The present review provides a compre-
hensive updates on artemisinin modes of action and strategies
for enhanced artemisinin production at global level.
Keywords Artemisia annua L.Artemisinin .Malaria .
Cancer .Mode of action .Yield enhancement
Introduction
Artemisinins (artemisinin and its chemical derivatives), a fam-
ily of unique sesquiterpene trioxane with endoperoxide
bridge, is surely a naturesgift to the mankind. Artemisinins
are derived from an Asteraceae member Artemisia annua and
has long been used in traditional medication in the form of
herbal preparations to cure intermittent fevers. With the redis-
covery of artemisinin in 1970s, its tremendous potential to
cure malaria infections have been established (Hsu 2006). In
the present scenario, WHO recommended that artemisinin
combination therapies (ACT) form the backbone of the global
struggle to cure malaria. According to WHO (2014), 97 coun-
tries worldwide had ongoing malaria transmission, and ap-
proximately 3.2 billion people are at risk of this disease.
Pharmacological action of artemisinin is now no more restrict-
ed to treat malaria but to other medical conditions such as
various cancers, inflammatory diseases, viral (e.g. Human
cytomegalovirus), protozoal (e.g. Toxoplasma gondii), hel-
minths (Schistosom sp., Fasciola hepatica)andfungal
(Cryptococcus neoformans) infections (Ho et al. 2013). Very
recently, an artemisinin derivative, artesunate has been shown
to inhibit autoimmune arthritis (Hou et al. 2014) and possess
antihistaminic activity (Favero Fde et al. 2014). The unique-
ness of artemisinin as potential drug for various diseases lies
in its complex structure. Various studies have revealed that
Handling Editor: Peter Nick
*Shashi Pandey-Rai
shashi.bhubotany@gmail.com
1
Laboratory of Morphogenesis, Department of Botany, Banaras
Hindu University, Varanasi, India
Protoplasma
DOI 10.1007/s00709-015-0805-6
there is no single standard mode of action that exists for
artemisinin. It executes its action through multiple mecha-
nisms, and this could be the probable reason for its wide spec-
trum of action against many diseases. Being a frontline treat-
ment for malaria and a promising drug for cancer treatment,
there are continuous efforts to elucidate antimalarial and anti-
cancer mechanisms of action for artemisinin. Therefore, there
is a need to review different modes of action in order to carry
forward the research of development and application of
artemisinin in a better way.
Artemisinin has a low ratio of its production to demand.
The various factors contributing to short falls in artemisinin
availability are its low concentration in plants (less than 0.01
to 1.4 % of the plant dry weight) (Liu et al. 2006) and depen-
dence of its yield on environmental variables such as temper-
ature, humidity and soil types (White 2008). Owing to its
complex structure, chemical synthesis is very uneconomical
and low yielding (Geldre et al. 1997). All these factors cumu-
latively make artemisinin relatively expensive causing diffi-
culty to fulfil demand of over 392 million courses of ACT
(equivalent to 180 metric tonnes of artemisinin) each year
(WHO 2014). Various efforts have been made, and research
is still on to find out the most suitable way to enhance
artemisinin production. However, any individual method is
not enough to produce higher artemisinin at global level,
and orchestrating various yield enhancing strategies is desir-
able to speed up the artemisinin supply to needy. The present
review summarizes recent progress on artemisininsmecha-
nism of action and strategies for enhanced artemisinin produc-
tion globally.
Mode of action of artemisinin
No single mechanism exists for artemisininsaction. It func-
tions through multiple modes. Despite of having plenty of
biological properties, artemisininsmodes of action have been
widely investigated in terms of its antimalarial and anticancer
effects. There exist a number of reviews dealing with mecha-
nism of antimalarial/anticancer individually. This review pro-
vides a compilation of latest advances in both action mecha-
nisms together.
Antimalarial modes of action
The exact mechanism of action of artemisinin is still a debated
topic. Multiple antimalarial modes of action have been postu-
lated for artemisinin (Fig. 1). Prior to its action, artemisinin
needs to be activated that generate free radical species. The
nature of artemisinin activatoris a bit controversial. Various
groups have demonstrated a covalent reaction between
artemisinin and haemoglobin-bound iron (Selmeczi et al.
2004;Kannanetal.2005; Messori et al. 2006); however, its
probability was denied due to steric hindrance caused by
bulky artemisinin when approaching heme moiety of
haemoglobin (Creek et al. 2009). Afterwards, it has been sug-
gested that ferrous heme (heme-Fe
2+
) is probably the most
efficient activator of artemisinin (Zhang and Gerhard
2008).
Two most acceptable models for artemisinin activation are
(i) reductive scission model and (ii) open peroxide model
(ONeill et al. 2010). In reductive scission, ferrous
heme/non-heme exogenous Fe
2+
bind to artemisinin and via
reductive scission of the peroxide bridge produce oxygen-
centred radicals which self-arranges to give carbon-centred
radicals. Further, iron-peroxide interaction occurs in different
ways to form either primary carbon-centred radicals (via C
3
-
C
4
bond scission) or secondary carbon-centred radicals (via 1,
5 H-shifts). In contrast, open peroxide model suggests in-
volvement of a Fanton reaction of Fe
2+
and the endoperoxide
to produce secondary carbon-centred radicals (Haynes et al.
2007) that are short-lived and are damaging to parasitesvital
biomolecules (Edikpo et al. 2013). This carbon-centred rad-
ical theorymay be strengthened by the fact that artemisinin
activity greatly diminished in the presence of free radical scav-
engers and antioxidants (Hamacher-Brady et al. 2011)where-
as pro-oxidants enhanced its activity (Krungkrai and
Yuthawang 1987). A recent report by Gopalakrishnan and
Kumar (2014) revealed that artesunate (an artemisinin deriv-
ative) induces DNA double-strand breaks in P. falciparum
with simultaneous increase in intercellular ROS level in the
parasite ultimately causing parasite death. In addition to oxi-
dative stress-mediated parasite damage, artemisinins have
been suggested to covalently interact with essential
plasmodial biomolecules (Asawamahasakda et al. 1994).
One such identified molecule is PfTCTP, a histamine releasing
factor that is involved in regulation of cell cycle, histamine
release, malignant transformation, immunological functions
(Bommer and Thiele 2004) as well as a target molecule in
tumour reversion and malaria treatment (Telerman and
Amson 2009; Bhisutthibhan et al. 1998). Though artemisinin
interacts covalently by alkylation of TCTP, it was unclear
whether it might function as target for artemisinin; however,
recently, Eichhorn et al. (2013) demonstrated a role of TCTP
in mediating artemisinin effect owing to its very high Kd
values (77120 μM) that are several orders of magnitude
higher than the artemisinin IC
50
s for parasite growth
inhibition.
A decade ago, another potential target for artemisinin ac-
tion was identified as PfATP6, the P. falciparum sarco-
endoplasmic reticulum calcium ATPase (SERCA) (Eckstein-
Ludwig et al. 2003). It reduces cytosolic free Ca
2+
by
pumping two Ca
2+
into the endoplasmic reticulum, in ex-
change for <4H
+
, a mechanism critical for parasite survival
(Brini and Carafoli 2009). Of various Ca
2+
-ATPases in
Plasmodium, only one enzyme orthologous to SERCA;
N. Pandey, S. Pandey-Rai
PfATP6 is found in P. falciparum (Eckstein-Ludwig et al.
2003) that consists of 10 transmembrane α-helix and three
cytosolic domainsthe effector (A) and the phosphorylation
(P) that are connected to the M domain and third, the
nucleotide-binding domain (N) that is connected to P domain
(Brini and Carafoli 2009). Eckstein-Ludwig et al. (2003)have
also suggested that mechanism of artemisinin-mediated
PfATP6 inhibition is highly specific, not affecting any other
malarial transporter including the non-SERCA Ca
2+
ATPase
(PfATP4) and possibly function through allostericmechanism
(Shandilya et al. 2013). Further experiments dealing with mu-
tation of a single residue (L263) that directly modulates
PfATP6 sensitivity to artemisinin suggest a potential resis-
tance mechanisms (Uhlemann et al. 2005). Artemisinin resis-
tance is generally characterized by a delayed clearance of ma-
larial parasite. Efforts are being implicated for using PfATP6
as a molecular marker for artemisinin resistance. However,
mutations in the Kelch 13 (K13) propeller domain have re-
cently been demonstrated as important determinants for
artemisinin resistance both in vivo and in vitro (Ariey et al.
2014;Straimeretal.2015). Mok et al. (2015) conducted pop-
ulation transcriptomics of 1043 human malaria parasites and
revealed that elevated expression of unfolded protein response
(UPR) pathways including the major PROSE and TRiC chap-
eron complexes is the possible underlying mechanism behind
artemisinin resistance. Further, they provided the mechanistic
evidence that artemisinin-resistant parasites upregulate UPR
pathways in order to overcome the protein damage caused by
artemisinin.
Recently, Hartwig et al. (2009) have demonstrated that
artemisinin may cause parasite membrane damage by accu-
mulating within its neutral lipids and suggested the role of
endoperoxide moiety of artemisinin behind such effects.
Experiments were conducted on Sacchramyces cereviciae
where targeted deletion of genes encoding mitochondrial
NADH dehydrogenases (NDE 1 or NDI 1) resulted in resis-
tance to inhibitory effects of artemisinin whereas their over-
expression increased sensitivity to artemisinin, suggesting a
Fig. 1 Different antimalarial
modesofactionofartemisinin
Artemisinin: modes of action and its enhanced global production
mode of action associated with mitochondrial functions (Li
et al. 2005). This idea was further attested by Wang et al.
(2010), demonstrating artemisinin-induced mitochondrial
membrane depolarization via locally generated reactive oxy-
gen species (ROS) that lead to mitochondrial malfunctioning
and ultimately parasite death. Additional mechanistic evi-
dence was provided by the study conducted on mammalian
cells with non-functional electron transport chain where ac-
tively respiring mitochondria have been shown to generate
reactive oxygen species that plays central role in
endoperoxide-induced cytotoxicity of artesunate (Mercer
et al. 2011).
Within the host erythrocytes, parasite degrades host
haemoglobin into hemeand globinmoiety through a series
of proteases in to its food vacuole. Further, catabolism releases
short peptides and amino acids which are needed for parasite
nutrition. This releases a toxic by-product hematinwhich
undergoes bio-mineralization to form non-toxic hemozoin
(malaria pigment). It has been proposed that Fe
2+
-heme-acti-
vated artemisinin interfere this catabolic pathway through
heme alkylation at α,βand δ-carbon atoms that later create
toxicity to parasite (Meunier et al. 2010; Cazelles et al. 2001).
Two proteins related to this catabolic pathwayheme detox-
ification protein (HDP) and histidine-rich protein II (HRP
II)have been posited as targets of artemisinin action
(Chugh et al. 2013).
An interesting demonstration regarding interaction of heme
and mitochondria with artemisinin has recently been revealed
(Sun et al. 2015). It has been suggested that heme and mito-
chondria playdistinct roles for mediating artemisinin damage.
In tumour cells, heme pathway has been shown to mediate
artemisinins killing whereas antimalarial action of artemisinin
is driven by specific mitochondrial pathways.
To further uncover the artemisininsmode of action, gene
expression analysis was undertaken that revealed alterations in
two genes (encoding heat shock protein, HSP70, and a hypo-
xanthine phosphoribosyltransferase, PF10_0121, related to
purine biosynthesis) associated with decreased sensitivity to
artemisinin. However, whether or not these proteins are direct-
ly a part of artemisinin mode of action is unclear.
Anticancer modes of action
Similar to its antimalarial actions, artemisinins exert their an-
ticancer effects through endoperoxide bridge-dependent ROS
generation and subsequent oxidative damage to the target
cell(s) (ONeill et al. 2010). Many workers have established
ROS-dependent artemisinin effects. Further boost to this the-
ory has recently been given by Lu et al. (2014)where
dihydroartemisinic acid (DHA) significantly reduced viability
of human colorectal cancer HCT-116 cells in vitro through G1
cell arrest, apoptotic cell death and ROS accumulation. A
consensus opinion is that artemisinin and its derivatives exert
their therapeutic effects through multiple modes (Fig. 2).
Though oxidative damage remains in the prime focus, italone
cannot describe everything about anticancer actions of
artemisinins. This hypothesis was strengthened when DHA
was demonstrated to activate p38 MAPK pathway in an
ROS-independent way (Lu et al. 2008) through inhibiting
vascular endothelial growth factor (VEGF)-induced endothe-
lial cell migration (Guo et al. 2014). Further refinements were
brought by workers who revealed an anticancer specific action
of artemisinin through modulation of transferrin receptor-1
(TfR1), a type II transmembrane protein (Ba et al. 2012).
TfR1 is required for iron import in to the cells by binding to
diferric transferrin and subsequent internalization of the Tf-
TfR1 complex through receptor-mediated endocytosis (Ponka
and Lok 1999). For rapid proliferation of cancer cells,
high iron concentration is required; therefore, TfR1
overexpresses in cancer. DHA deplete cellular iron
levels via palmitoylation and subsequent lipid raft-
mediated non-classical endocytosis of TfR1, leading to
iron deficiency in cancer cells (Ba et al. 2012).
To uncover the probable molecular network governing an-
ticancer mechanisms of artemisinins, recent efforts by Huang
et al. (2013) have revealed key signalling pathways of
artemisinin actions including Pi3k-akt, T cell receptor, toll-
like receptor and TGF-beta signalling pathways. Further, dif-
ferent targets of artemisinins from these pathways were also
predicted. Pi3k-akt (phosphatidylinositol-3 kinasesAkt), an
intracellular signalling pathway, is activated in cancers and
plays central role in cancer cell progression and survival
(Courtney et al. 2010). However, a current report by Chen
et al. (2014) suggests that dihydroartemisinin inhibits glioma
cell proliferation and invasion possibly via suppressing
disintegrin and metalloproteinase 17 (ADAM17) levels and
downregulating epidermal growth factor receptor (EGFR)-
Pi3k-akt signalling. Consequently, Pi3k-akt pathway is being
developed as a potential drug target to cure cancers.
Artemisinins, recently emerged potent anticancer drugs, target
this pathway as a major mechanism of action and thereby
efficiently regulate various cellular processes like cell growth,
proliferation, survival and immunity probably through
targeting several membrane receptors such as EGFR, EPOR,
IGF1R, IL2RA, IL4R, ITGAV, KDR, KIT, MET, NGFR,
PDGFRB and TLR4 (Huang et al. 2013). Additionally,
artemisininsderivatives such as DHA and artesunate were
also postulated to induce apoptosis in prostate cancer cells
and regulate immunological responses via modulating Pi3k-
akt pathway (He et al. 2010;Xuetal.2007).
In addition to Pi3k-akt, more targeted approaches for can-
cer treatment involve inhibition of molecular/signalling path-
ways that are crucial for tumour cell progression and mainte-
nance, and hence, artemisinins are making their place for ad-
vance cancer therapies owing to its capability to target toll-like
receptors (TLRs), TFG-beta and T cell signalling pathways.
N. Pandey, S. Pandey-Rai
TLRs are integral membrane protein receptors which are be-
lieved to promote tumour growth and progression (Muccioli
et al. 2012). Similarly, due to their growth promoting abilities,
TFG-beta (transforming growth factor-beta) and T cell recep-
tor signalling pathways are believed as excellent targets for
treatment of cancers and other diseases (Akhurst and Hata
2012; Watanabe et al. 2014). Artemisinin-mediated targets
for T cell receptor signalling pathways have been predicted
as CD28, CD4, CD8A and CTLA4 whereas for toll-like re-
ceptor and TGF-beta signalling pathways, TLR4 and
TGFBR1, respectively, are predicted targets (Huang et al.
2013). A recent report by Lee et al. (2014) suggests that
artesunate, an artemisinin derivative, has a negative mitogenic
effect on CD+ T cells as it inhibits CD4+ T cell proliferation
and IL-2 (T cell growth factor) production and reduction in
expression of cell surface protein CD25 (IL-2 receptor alpha
chain) and CD69 on CD4+ T cells.
Emerging resistance to artemisinin: an unfortunate
condition
Although tremendous efforts have been implicated to eluci-
date possible mechanisms of action, many points are still con-
troversial and unclear which are needed to be looked for.
While attempts have been made and are still in process to
eliminate malaria, unfortunately, spread of artemisinin resis-
tance in malarial parasites has imperilled efforts for malaria
control and elimination. As per the latest WHO report
(February 2015), artemisinin resistance has been confirmed
in Combodia, the Lao Peoples Democratic Republic,
Myanmar, Thailand and Viet Nam, and many other regions
are at high risk of emergence of multidrug resistance. Of var-
ious factors that have contributed towards evolution of
artemisinin resistance, widely available oral artemisinin
monotherapies and other substandard antimalarial drugs have
played major roles. WHO along with other organizations are
working with the strategy to remove oral artemisinin-based
monotherapies and substandard antimalarial drugs from the
market and emphasizing the use of artemisinin combination
therapies to check further spread of drug resistance. However,
this requires strong financial capital, long-term political com-
mitments and cross-border cooperation. Besides various ma-
laria elimination and control policies that is being developed,
it is now highly desirable to uncover depth of every possible
mechanism underlying artemisinin action so as to open new
doors for development of accurate and potential new genera-
tion antimalarial drugs and modes of malaria treatment in case
of further spreading of artemisinin resistance in future.
Strategies for enhanced artemisinin production
Extensive efforts have been made to improve artemisinin syn-
thesis through many approaches. The present review provides
compilation of recent advancement in both conventional and
modern approaches of yield enhancement such as (i) metabol-
ic engineering of the plant using various genetic engineering
tools; (ii) regulation of artemisinin biosynthesis through stress
signals to plants, hairy root cultures and cell cultures and (iii)
conventional, mutational and molecular breeding of A. annua.
Fig. 2 Proposed model for
various anticancer modes of
action of artemisinin
Artemisinin: modes of action and its enhanced global production
Covello (2008) had suggested the implementation of mix
mode production system involving both biological and chem-
ical synthesis, as a good option for artemisinin production.
Besides various approaches for improved artemisinin synthe-
sis, recent success has been achieved in its semi-synthetic
production with the completion of The Bill and Melinda
Gates Foundation funded The semi-synthetic Artemisinin
Projectin a partnership between University of California
(Berkeley, USA), Amyris Inc. and the Institute for One
World Health (a non-profit pharmaceutical company and
now known as PATH Drug Solutions). The project involved
combined use of metabolic engineering of microorganisms
and synthetic biology to produce semi-synthetic artemisinin.
They achieved 25 gm per litre artemisinic acid production
from engineered S. cerevisiae (Bakers yeast) as well as 40
45 % conversion of artemisinic acid into artemisinin (Paddon
et al. 2013). Details of the complete scheme and approaches
adopted for semi-synthetic artemisinin production have been
very recently reviewed by the project leaders themselves
(Paddon and Keasling 2014). Even though successful produc-
tion of semi-synthetic artemisinin to supplement the plant-
derived production, research is still on to produce WHO rec-
ommended potent artemisinin derivatives at large scale to be
used in ACTs. Therefore, it is still highly desirable not to be
dependent solely on semi-synthetic artemisinin, and efforts
should be implemented to enhance drug production at global
level by other means as well.
It is interesting to note that dried leaves of A. annua con-
taining artemisinin and artemisinin synergistic flavonoids are
more effective to cure malaria than a comparable dose of pure
artemisinin as ACT oral therapy (Weathers et al. 2014). As
discussed earlier in this review, use ofcombination therapy (as
ACT) was promoted to overcome the increasing resistance to
existing antimalarial drugs. However, very recently, consump-
tion of dried A. annua leaves has been demonstrated to over-
come existing resistance to pure artemisinin in Plasmodium
yoelii (Elfawal et al. 2015). These recent reports promote the
idea of increasing artemisinin synthesis through in-planta
yield enhancement over the semi-synthetic production of pure
artemisinin for better and more effective cure of human
malaria.
Artemisinin biosynthetic pathway and overexpression
of concerned pathway gene(s)
Like other terpenes, artemisinin biosynthesis involves two
pathways, the cytosolic mevalonate (MVA) pathwaystem-
ming from acetyl co-A and the plastidic non-mevalonte/MEP
pathwayoriginating from glyceraldehyde-3-phosphate and
pyruvate. These two pathways result in the formation of two
isoprenoid precursors, isopentenyl diphosphate (IPP) and di-
methyl allyl diphosphte (DMAPP). These two precursors de-
rived from two pathways do not lead rest of the pathway
separately and independently, rather it has been demonstrated
that cytosolic DMAPP (mevalonate origin) is transferred to
the plastid and it is the plastidic IPP unit that form the central
isoprenoid unit of majority of fernesyl diphosphate (FPP) en-
gaged in artemisinin biosynthesis (Schramek et al. 2009). Two
stages of artemisinin biosynthesis have now been completely
elucidated: first stage involves the formation of FPP while
second is the cyclization of FPP to form amorpha-4,11-diene,
which ultimately lead to the synthesis of artemisinin and its
various derivatives, all steps catalyzed by a series of enzymes.
Key enzymes involved are as follows: HMGS, HMGR, DXS,
DXR, FPS, ADS, CYP71AV1, DBR2 and ALDH1 (Fig. 3).
All the biosynthetic pathway genes have been shown to be
differentially expressed in various plant parts. Glandular tri-
chomes of A. annua have been well characterized as the site of
artemisinin biosynthesis and tissues with high density of glan-
dular trichomes such as flower buds and young leaves show
relatively higher expression of biosynthetic pathway genes
(Olofsson et al. 2011). In contrast to the all enzyme-
catalyzed steps of artemisinin biosynthesis, the conversion
of dihydroartemisinic acid to artemisinin is, however, a non-
enzymatic photooxidative step. It has been hypothesized that
in A. annua, there are many oxygenated sesquiterpenoids in-
cluding artemisinin, which are formed by spontaneous autox-
idation of terpene precursors via highly reactive allyllic hydro-
peroxide intermediate (Brown 2010).
Manipulations of biosynthetic pathways to channelize
the carbon flux for enhanced synthesis of a particular
metabolite involves either upregulation of desired path-
way or downregulation of competing pathways through
overexpression and suppression of their respective path-
way genes, respectively.
Overexpression of key biosynthetic genes for hyperproduc-
tion of precursors of different stages and hence greater syn-
thesis of product (artemisinin) is an excellent idea. Much suc-
cess has been achieved in engineering almost all the key genes
from artemisinin biosynthetic pathway. Basic chemical skele-
ton of artemisinin involves fusion of three C-5 isoprenoid unit
(IPP or DMAPP). Two important key enzymes leading to C-5
isoprenoid unit are HMGR and DXR. HMGR shunts HMG
Co-A into the mevalonate pathway, leading to the synthesis of
isoprenoid unit in cytosol, whereas in plastidic non-
mevalonate pathway DXR catalyzes the first committed step
of isoprenoid (C5) synthesis by converting 1-deoxy-D-
xylulose-5-phosphate in to 2-C-methyl-D-erythritol-4-
phosphate.
These two key genes have been overexpressed in A. annua
by many workers. Agrobacterium-mediated HMGR gene
transfer from Catheranthus roseus to A. annua was first per-
formed by Aquil et al. (2009), and this resulted in 38.9 %
higher artemisinin level in one of the transgenic line.
However, when HMGR was co-expressed along with ADS
gene, artemisinin enhanced up to 7.65-folds in one of the
N. Pandey, S. Pandey-Rai
transgenic line as compared to non-transgenic plants (Alam
and Abdin 2011).
Isotopologue (
13
CO
2
) profiling (Schramek et al. 2009)and
fosmidomycin (MEP pathway blocker) treatment (Towler and
Weathe rs 2007) confirmed DXR to be an important regulator
of artemisinin biosynthesis. Further, genetic map (Graham
et al. 2010) itself did not bring direct evidence for DXR to
be regulator of artemisinin biosynthesis; however, it
pointed the fact that DXR locus co-localizes with QTL
for artemisinin yield and concentration. Recently, CaMV
35S promoter driven overexpression of DXR gene in
A. annua depicted 1.212.35-folds higher artemisinin
in transgenic plants (Xiang et al. 2012).
C5 isoprenoid unit thus produced (via actions of key en-
zymes HMGR and DXR) undergo two sequential 1-4 conden-
sations catalyzed by fernesyl diphosphate synthase (FPS) to
give C15 fernesyl diphosphate (FPP), which is further utilized
to produce variety of sesquiterpenoids and triterpenoids in-
cluding artemisinin. Thus, overexpression of FPS gene may
probably contribute to greater metabolic flux towards
artemisinin synthesis and both heterologous (from
Gossypium arboreum) and homologous (from A. annua)
FPS gene overexpression resulted in 23-folds higher
artemisinin content in transgenic lines in comparison to
wild-type plants (Chen et al. 2000;Banyaietal.2010).
Further, when FPS gene was coexpressed with HMGR gene
(both were homologous), they boosted artemisinin content
1.8-fold in transgenic plants.
The first committed step of artemisinin biosynthesis is the
cyclization of FPP in to amorpha-4,11-diene catalyzed by
amorpha-4,11-diene synthase (ADS). Thus, the overexpres-
sion of ADS is a promising approach for upregulated
artemisinin biosynthesis in A. annua. Tang et al. (2008)cloned
and expressed ADS gene in A. annua and observed 2.3-folds
more artemisinin as compared to control. However, spraying
of artemisinic acid to A. annua plants reduced ADS transcript
level suggesting a possibility of some feedback inhibition on
artemisinin biosynthesis (Arsenault et al. 2010a)Further,
Fig. 3 Transgenic approaches for enhanced artemisinin biosynthesis in
A. annua. (3-hydroxy-3-methylglutaryl-CoA synthase: HMGS, 3-
hydroxy-3-methylglutaryl-CoA reductase: HMGR, 1-deoxyxylulose 5-
phosphate synthase: DXS, 1-deoxyxylulouse 5-phosphate
reductoisomerase: DXR, amorpha-4,11-diene synthaase: ADS,
cytochrome P 450 CYP71AV1: CYP, double bond reductase 2: DBR2,
dihydroartemisinic aldehyde reductase 1: RED1, aldehyde
dehydrogenase 1: ALDH1, farnesyl diphosphate synthase: FPS,
squalene synthase: SQS, caryophyllene synthase: CPS)
Artemisinin: modes of action and its enhanced global production
diversion of amorpha-4,11-diene towards artemisinin synthe-
sis is carried out by cytochrome P450 monooxygenase
(CYP71AV1) that catalyzes three-step sequential conversion
of amorpha-4,11-diene to artemisinin alcohol, then artemisinic
aldehyde, and ultimately to artemisinic acid. The activity of
CYP71AV1 also requires a reducing companion cytochrome
P450 oxidoreductase (CPR), which is considered to be co-
expressed with CYP71AV1 (Ro et al. 2006). In addition,
Chen et al. (2012) tried to co-overexpress CYP71AV1 and
CRP with FPS gene. This led to 3.6-fold greater accumulation
of artemisinin in FPS+CYP71AV1+CPR co-overexpressing
transgenic lines in comparison to wild-type plants. In a similar
approach, CYP71AV1 and CPR were co-overexpressed with
ADS and resulted in 2.4-fold higher artemisinin in one of the
transgenic line as compared to control plants. Further in the
artemisinin biosynthetic pathway, reduction of artemisinic al-
dehyde in to dihydroartemisinic aldehyde has been proposed
to be an important step which is catalyzed by artemisinic al-
dehyde Δ11(13) reductase (DBR2). Very recently, Yuan et al.
(2014)madeanefforttooverexpressDBR2 gene driven by the
CaMV 35S promoter and observed remarkable increase in
artemisinin and its direct precursor DHA, as confirmed by
HPLC analysis. In addition, DBR2-overexpressing lines also
produced more arteannuin B and its direct precursor
artemisinic acid.
All the important genes encoding key enzymes of upstream
and downstream artemisinin biosynthetic pathway have been
cloned and transgenic lines overexpressing single or multiple
genes have so far resulted in 1.8to 7.6-folds elevated
artemisinin production, suggesting overexpression of pathway
genes as a promising approach to effectively improve
artemisinin synthesis.
Indirect upregulation of artemisinin biosynthetic pathway
genes
The expression of genes is largely regulated by specific tran-
scription factors (TFs). Therefore, overexpression of TFs of-
fers an alternative and complementary strategy to upregulate
biosynthetic pathway genes (Fig. 3). Promoters of three key
genes of artemisinin biosynthesis, i.e. ADS,CYP71AV1 and
DBR2, have been characterized till date. The first ever tran-
scription factor of A. annua that has been characterized is
AaWRKY1 (a WRKY type transcription factor) and is pro-
posed to bind to the W-boxes in ADS and CYP71AV1 pro-
moters. Transient expression of AaWRKY cDNA in leaves
of A. annua was shown to increase the transcript level of the
majority of artemisinin biosynthetic genes (Ma et al. 2009)
and AaWRKY-overexpressing transgenic lines produced
4.4-folds higher artemisinin as compared to control lines
(Tang et al. 2012). Recently, in a more targeted approach,
trichome-specific overexpression of AaWRKY1 resulted in
more effective activation of CYP71AV1 transcription as
compared to the constitutive overexpression of AaWRKY
(Han et al. 2014). However, this did not lead to any significant
improvement in transcript level of other key genes such as
FDS,ADS and DBR2 and promoted artemisinin content 1.8
times as compared to wild-type plants. Further, CRTD
REHVCBF2 (CBF2) and RAV1AAT (RAA) motifs in pro-
moters of both ADS and CYP71AV1 have been proposed to
be the binding sites of two TFs, AaERF1 and AaERF2, which
belong to JA-responsive AP2 family of TF. Both the TFs were
cloned from A. annua and overexpression of either of the two
promoted ADS and CYP71AV1 transcripts level and
artemisinin as well. The fourth TF that has been successfully
cloned and overexpressed in A. annua is AaORA which is a
trichome-specific APETALA2/ethylene response factor (AP2/
ERF) family TF. Its overexpression in A. annua significantly
boosted transcription of ADS,CYP71AV1,DBR2 and
AaERF1. AaORA-overexpressing and AaORA-RNAi trans-
genic lines showed significant increase (53 %) and decrease in
artemisinin content, respectively, as compared to control (Lu
et al. 2013). Recently, a fifth TF, AabHLH1, a bHLH TF has
been successfully isolated from a cDNA library of glandular
secretary trichomes (GSTs) of A. annua and has been
demonstrated to be able to bind to the E-box-cis ele-
ment in the promoter of ADS and CYP71AV1.Transient
expression of AabHLH1 in the leaves of A. annua sig-
nificantly enhanced transcript levels of ADS,CYP71AV1
and HMGR (Jietal.2014).
In addition to TF engineering, artemisinin biosynthesis
may also be improved by blocking/silencing artemisinin com-
petitive pathway/genes and hence ensuring diversion of com-
mon isoprenoid precursors more towards artemisinin biosyn-
thetic pathway (Fig. 3). So far, two competitive pathway en-
zymes have been engineered in A. annua: squalene synthase
(SQS) and β-caryophyllene synthase (CPS). RNA interfer-
ence (RNAi) offers a potential reverse geneticsapproach to
knock down the expression of particular gene(s) and in
A. annua Zhang et al. (2009) applied RNAi technology for
the first time to knock down the expression of SQS. SQS
catalyzes the first committed step in sterol biosynthetic path-
way by converting FPP into squalene and thus competes with
ADS for utilizing FPP in the biosynthesis of sterols.
Suppression of SQS expression in A. annua through hairpin
RNA-mediated RNAi technique reported a 3.14-fold higher
artemisinin content along with significantly lower sterol level
in transgenic plants as compared to control. Another gene that
has been artificially modulated for its expression in A. annua
is CPS. CPS, being a sesquiterpene synthase, competes with
ADS for utilizing FPP and converts FPP into β-caryophyllene.
Using antisense RNA technology, Chen et al. (2011)triedto
suppress the expression of CPS gene by Agrobacterium-me-
diated insertion of 750-bp antisense fragment of CPS cDNA
into A. annua via pBI121 plant expression vector. This
lowered the endogenous CPS expression and elevated
N. Pandey, S. Pandey-Rai
artemisinin content by 54.9 % in one of the transgenic line as
compared to wild-type plants.
Regulation of artemisinin biosynthesis through stress
signals
Various environmental stresses in a moderate dose and dura-
tion can act as a potential elicitor for enhanced secondary
metabolite production. When plants recognize stress signals
at cellular level, a stress response is induced. A number of
researchers have exposed A. annua to a number of stress sig-
nals and had reported alterations in artemisinin content along
with transcript level of its biosynthetic genes which is sum-
marized in Table 1. Environmental stress signals such as tem-
perature (low or high), salinity, water (drought or flooding),
radiations, chemical, mechanical and various kinds of biotic
stresses often result in induced accumulation of
phenylpropanoids (Dixon and Paiva 1995). In recent years,
many of them have been explored in terms of their potential
to boost artemisinin level in A. annua. Chilling has been dem-
onstrated to affect artemisinin content in A. annua.Yangetal.
(2010) had revealed chilling-mediated induction in
artemisinin level along with its biosynthetic genes such as
ADS,CYP and DXS. Similarly, night frost also increased
artemisinin content (Wallaart et al. 2000). Influence of salinity
on artemisinin accumulation has been investigated by differ-
ent groups. Qureshi et al. (2005) showed that salt treatment
(0160 mM) increased artemisinin content during the early
phase of plant growth, possibly due to sudden non-
enzymatic conversion of artemisinic acid/DHA (artemisinin
precursors) in to artemisinin under increasing oxidative stress.
However, at later stages, it reduced artemisinin. In contrast,
when A. annua seedlings were subjected to salinity stress (4
6 g/l NaCl) significantly boosted artemisinin (23%dry
weight) compared to non-treated plants (1 % dry weight)
(Qian et al. 2007). Role of nutrient deficiency in artemisinin
production was first reported by Ferreira (2007). Higher level
of artemisinin was found in plants grown in potassium defi-
ciency as compared to plants grown with full supplementation
of macronutrients and had suggested that artemisinin produc-
tion per hectare may be enhanced by growing A. annua under
a mild potassium deficiency. In contrast, supplementation of
few elements reportedly enhances artemisinin content. Mild
boron stress (1.0 mM) enhanced artemisinin as compared to
control (Aftab et al. 2010). Similarly, arsenic (As) (3000 μg/l)
for 7 days and cadmium (Cd) treatment (20 and 100 μmol/l) to
Tabl e 1 Different stress signals for enhanced artemisinin biosynthesis in-planta
Stress signals for artemisinin enhancement Experiments conducted on Reference
Chilling Soil-grown plants/In vitro propagated plants Yang et al. 2010;Luluetal.2008
Night frost Soil-grown plants Wallaart et al. 2000
Salinity Soil-grown plants Qian et al. 2007; Qureshi et al. 2005
Potassium deficiency Soil-grown plants Ferreira 2007
Boron Soil-grown plants Aftab et al. 2010
Arsenic Soil grown plants/hydroponic culture Rai et al. 2011a; Paul and Shakya 2013
Phosphorus Soil grown plants Kapoor et al. 2007
Cadmium Hydroponic culture Li et al. 2012
Drought Soil-grown plants Marchese et al. 2010
Post harvest drying Soil grown plants Laughlin 2002
Miconazole Hydroponic culture Towler and Weathers 2007
UV-B In vitro propagated plants Pandey and Pandey-Rai 2014a
UV-C Soil-grown plants Rai et al. 2011b
Jasmonic acid Exogenous application to soil-grown plants Wang et al. 2009;Maesetal.2011
Salicylic acid Exogenous application to soil-grown plants Pu et al. 2009; Guo et al. 2010
Phytohormones (GA3+BAP+ ABA) Soil-grown plants Maes et al. 2011
Abscisic acid (alone) Pot culture of in vitro germinated seedlings Jing et al. 2009
Sugars In-vitro propagated plants Wang and Weathers; Arsenault et al. 2010b
Methyl jasmonate (MeJA)+Miconzole Cell suspension culture Caretto et al. 2011
MeJA (alone) Hairy root culture Patra et al. 2013
Oligosaccharide elicitor (OE) of fungal origin Hairy root culture Zheng et al. 2008;Wangetal.2006; Wang et al. 2009
Yeast extract+MeJA+Chitosan Hairy root culture Putalun et al. 2007
MeJA+ cell homogenate of Piriformospora indica Hairy root culture Ahlawat et al. 2014
Artemisinin: modes of action and its enhanced global production
hydroponically grown A. annua promoted synthesis and ac-
cumulation of artemisinin (Rai et al. 2011a;Lietal.2012).
RT-PCR analysis revealed upregulated transcript levels of
HMGR,FDS,ADS and CYP71AV1 genes under arsenic stress
(Rai et al. 2011a). In tune with boron and As, phosphorus also
enhanced artemisinin level (Kapoor et al. 2007). Not all the
metal stress signals affect artemisinin positively. A re-
cent study by Paul and Shakya (2013) has demonstrated
that As(III) (at 5 and 7.5 μg/ml) and NaCl favour
growth of A. annua and artemisinin biosynthesis while
higher doses of As(III) (10 μg/ml) and Cr(VI) even at
mild doses have adverse effects on growth and
artemisinin level (Paul and Shakya 2013).
There are reports in support of drought stress affecting
artemisinin accumulation (Marchese et al. 2010). Water deficit
of 38 h (W=1.39 MPa) significantly increased both leaf and
plant artemisinin up to 29 % without causing any serious harm
to biomass production. This suggests that pre-harvest water
deficit may reduce time and costs in crop drying along with
gain in artemisinin content. Further, different ways of post-
harvest drying of crops also play role in overall artemisinin
availability. Field crop drying is beneficial over artificial dry-
ing (such as oven-drying) and gives higher artemisinin; how-
ever, mode of post-harvest drying does not seem to affect
artemisinic acid level (Laughlin 2002). The exact mechanism
behind such abiotic factor-mediated elevation in artemisinin is
still not clear. However, a common factor behind these stresses
is the ROS-mediated oxidative stress which may probably be
one of the reasons for greater artemisinin accumulation.
Though there are many reports of deciphering roles of abi-
otic stress signals in influencing artemisinin content in
A. annua, relatively fewer work have been conducted to depict
artemisinin regulation through biotic stress signals. A study by
Kapoor et al. (2007)showedthatinoculationbytwo
arbuscular mycorrhizal fungi, Glomus macrocarpum and
Glomus fassiculatum, significantly enhanced artemisinin con-
centration density of leaf glandular trichomes (GT) which are
the accumulation sites of artemisinin. They also observed a
strong positive linear correlation between leaf GT density and
artemisinin, suggesting that certain biotic stress signals may
be potential tools to improve in-planta improvement of
artemisinin synthesis and accumulation. Few more biotic elic-
itors have been demonstrated to improve artemisinin but in
hairy root culture which will be discussed later in this section.
Exogenous application of signalling molecules such as
jasmonic acid (Wang et al. 2009;Maesetal.2011)and
salicylic acid (Pu et al. 2009; Guo et al. 2010) also has role
in boosting biosynthesis of artemisinin probably due to in-
duced burst of ROS and subsequent conversion of DHA in
to artemisinin as well as upregulation of key artemisinin bio-
synthetic genes. In addition, few phytohormones such as
gibberellic acid cytokinins (BAP) and abscisic acid also boost
artemisinin level (Maes et al. 2011;Jingetal.2009).
Elicitation and plant cell/tissue culture
Plants respond quickly to even a small fluctuation in its phys-
ical or chemical environment. These fluctuations are very well
known to elicit the plants secondary metabolism and hence
are exploited to get the desired secondary metabolite. Plant
tissue culture offers an opportunity to grow plants in vitro in
a controlled environment and manipulation of plant physical
and chemical environment to get the desired product. Rapidly
increasing world population is putting more and more pres-
sure on availability of cultivation land. Therefore, in the pres-
ent scenario, there is a need to explore plant cell/tissue culture
technique in integration with elicitation to get higher second-
ary metabolite production with minimum land usage. This will
reduce the land use by medicinal plants which are largely
grown for production of pharmaceuticals and other phyto-
chemicals and better utilization of lands for various other
motives.
There are established standard protocols for in vitro prop-
agation of A. annua. It can easily be propagated through
micro-cuttings in MS media. Since accumulation of secondary
metabolites in plants in response to abiotic factors is a com-
mon phenomenon, manipulation of abiotic environment of
in vitro propagated A. annua plantlets may lead to differential
accumulation of artemisinin. We applied this hypothesis in our
recent work and had demonstrated that UV-B radiation expo-
sure (3 h) to in vitro propagated A. annua plantlets can double
the artemisinin content (Pandey and Pandey-Rai 2014a)with-
out causing any serious damage to physiology including pho-
tosynthesis and hence survival (Pandey and Pandey-Rai
2014b). Chilling (4 °C) treatment for 30 min also affected
artemisinin content as well as transcript level of ADS (11-
fold) and CYP (7-fold) genes in in vitro cultured plants as
revealed by HPLC and qPCR analysis, respectively (Lulu
et al. 2008). Different sugar supplementation during in vitro
plantlet development also affects artemisinin content in plants.
Seedlings were inoculated in GamborgsB5mediumcontain-
ing3%(w/v) sucrose, glucose or fructose. Fructose inhibited
artemisinin synthesis while seedlings growing in 3 % glucose
produced more artemisinin (Wang and Weathers 2007). These
results were further confirmed by Arsenault et al. (2010)
through expression analysis of biosynthetic genes as well as
quantification of artemisinin and its various derivatives.
Various efforts have been made to get higher artemisinin
through integrated approach of elicitation with hairy root or
cell suspension culture in A. annua. Among these, hairy root
culture seems advantageous over cell suspension culture in
terms of high growth rate, better genetic stability and indepen-
dency of hormone supplemented media requirement for
growth (Guillon et al. 2006). Plant signalling compounds,
precursors, inhibitors and various other kinds of molecules
have been explored in terms of their potential to elicit
artemisinin biosynthesis. Caretto et al. (2011) investigated
N. Pandey, S. Pandey-Rai
efficacy of methyl jasmonate (MeJA) and miconazole on
artemisinin biosynthesis in cell suspension cultures of
A. annua. Three-fold increment in just 30 min of MeJA
(22 μmol) treatment and 2.5-fold after 24 h of miconazole
treatment were observed. However, miconazole treatment se-
verely affected cell viability. Earlier, this antifungal compound
miconazole significantly induced artemisinin level in
A. annua seedlings (Towler and Weathers 2007).
Miconazole-mediated induction in artemisinin is due to its
ability to inhibit sterol biosynthesis and redirection of carbon
flux towards terpenoid pathway (Zarn et al. 2003). Using
MeJA alone (40 μg/l, 15 days) as elicitor, 3.45 mg/g
artemisinin was produced (Patra et al. 2013). Nitric oxide
(NO) also stimulates artemisinin synthesis in hairy root cul-
ture. Zheng et al. (2008) studied that NO generation induced
by an oligosaccharide elicitor (OE) from Fusarium oxysporum
mycelium elevated artemisinin from 0.7 to 1.3 mg/g DW by
treating 20-day-old hairy roots with OE for 4 days. Further,
combined treatment of OE with NO donor sodium nitroprus-
side increased artemisinin from 1.2 to 2.2 mg/g DW. Various
other elicitors have been reported to stimulate artemisinin bio-
synthesis in hairy root culture such as OE from Colletotrichum
gloeosporioides (Wang et al. 2006), cerebroside (Wang et al.
2009), yeast extract, methyl jasmonate and chitosan (Putalun
et al. 2007). Very recently, in an interesting approach of com-
bined application of abiotic and biotic elicitor (MeJA and cell
homogenate of Piriformospora indica) to hairy root culture,
2.44 times greater artemisinin was obtained (Ahlawat et al.
2014). This increase was probably due to upregulation of bio-
synthetic genes. They also reported for the first time a positive
correlation between elicitor application and expression of bio-
synthetic pathway genes viz. HMGR,ADS,CYP71V1,ALDH1,
DXR,DXS and DBR2 in hairy root cultures of A. annua.
Combined approaches utilizing in vitro techniques and
elicitation is a promising idea for large-scale production of
artemisinin through bioreactor cultivation technology.
Breeding approaches for enhanced artemisinin
production
According to the factsheets on the World Malaria Report 2014,
issued by WHO (December 2014), in the year 2013, there
were an estimated 198 million cases of malaria (uncertainty
range 124283 million) and an estimated 584,000 deaths (un-
certainty range 367,000755,000) in the world, of which 90 %
deaths occur in Africa. Further, the second most affected part
of the world is South-East Asia where India has the highest
malaria burden, followed by Indonesia and Myanmar. These
facts suggest that major risk and burden of this disease is inthe
developing countries which are generally with insufficient ad-
vanced infrastructure/techniques. Therefore, despite of various
efforts to produce high artemisinin by diverse means, there is
still a need to think on ground level as well because
agricultural production is not a problem or limiting factor in
developing countries. Integrating cultivation of A. annua with
different breeding techniques may be a viable alternative to
produce more artemisinin thereby aiding in overall enhanced
production of artemisinin globally. By bringing Artemisia
annua into cultivation, conventional and new biotechnological
plant-breeding techniques can be applied at the genetic level to
improve yield and uniformity with stable high performing
phenotypes across adverse/variable environments and to mod-
ify desired valuable compounds. Conventional plant breeding
practices include collection/creation of population or germ-
plasm with useful desired genetic variation, identification of
superior individuals and development of improved variety
from selected individuals. Success of conventional breeding
is dependent on the selection process. Numerous selection
methods can be adopted for A. annua. Mass selection is de-
pendent mainly on selection of plants according to their phe-
notypes and performance and can be used to improve the
overall population of A. annua by positive or negative mass
selection. One drawback of mass selection is the influence of
the environment on the development, phenotype and perfor-
mance of single plants. Another method is the recurrent selec-
tion, which is more suitable for the A. annua as it is cross-
pollinating species. In this method, the seed from selected
plants is not added together but is kept apart and used to
perform offspring tests. Through different selection strategies,
certain high yielding varieties of A. annua such as CIM-
ArogyaJeevan Rakshaand Asha were released by Central
Institute for Medicinal and Aromatic Plants (CIMAP), India,
as superior lines rich in artemisinin (Patra and Kumar 2005).
Recently, Townsend et al. (2013) have reported that selection
of material for breeding using combining ability analysis of a
diallele cross can be used for the identification of elite parents
to produce improved A. annua hybrids. This selection method
was reportedly consistent with advanced QTL-based molecu-
lar breeding approaches.
Modernization of plant breeding using different biotechno-
logical tools is opening new avenues for crop improvement.
Two such tools are mutation and molecular breeding (Fig. 4).
Mutation breeding involves induced mutation through
chemicals or radiation with the advantage of improving one
or two characters without modifying the rest of genotypes.
A successful breeding effort was made by Mediplant (a
swiss not-for-profit organization) by developing high-
yielding F1 hybrid populations of A. annua known as
Artemiswith mean annual artemisinin production of about
32 kg/ha. Further, they also created new high yielding hybrids
with 40.552.0 kg/ha artemisinin (Simonnet et al. 2008).
There has been significant progress in the development of
biotechnological plant-breeding techniques using various mo-
lecular tools. Different DNA markers (for example, RFLP,
RAPD, AFLP, SSR, SNP, STMS etc) and functional markers
(such as EST, microarray and qRT-PCR) can be variously
Artemisinin: modes of action and its enhanced global production
used to speed up the selection for recognition of desired ge-
notypes at an early stage. Although these marker-assisted se-
lections are being applied in various crops, there are relatively
few studies of molecular marker-based approaches in medic-
inal plants. One such effort was made by CIMAP (India) for
developing high-artemisinin variety CIM-arogya, through
marker-assisted selection breeding.
For the production of high-yield varieties of A. annua,afast
track molecular breeding project known as The CNAP
Artemisia Research Projecthas been initiated at the Centre
for Novel Agricultural Products (CNAP) which is a part of
university of York. This project led by CNAPs Director
Dianna Bowles and Deputy Director Ian Graham was funded
by Bill & Melinda Gates Foundation. With the aim to produce
high yielding non-GM variety of A. annua, over 23,000 pa-
rental lines were screened for desired traits and 768 different
hybrid crosses were made, of which 268 most promising hy-
brids were forwarded to field trials. After rigorous selection
procedure, two best performing hybrids viz. Hyb1209r
(Shennong) and Hyb8001r (Zenith) with 36.3 and 54.5 kg/ha
artemisinin have been commercially released. Using various
molecular markers, they also developed the genetic map of
A. annua with nine linkage groups (Graham et al. 2010). For
this, they used the Artemis (high yielding variety) pedigree and
established genetic linkage and QTL maps. Positive QTL re-
lated to artemisinin yield were also independently validated.
Conclusion and future directions
Various approaches have been adopted for improvement in the
yield of artemisinin. However, it is still a challenge for any
individual approach alone to meet the global demand of
artemisinin. Though recent achievement of semi-synthetic
production of artemisinin by Keaslings group is a major
breakthrough in bringing artemisinin production at industrial
level, there still exist many hurdles for sustainable production
of this drug to be used in ACTs. Its insecure supply leads to
price hike and to reduce the drug cost there is a need to speed
up the artemisinin production worldwide.
Therefore, in-planta yield enhancement through various
ways as discussed in this review remains the prime focus for
sustainable drug supply. It is worth mentioning that the all
Fig. 4 Different proposed
breeding approaches in A. annua
for enhanced production of
artemisinin
Fig. 5 Intercropping model with special reference to Indian cropping for
growing A. annua
N. Pandey, S. Pandey-Rai
transgenic or non-transgenic tools for enhancement of
artemisinin biosynthesis can be more fruitful if implemented
on high yielding chemotypes/hybrids of A. annua. Integration
of conventional methods with advanced techniques will serve
to extend and enhance the continued production of drug. As
discussed earlier, more than 90 % burden of malaria is in
African and South-East Asian developing countries. As ma-
jority of these countries have agricultural economies, agri-
based enhancement of artemisinin production in-planta can
be a promising approach for overall improved production.
Various agricultural strategies can be posited depending on
the local scenario of agricultural practices in different coun-
tries. Deficiency in the supply chain of artemisinin can be
remediated by integrating agri-based technologies with mo-
lecular tools for affordable drug production. As for example
in India, we hope that intercropping of high yielding varieties
of A. annua (developed through molecular/mutation breeding)
with staple food crops can be good option for its cultivation
(Fig. 5) without engaging additional cultivable land and mean-
time execution of multiharvesting and advanced drug extrac-
tion practices may result in greater artemisinin production
with lower production cost. Similar approaches may also be
applied in other countries as well which can give a further
boost to artemisinin production at the same time bettering life
standards of their farmers.
Keeping in mind, benefits and limitations associated with
every yield enhancement strategies, we emphasize that to
bring artemisinin supply up to the level of its demand requires
its production collectively through semi-synthesis, in-planta
yield enhancement by both transgenic and non-transgenic
methods as well as modern breeding.
Conflict of interest The authors (NP and SPR) of the review article
entitled BUpdates on Artemisinin: an insight to mode of actions and
strategies for enhanced global production^have no conflict of interest.
References
Aftab T, Khan MMA, Idrees M, Naeem M, Ram M (2010) Boron in-
duced oxidative stress, antioxidant defence response and changes in
artemisinin content in Artemisia annua L. J Agron Crop Sci 196:
423430. doi:10.1111/j.1439-037X.2010.00427.x
Ahlawat S, Saxena P, Alam P, Wajid S, Abdin MZ (2014) Modulation of
artemisinin biosynthesis by elicitors, inhibitor, and precursor in
hairy root cultures of Artemisia annua L. J Plant Interact 9(1):811
824. doi:10.1080/17429145.2014.949885
Akhurst RJ, Hata A (2012) Targeting the TGFβsignalling pathway in
disease. Nat Rev Drug Discov 11(10):790811. doi:10.1038/
nrd3810
Alam P, Abdin MZ (2011) Over-expression of HMG-CoA reductase and
amorpha-4,11-diene synthase genes in Artimisia annua L. and its
influence on artimisinin content. Plant Cell Rep 30(10):1919
1928. doi:10.1007/s00299-011-1099-6
Aquil S, Husaini AM, Abdin MZ, Rather GM (2009) Overexpression of
HMG-CoA reductase gene leads to enhanced artimisinin biosynthe-
sis in transgenic Artemisia annua plants. Planta Med 75(13):1453
1458. doi:10.1055/s-0029-1185775
Ariey F, Witkowski B, Amaratunga C, Beghain J, Langlois AC, Khim N
et al (2014) A molecular marker of artemisinin resistant Plasmodium
falciparum malaria. Nature 505(7481):5055. doi:10.1038/
nature12876
ArsenaultPR, Vail DR, Wobbe KK, Weathers PJ (2010a) Effect of sugars
on artemisinin production in Artemisia annua L.: transcription and
metabolite measurements. Molecules 15:23022318
Arsenault PR, Vail DR, Wobbe KK, Erickson K, Weathers PJ (2010b)
Reproductive development modulates gene expression and metabo-
lite levels with possible feedback inhibition of artemisinin in
Artemisia annua L. Plant Physiol 154:958968
Asawamahasakda W, Ittarat I, Chang C-C, McElroy P et al (1994) Effects
of antimalarials and protease inhibitors on plasmodial hemozoin
production. Mol Biochem Parasitol 67:183191
Ba Q, Zhou N, Duan J, Chen T, Hao M et al (2012) Dihydroartemisinin
Exerts Its Anticancer Activity through Depleting Cellular Iron via
Transferrin Receptor-1. PLoS ONE 7(8):e42703. doi:10.1371/
journal.pone.0042703
Banyai W, Kirdmanee C, Mii M, Supaibulwatana K (2010)
Overexpression of farnesyl pyrophosphate synthase(FPS) gene af-
fected artemisinin content and growth of Artemisia annua L. Plant
Cell Tiss Org Cult 103:255265. doi:10.1007/s11240-010-9775-8
Bhisutthibhan J, Pan XQ, Hossler PA, Walker DJ, Yowell CA, Carlton J
et al (1998) The Plasmodium falciparum translationally controlled
tumor protein homolog and its reaction with the antimalarial drug
artemisinin. J Biol Chem 273:1619216198
Bommer UA, Thiele BJ (2004) The translationally controlled tumour
protein (TCTP). Int J Biochem Cell Biol 36:379385
Brini M, Carafoli E (2009) Calcium pumps in health and disease. Physiol
Rev 89(4):13411378. doi:10.1152/physrev.00032.2008
Brown GD (2010) The Biosynthesis of Artemisinin (Qinghaosu) and the
Phytochemistry of Artemisia annua L. (Qinghao). Molecules 15:
76037698. doi:10.3390/molecules15117603
Caretto S, Quarta A, Durante M, Nisi R, de Paolis A, Blando F, Mita G
(2011) Methyl jasmonate and miconazole differently affect
arteminisin production and gene expression in Artemisia annua sus-
pension cultures. Plant Biol 13:5158
Cazelles J, Robert A, Meunier B (2001) Alkylation of heme by
artemisinin, an antimalarial drug. C R Acad Sc Se Li Fascicule C
Chim 4:8589
Chen D, Ye H, Li G (2000) Expression of a chimeric farnesyl diphosphate
synthase gene in Artemisia annua L. Plant via Agrobacterium
tumefaciens mediated transformation. Plant Sci 155(2):179185.
doi:10.1016/S0168-9452(00)00217-X
Chen JL, FangHM, Ji YP, Pu GB, Guo YW, HuangLL, Du ZG, Liu BY,
Ye HC, Li GF, Wang H (2011) Artemisinin biosynthesis enhance-
ment in transgenic Artemisia annua plants by downregulation ofthe
beta-caryophyllene synthase gene. Planta Med 77:17591765
Chen Y, Shen Q, Wang Y, Wang T et al (2012) The stacked over-
expression of FPS, CYP71AV1 and CPR genes leads to the increase
of artimisinin level in Artimisia annua L. Plant Biotechnol Rep 7:
287295. doi:10.1007/s11816-012-0262-z
Chen J, Chen X, Wang F, Gao H, Hu W (2014) Dihydroartemisinin
suppresses glioma proliferation and invasion via inhibition of the
ADAM17 pathway. Neurol Sci. doi:10.1007/s10072-014-1963-6
Chugh M, Sundararaman V, Kumar S, Reddy VS et al (2013) Protein
complex directs haemoglobin to hemozoin formation in
Plasmodium falciparum. Proc Natl Acad Sci 110:53925397. doi:
10.1073/pnas.1218412110
Courtney KD, Corcoran RB, Engelman JA (2010) The PI3K Pathway As
Drug Target in Human Cancer. J Clin Oncol 28(6):10751083. doi:
10.1200/JCO.2009.25.3641
Artemisinin: modes of action and its enhanced global production
Covello PS (2008) Making artemisinin. Phytochemistry 69(17):2881
2885. doi:10.1016/j.phytochem.2008.10.001
Creek DJ, Ryan E, Charman WN, Chiu FCK et al (2009) Stability of
peroxide antimalarials in the presence of human haemoglobin.
Antimicrob Agents Chemother 53(8):34963500. doi:10.1128/
AAC. 00363-09
Dixon RA, Paiva N (1995) Stressed induced phenyl propanoid metabo-
lism. Plant Cell 7:10851097. doi:10.1105/tpc.7.7.1085
Eckstein-Ludwig U, Webb RJ, Van GID, East JM et al (2003) Artimisinin
target the SERCA of Plasmodium falciparum. Nature 424:957961.
doi:10.1038/nature01813
Edikpo N, Ghasi S, Elias A, Oguanobi N (2013) Artemisinin and bio-
molecules: The continuing search for mechanism of action. MolCell
Pharmacol 5:7589
Eichhorn T, Winter D, Büchele B, Dirdjaja N, Frank M et al (2013)
Molecular interaction of artemisinin with translationally controlled
tumor protein (TCTP) of Plasmodium falciparum. Biochem
Pharmacol 85:3845. doi:10.1016/j.bcp.2012.10.006
Elfawal MA, Towler MJ, Reich NG, Weathers PJ, Rich SM (2015) Dried
whole-plant Artemisia annua slows evolution of malaria drug resis-
tance and overcomes resistance toartemisinin. Proc Natl Acad Sci
112(3):821-826. doi: 10.1073/pnas1413127112
Favero Fde F, Grando R, Nonato FR, Sousa IM, Queiroz NC, Longato
GB, Zafred RR, Carvalho JE, Spindola HM, Foglio MA (2014)
Artemisia annua L.: evidence of sesquiterpene lactonesfraction
antinociceptive activity. BMC Complement Alternat Med 14(1):
266. doi:10.1186/1472-6882-14-266
Ferreira JF (2007) Nutrient deficiency in the production of artemisinin,
dihydroartemisinic acid, and artemisinic acid in Artemisia annua L. J
Agric Food Chem 55(5):16861694
Geldre EV, Vergauwe A, Eekhout EV (1997) state of the art of production
of antimalarial compound artemisinin in plants. Plant Mol Biol 33:
199209
Gopalakrishnan AM, Kumar N (2014) Anti-malarial action of Artesunate
involves DNA damage mediated by Reactive Oxygen Species.
Antimicrob Agents Chemother. doi:10.1128/AAC. 03663-14
Graham IA, Besser K, Blumar S, Branigan CA et al (2010) The genetic
map of Artemisia annua L. Identifies loci affecting yield of the
antimalarial drug artemisinin. Science 327(5963):328331. doi:10.
1126/science.1182612
Guillon S, Tremouillaux GJ, Pati PK, Rideau M, Gantet P (2006) Hairy
root research: recent scenario and exciting prospects - Commentary.
Curr Opin Plant Biol 9:341346
Guo XX, Yang XQ, Yang RY, Zeng QP (2010) Salicylic acid and methyl
jasmonate but not Rose Bengal enhance artemisinin production
through involving burst of endogenous singlet oxygen. Plant Sci
178:390397
Guo L, Dong F, Hou Y, Cai W, Zhou X, Huang AL, Yang M, Allen TD,
Liu J (2014) Dihydroartemisinin inhibits vascular endothelial
growth factor-induced endothelial cell migration by a p38
mitogen-activated protein kinase-independent pathway. Exp Ther
Med 8(6):17071712. doi:10.3892/etm.2014.1997
Hamacher-Brady A, Stein HA, Turschner S, ToegelI, Mora R, Jennewein
N, Efferth T, Eils R, Brady NR (2011) Artesunate activates mito-
chondrial apoptosis in breast cancer cells via iron-catalyzed lyso-
somal reactive oxygen species production. J Biol Chem 286(8):
65876601. doi:10.1074/jbc.M110.210047
Han J, Wang H, Lundgren A, Brodelius PE (2014) Effects of overexpres-
sion of AaWRKY1 on artemisinin biosynthesis in
transgenic Artemisia annua plants. Phytochemistry 102:8996.
doi:10.1016/j.phytochem.2014.02.011
Hartwig CL, Rosenthal AS, Dangelo J, Griffin CE, Posner GH, Cooper
RA (2009) Accumulation of artemisinintrioxane derivatives within
neutral lipids of Plasmodium falciparum malaria parasites is endo-
peroxide-dependent. Biochem Pharmacol 77(3):322336
Haynes RK, Chan WC, Lung CM, Uhlemann AC et al (2007) The Fe 2+
mediated decomposition, PfATP6, binding and antimalarial activi-
ties of artimisone and other artemisinins: The unlikelihood of C-
centered radicals as bioactive intermediates. ChemMedChem
2(10):14801497. doi:10.1002/cmdc.200700108
He Q, Shi J, Shen XL, An J, Sun H, Wang L, Hu YJ, Sun Q, Fu LC,
Sheikh MS, Huang Y (2010) Dihydroartemisinin upregulates death
receptor 5 expression and cooperates with TRAIL to induce apopto-
sis in human prostate cancer cells. Cancer Biol Ther 9:819824. doi:
10.4161/cbt.9.10.11552
Ho WE, Peh HY, Chan TK, Wong WS (2013) Artemisinins:
Pharmacological actions beyond anti-malarial. Pharmacol Ther
142(1):126139. doi:10.1016/j.pharmthera.2013.12.001
Hou L, Block KE, Huang H (2014) Artesunate abolishes germinal center
B cells and inhibits autoimmune arthritis. PLoS One 9(8):e104762.
doi:10.1371/journal.pone.0104762
Hsu E (2006) Reflections on the discoveryof the antimalarial qinghao.
Br J Clin Pharmacol 61(6):666670. doi:10.1111/j.1365-2125.2006.
02673.x
Huang M, Lu JJ, Huang MQ, Bao JL et al (2013) Terpenoids: natural
products for cancer therapy. Expert Opin Investig Drug 21:1801
1818. doi:10.1517/13543784.2012.727395
Ji Y, Xiao J, Shen Y, Ma D, Li Z, Pu G, Li X, Huang L, Liu B, Ye H,
Wang H (2014) Cloning and characterization of AabHLH1, a
bHLH transcription factor that positively regulates artemisinin bio-
synthesis inArtemisia annua. Plant Cell Physiol 55(9):15921604.
doi:10.1093/pcp/pcu090
Jing F, Zhang L, Li M, Tang Y, Wang Y, Wang Y, Wang Q, Pan Q, Wang
G, Tang K (2009) Abscisic acid(ABA) treatment increases
artemisinin content in Artemisia annua by enhancing the expression
of genes in artemisinin biosynthesis pathway. Biologia 64:319323
Kannan R, Kumar K, Sahal D, Kukreti S et al (2005) Reaction of
artemisinin with haemoglobin: implication for antimalarial activity.
Biochem J 385(2):409418. doi:10.1042/BJ20041170
Kapoor RR, Chaudhary V, Bhatnagar AK (2007) Effects of arbuscular
mycorrhiza and phosphorus application on artemisinin concentra-
tion in Artemisia annua L. Mycorrhiza 17:581587. doi:10.1007/
s00572-007-0135-4
Krungkrai SR, Yuthawang Y (1987) The antimalarial action of
Plasmodium falciparum of quinghaoso and artesunate in combina-
tion with agents which modulate oxidant stress. Trans R Soc Trop
Med Hyg 81:710714
Laughlin JC (2002) Post-harvest drying treatment effects on antimalarial
constituents of Artemisia annua L. Acta Horticult 576:315320
Lee SH, Cho YC, Kim KH, Lee IS, Choi HJ, Kang BY (2014) Artesunate
inhibits proliferation of naïve CD4+ T cells but enhances function of
effector T cells. Arch Pharm Res. doi:10.1007/s12272-014-0491-5
Li W, Mo W, Shen D, Sun L et al (2005) Yeast model uncovers dual roles
of mitochondria in the action of artemisinin. PLoS Genet 1:329
334. doi:10.1371/journal.pgen.0010036
Li X, Zhao M, Guo L, Huang L (2012) Effect of cadmium on photosyn-
thetic pigments, lipid peroxidation, antioxidants, and artemisinin in
hydroponically grown Artemisia annua. J Environ Sci (China)
24(8):15111518
Liu C, Zhao Y, Wang Y (2006) Artemisinin: current state and perspec-
tives for biotechnological production of an antimalarial drug. Appl
Microbiol Biotechnol 72:1120
Lu JJ, Meng LH, Cai YJ, Chen Q, Tong LJ, Lin LP, Ding J (2008)
Dihydroartemisinin induces apoptosis in HL-60 leukemia cells de-
pendent of iron and p38 mitogen-activated protein kinase activation
but independent of reactive oxygen species. Cancer Biol Ther 7:
10171023. doi:10.4161/cbt.7.7.6035
Lu X, Zhang L, Zhang F, Jiang W, Shen Q, Zhang L, Lv Z, Wang G, Tang
K (2013) AaORA, a trichome-specific AP2/ERF transcriptionfactor
of Artemisia annua, is a positive regulator in the artemisinin
N. Pandey, S. Pandey-Rai
biosynthetic pathway and in disease resistance to Botrytis cinerea.
New Phytol 198:11911202
Lu M, Sun L, Zhou J, Yang J (2014) Dihydroartemisinin induces apopto-
sis in colorectal cancer cells through the mitochondria-dependent
pathway. Tumour Biol 35(6):53075314. doi:10.1007/s13277-014-
1691-9
Lulu Y, Chang Z, Ying H, Ruiyi Y, Qingping Z (2008) Abiotic stress-
induced expression of artemisinin biosynthesis genes inArtemisia
annuaL. Chin J Appl Environ Biol 14:001005
Ma DM, Pu GB, Lei CY, MaLQ WHH, Guo YW, Chen JL, Du ZG, Wang
H, Li GF, Ye HC, Liu BY (2009) Isolation and characterization of
AaWRKY1, an Artemisia annua transcription factor that regulates
the amorpha-4,11-diene synthase gene, a key gene of artemisinin
biosynthesis. Plant Cell Physiol 50:21462161
Maes L, Van Nieuwerburgh FC, Zhang Y, Reed DW, Pollier J, Vande
Casteele SR, Inze D, Covello PS, Deforce DL, Goossens A (2011)
Dissection of the phytohormonal regulation of trichome formation
and biosynthesis of the antimalarial compound artemisinin in
Artemisia annua plants. New Phytol 189(1):176189
Marchese J, Ferreira JFS, Rehder VLG, Rodrigues O (2010) Water deficit
effect on the accumulation of biomass and artemisinin in annual
wormwood (Artemisia annua L., asteraceae). Braz J Plant Physiol
22(1):19
Mercer AE, Copple IM, Maggs JL, ONeil PM et al (2011) The role of
heme and the mitochondrion in the chemical and molecular mecha-
nism of mammalian cell death induced by the artemisinin antima-
larials. J Biol Chem 286(2):987996. doi:10.1074/jbc.M110.
144188
Messori L, Gabbiani C, Casini A, Siraqusa M et al (2006) The reaction of
artemisinins with haemoglobin: A unified picture. Bioorg Med
Chem 14(9):29722977. doi:10.1016/j.bmc.2005.12.038
Meunier FA, Nguyen TH, Colasante C, Luo F et al (2010) Sustained
synaptic vesicle recycling by bulk endocytosis contribute to the
maintenance of high rate neurotransmitter release stimulated by
glycerotoxin. J Cell Sci 123:11311140. doi:10.1242/jcs.049296
Mok S, Ashley EA, Ferreira PE, Zhu L et al (2015) Population tran-
scriptomics of human malaria parasites reveals the mechanism of
artemisinin resistance. Science 347(6220):431435. doi:10.1126/
science.1260403
Muccioli M, Sprague L, Nandigam H, Pate M, Benencia F (2012) Toll-
like receptors as novel therapeutic targets for ovarian cancer. ISRN
Oncol 2012: Article ID 642141, 8 pages doi:10.5402/2012/642141
ONeill PM, Barton VE, Ward SA (2010) The Molecular Mechanism of
Action of ArtemisininThe Debate Continues. Molecules 15:
17051721. doi:10.3390/molecules15031705
Olofsson L, Engström A, Lundgren A, Brodelius PE (2011) Relative
expression of genes of terpene metabolism in different tissues of
Artemisia annua L. BMC Plant Biol 11:45. doi:10.1186/1471-
2229-11-45
Paddon CJ, Keasling JD (2014) Semi-synthetic artemisinin: a model for
the use of synthetic biology in pharmaceutical development. Nat
Rev Microbiol 12(5):355367. doi:10.1038/nrmicro3240
Paddon CJ, Westfall PJ, Pitera DJ, Benjamin K, Fisher K, McPhee D,
Leavell MD, Tai A et al (2013) High-level semi-synthetic produc-
tion of the potent antimalarial artemisinin. Nature 496(7446):528
532. doi:10.1038/nature12051
Pandey N, Pandey-Rai S (2014a) Short-term UV-B radiation-mediated
transcriptional responses and altered secondary metabolism of
in vitro propagated plantlets of Artemisia annua L. Plant Cell Tiss
Organ Cult 116:371385. doi:10.1007/s11240-013-0413-0
Pandey N, Pandey-Rai S (2014b) Modulations ofphysiological responses
and possible involvement of defense-related secondary metabolites
in acclimation of Artemisia annua L. against short-term UV-B radi-
ation. Planta 240(3):611627. doi:10.1007/s00425-014-2114-2
Patra NK, Kumar B (2005) Improved varieties and genetic research in
medicinal and aromatic plants (MAPs). In: Kumar A, Mathur AK,
Sharma A, Singh AK, Khanuja SPS (eds) Proceeding of second
national interactive meet on medicinal and aromatic plants. CSIR-
CIMAP, Lucknow, pp 5361
Patra N, Srivastava AK, Sharma S (2013) Study of various factors for
enhancement of artemisinin in Artemisia annua hairy roots. Int J
Chem Eng Appl 4:157160
Paul S, Shakya K (2013) Arsenic, chromium and NaCl induced
artemisinin biosynthesis in Artemisia annua L.: a valuable antima-
larial plant. Ecotoxicol Environ Saf 98:5965
Ponka P, Lok CN (1999) The transferrin receptor: role in health and
disease. Int J Biochem Cell Biol 31:11111137. doi:10.1016/
S1357-2725(99)00070-9
Pu GB, Ma DM, Chen JL, Ma LQ, Wang H, Li GF, Ye HC, Liu BY
(2009) Salicylic acid activates artemisinin biosynthesis in Artemisia
annua L. Plant Cell Rep 7:11271135
Putalun W, Luealon W, De-Eknamkul W, Tanaka H, Shoyama Y (2007)
Improvement of artemisinin production by chitosan in hairy root
cultures of Artemisia annua L. Biotechnol Lett 29:11431146
Qian Z, Gong K, Zhang L, Lv J, Jing F, Wang Y, Guan S, Wang G, Tang
K (2007) A simple and efficient procedure to enhance artemisinin
content inArtemisia annuaL. by seeding to salinity stress. Afr J
Biotechnol 6(12):14101413
Qureshi MI, Israr M, Abdin MZ, Iqbal M (2005) Responses of Artemisia
annua L. to lead and salt-induced oxidative stress. Environ Exp Bot
53:185193
Rai R, Pandey S, Rai SP (2011a) Arsenic-induced changes in morpho-
logical, physiological, and biochemical attributes and artemisinin
biosynthesis in Artemisia annua, an antimalarial plant.
Ecotoxicology 20(8):19001913. doi:10.1007/s10646-011-0728-8
Rai R, Meena RP, Smita SS, Shukla A, Rai SK, Pandey-Rai S (2011b)
UV-b and UV-C pre-treatments induce physilogical changes and
artemisinin biosynthesis in Artemisia annua L.- An antimalarial
plant. J Photochem Photobiol B 105:216225
Ro DK, Paradise EM, Ouellet M, Fisher KJ et al (2006) Production of the
antimalarial drug precursor artimisinic acid in engineered yeast.
Nature 440:940943. doi:10.1038/nature04640
Schramek N, Wang HH, RÖmisch-Margl W, Keil B, Radykewicz T,
WinzenhÖrlein B et al (2009) Artemisinin biosynthesis in growing
plants of Artemisia annua. a CO study. Phytochemitry 71:179187
Selmeczi K, Robert A, Claparols C, Meunier B (2004) Alkylation of
human haemoglobin A
0
by the antimalarial drug artemisinin.
FEBS Lett 556:245248
Shandilya A, Chacko S, Jayaram B, Ghosh I (2013) A plausible mecha-
nism for the anti- malarial activity of artimisinin: A computational
approach. Sci Rep 3:251. doi:10.1038/srep02513
Simonnet X, Quennoz M, Carlen C (2008) New Artemisia annua hybrids
with high artemisinin content. Acta Horticult 769:371373
Straimer J, Gnädig NF, Witkowski B, Amaratunga C, Duru Vet al (2015)
Drug resistance. K13-propeller mutations confer artemisinin resis-
tance in Plasmodium falciparum clinical isolates. Science
347(6220):428431. doi:10.1126/science.1260867
Sun C, Li J, Cao Y, Long G, Zhou B (2015) Two distinct and competitive
pathways confer the cellcidal actions of artemisinins. Microb Cell
2(1):1425. doi:10.15698/mic2015.01.181
TangKX,JingFY,ZhangL,WangGF(2008)Methodforincreasing
artemisinin content by co-transferring genes HMGR and FPS into
Artemisia annua [P] Faming Zhuanli Shenqing. CN101182545A
Tang KX, Jiang WM, Lu X, Qiu B, Wang GF (2012) Overexpression
AaWRYK1 gene increased artemisinin content in Artemisia annua L.
Shanghai Jiao Tong University, China. Patent CN201210249469.X,
14 Novemb 2012
Telerman A, Amson R (2009) The molecular programme of tumour re-
version: the steps beyond malignant transformation. Nat RevCancer
9:206216
Artemisinin: modes of action and its enhanced global production
Towler MJ, Weathers PJ (2007) Evidence of artemisinin production from
IPP stemming from both the mevalonate and the nonmevalonate
pathways. Plant Cell Rep 26:21292136
Townsend T, Segura V, Chigeza G, Penfield T, Rae A (2013) The Use of
Combining Ability Analysis to Identify Elite Parents forArtemisia
annuaF1 Hybrid Production. PLoS ONE 8(4):e61989
Uhlemann AC, Ramharter M, Lell B, Kremsner PG, Krishna S (2005)
Amplification of Plasmodium falciparum multidrug resistance gene
1 in isolates from Gabon. J Infect Dis 192:18301835
Wallaart TE, Pras N, Beekman AC, Quax WJ (2000) Seasonal variation
of artemisinin and its biosynthetic precursors in plants of Artemisia
annua of different geographical origin: proof for the existence of
chemotypes. Planta Med 66:5762
Wang YY, Weathers PJ (2007) Sugars proportionately affect artemisinin
production. Plant Cell Rep 26:10731081
Wang JW, Zheng LP, Tan RX (2006) The preparation of an elicitor from a
fungal endophyte to enhance artemisinin production in hairy root
cultures of Artemisia annua L. Sheng Wu Gong Cheng Xue Bao 22:
829834
Wang JW, Zheng LP, Zhang B, Zou T (2009) Stimulation of artemisinin
synthesis by combined cerebroside and nitric oxide elicitation in
Artemisia annua hairy roots. Appl Microbiol Biotechnol 85(2):
285292
Wang J, Huang L, Li J, Fan Q et al (2010) Artemisinin directly targets
malarial mitochondria through its specific mitochondrial activation.
PLoS One 5:e 9582. doi:10.1371/journal.pone.0009582
Watanabe M, Moon KD, Vacchio MS, Hathcock KS, Hodes RJ (2014)
Downmodulation of Tumor Suppressor p53 by T Cell Receptor
Signaling Is Critical for Antigen-Specific CD4
+
T Cell Responses.
Immunity 40(5):681691. doi:10.1016/j.immuni.2014.04.006
Weathers PJ, Elfawal MA, Towler MJ, Acquaah-Mensah GK, Rich SM
(2014) Pharmacokinetics of artemisinin delivered by oral consump-
tion of Artemisia annua dried leaves in healthy vs. Plasmodium
chabaudi-infected mice. J Ethnopharmacol 153(3):732736
White NJ (2008) Qinghaosu (Artemisinin) The price of success. Science
320:331334. doi:10.1126/science.1155165
WHO (2014) Facts sheet on the world malaria report 2013. http://www.
who.int/malaria/media/world_malaria_report_2013/en/
WHO (2015) WHO updates on artemisinin resistance (February 2015).
http://www.who.int/malaria/media/artemisinin_resistance_qa/en/
Xiang L, Zeng L, Yuan Y, Chen M et al (2012) Enhancement of
artimisinin biosynthesis by overexpressing dxr, cyp71av1 and cpr
in the plants of Artemisia annua L. Plant Omics 5:503507
Xu H, He Y, Yang X, Liang L, Zhan Z, Ye Y, Yang X, Lian F, Sun L
(2007) Anti-malarial agent artesunate inhibits TNF-alpha-induced
production of proinflammatory cytokines via inhibition of NF-
kappaB and PI3 kinase/Akt signal pathway in human rheumatoid
arthritis fibroblast-like synoviocytes. Rheumatology 46:920926.
doi:10.1093/rheumatology/kem014
Yang RY, Zeng XM, Lu YY, Lu WJ, Feng LL, Yang XQ, Zeng QP (2010)
Senescent leaves ofArtemisia annuaare one of the most active or-
gans for overexpression of artemisinin biosynthesis responsible
genes upon burst of singlet oxygen. Planta Med 76:734742
Yuan Y, Liu W, Zhang Q, Xiang L,Liu X, Chen M, Lin Z, Wang Q, Liao
Z (2014) Overexpression of artemisinic aldehyde Δ11 (13) reduc-
tase gene-enhanced artemisinin and its relative metabolite biosyn-
thesis in transgenic Artemisia annua L. Biotechnol Appl Biochem.
doi:10.1002/bab.1234
Zarn JA, Brüschweiler BJ, Schlatter JR (2003) Azole fungicides affect
mammalian steroidogenesis by inhibiting sterol 14-α-demethylase
and aromatase. Environ Health Perspect 111:255261
Zhang S, Gerhard G (2008) Heme activates artemisinin more efficiently
than hemin, inorganic iron or haemoglobin. Bioorg Med Chem 16:
78537861. doi:10.1016/j.bmc.2008.02.034
Zhang L, JingFY, Li FP, Li MY, Wang YL, Wang GF, Sun XF, Tang KX
(2009) Development of transgenic Artemisia annua(Chinese worm-
wood) plants with an enhanced content of artemisinin, an effective
anti-malarial drug, by hairpin-RNA-mediated gene silencing.
Biotechnol Appl Biochem 52:199207
Zheng LP, Guo YT, Wang JW, Tan RX (2008) Nitric oxide potentiates
oligosaccharide-induced artemisinin production in Artemisia annua
hairy roots. J Integr Plant Biol 50:4955
N. Pandey, S. Pandey-Rai
... 적절한 수준의 환경 스트레스의 처리로 유도된 식물의 방어기작을 목적 물질의 생합성 증진 가능성을 확인하였다. 이에 정교하게 조절된 환경스트레스 처리는 식물의 생장저해로 인한 수량감소 없이 artemisinin 의 합성을 유도할 수 있다 (Pandey and Pandey-Rai 2015). 물리 적 처리에 의한 hormesis 유도는 재배환경 및 생산 식물체에 화학적 오염으로부터 안전하게 이차대사물의 생산 및 축적 을 높이기 위한 재배생산환경 제어 방식으로 식물공장에 간 편하게 적용할 수 있다. ...
... Artemisinin 생합성 효소들 중에서 ADS (amorpha-4,11-diene synthase), CYP (cytochrome P450 mono-oxygenase), ALDH1 (alcohol dehydreogeanse I)는 과발현을 통한 유전공학적 방법 으로 artemisinin 생합성의 주효소로 확인되었다 (Lv et al. 2016;Pandey and Pandey-Rai 2015). 이외에 artemisinin 생합성이 활 발한 trichome에서 3종의 유전자들의 높은 발현량과 artemisinin 생합성과의 연관성이 관찰하였다 (Olofsson et al. 2012 (Fig. 2-A). ...
... Artemisinin, a sesquiterpene trioxane lactone with an endoperoxide bridge (Martínez et al., 2014) has been used in traditional medication to cure intermittent fevers and malaria for over 2000 years (Pandey and Pandey-Rai, 2016). According to World Health Organization (WHO), artemisinin combination therapies form the backbone of the global struggle to cure malaria. ...
... The maximum day temperatures of 35 • C was reported to be most suitable for flowering (Bagchi et al., 1997). Because chemical synthesis of artemisinin is uneconomical and low yielding (Geldre et al., 1997;Pandey and Pandey-Rai, 2016) and biosynthetic processes have proven to be the most efficient synthetic methods to produce it (Tang et al., 2018), field production of A. annua continues to be the main source of artemisinin (Hommel, 2008). Meeting high global demand of artemisinin, therefore, requires expansion of the area cultivated with A. annua cultivars containing high artemisinin levels, enhanced methods for its cultivation and processing and more efficient techniques/methods for extraction and purification of artemisinin from the plant parts (Wetzstein et al., 2018). ...
Article
Artemisinin, the main active compound in Artemisia annua L., has been used for antimalarial properties for centuries and currently attracting increasing interest for its antiviral activities. In addition, several recent publications indicated that this valuable compound can be effective on Sars-CoV-2 virus. In the study, a high-performance liquid chromatography (HPLC) method was optimized in terms of mobile phase compositions, column temperature and flow rate using response surface methodology for the determination of artemisinin from A. annua samples. The method was also validated for some parameters according to the Eurachem guideline. Validated method was applied on A. annua plant samples, cultivated in a controlled condition, and content of artemisinin was found in the range from 5825 to 7972 mg/kg (n=20). Extraction conditions of artemisinin from the plant samples were also optimized. In the first step of the extraction, solvents with different polarities were applied to the samples for the evaluation of artemisinin solubility. Then, ethanol was chosen for extraction solvent due the high extraction yield and classification in safer chemical ingredients list by EPA. After the extraction, a purification step using various adsorbents was studied to remove remnant impurities such as chlorophyl. The results showed that powdered charcoal was found to be the most effective adsorbent. Amount of the adsorbent was also studied to evaluate for the reduction of chlorophyl without reducing the artemisinin concentration. Finally, purified solvent was dewaxed, evaporated, and dried under nitrogen to concentrate the artemisinin content. In conclusion, the optimized conditions could be regarded as a new alternative technique in pharmaceutical industry for the extraction of artemisinin from A. annua samples.
... Previous studies have found that several transcription factor (TF) families can participate in the regulation of artemisinin biosynthesis, including AP2/ERF (Lu et al., 2013), HD-ZIP-IV , WRKY , TCP , MYB , bHLH and bZIP gene family, which can directly or indirectly upregulate or downregulate the expression of artemisinin biosynthesis genes, or increase the density of GSTs by regulating the glandular hair initiation and development. Numerous studies suggest that artemisinin biosynthesis is influenced by multiple environmental factors, including environmental stress, plant hormones, and various biological, chemical, and physical signal-inducing factors (Lu et al., 2000;Maes et al., 2011;Mannan et al., 2010;Pandey and Pandey-Rai, 2016). For instance, exogenous application of ABA and MeJA can upregulate the expression levels of the four key enzyme-encoding genes such as AaADS, AaCYP71AV1, AaDBR2 and AaALDH1, resulting in an increase in artemisinin content in A. annua plants . ...
... Artemisinin derivatives mainly worked on the membrane structure of the malaria parasite, and induced cytotoxicity by producing reactive oxygen species (ROS), which can cause depolarization of mitochondria and plasma membrane [7]. It was also found that DHA contained an endoperoxide group which was critical to the generation of free carbon-centred intermediates that could kill malarial parasite via oxidative stress [8]. Apart from malarial, artemisinins had broad-spectrum inhibitory effects on pathogenic parasites, such as protozoan parasites schistosomiasis [9]. ...
Article
Full-text available
Dihydroartemisinin (DHA) is a derivative of artemisinin and is toxic to parasites. We used the Tetrahymena thermophila (T. thermophila) as a model to explore DHA toxicity. Results showed that low concentration of DHA (20 μmol/L) promoted cell proliferation, whereas high concentrations of DHA (40-1280 μmol/L) inhibited that. Appearance of nucleus was pycnosis by laser scanning confocal microscope. DHA significantly elevated activities of SOD and GSH-Px (P < 0.01) and MDA was markedly increased at high level but decreased at low level (P < 0.01). Further results of transcriptome in T. thermophila treated with different concentration DHA group (0, 20, 160 μmol/L) showed that differentially expressed genes (DEGs) were involved in oxidation-reduction and metabolism of exogenous substances indicated oxidative stress stimulation. Kyoto Encyclopedia of Genes and Genomes showed that DEGs were involved in the cytochrome P450-mediated metabolism of exogenous substances, glutathione metabolism and ABC transport. Remarkably, DNA replication was significantly enriched in low concentration DHA, energy metabolism related pathways and necrotic process were considerably enriched in high concentration DHA. The results of RT-qPCR of 13 DEGs were the same as that of transcriptome, in which the expression of GST and GPx family genes were significantly altered after exposed to high-DHA group. DHA induced oxidative stress damage through disturbing with energy. However, detoxification pathways in T. thermophila to resist oxidative damage and cell alleviated low concentration DHA stress by regulating antioxidant enzyme. This study provides good practice on pharmacological mechanism of artemisinin-based drugs in antiparasitic.
... The genetic factors that play a role in forming the resistance of Plasmodium falciparum to artemisinin are mutations in the Kelch-13 gene. [3] ...
Article
Full-text available
Artemisinin class of antimalarial drugs play an important role in controlling falciparum malaria after the emergence of resistance of Plasmodium falciparum to other antimalarial drugs such as chloroquine, sulfadoxine-pyrimethamine and mefloquine. Therefore, the presence of Plasmodium falciparum resistance to this class of drugs is threat to global efforts to eliminate this disease. Resistance of Plasmodium falciparum to artemisinin recently known to be associated with mutations in the propeller domain of the kelch-13 (K13) Plasmodium falciparum gene. The incidence of Plasmodium falciparum resistance due to mutations in the K13 gene, among others, can be found in Cambodia, Laos, Vietnam, China, Myanmar, Thailand and Africa. The presence of mutations in this gene will change the response of Plasmodium falciparum against oxidative stress induced by artemisinin by involving the proteasome-ubiquitin pathway. In addition, mutation K13 will also change the levels of PI3K and PI3P in the body of Plasmodium falciparum. PI3K and PI3P are lipids that essential for the development of Plasmodium falciparum from ring stage to schizont. Resistance to artemisinin will also provide phenotypic changes in the life cycle of Plasmodium falciparum in the form of elongation at the stage ring and transient shortening in trophozoite development. This resistance incident can be overcome, among others by prolonging the duration of treatment (from a 3-day regimen to a 4-day regimen) and combining artemisinin with proteasome inhibitors.
... It is documented that highly saline concentrations limit plant life events through osmotic stress, ion toxicity, and secondary oxidative stress by reducing water absorption and inducing the accumulation of ions and reactive oxygen species (ROS) . However, plants have evolved a series of adaptive mechanisms in response to salt stress effects by regulating gene expression and metabolite synthesis (Pandey and Pandeyrai, 2016). Some key osmoprotectants, such as amino acids, organic acids, and sugars, are specifically synthesized under salt stress and function in regulating cellular osmotic potential . ...
Article
Full-text available
Salt stress is a major environmental factor that seriously restricts quinoa seed germination. However, the key regulatory mechanisms underlying the effect of salt stress on the initial imbibition stage of quinoa seeds are unclear. In this study, dry seeds (0 h) and imbibed (8 h) seeds with 450 mM NaCl (artificial salt) and 100% brackish water of Yellow River Estuary (BW, natural salt) were used to assess the key salt responses based on germination, transcriptome, and metabolome analyses. The results indicated that the capacity of germinating seeds to withstand these two salt stresses was similar due to the similarities in the germination percentage, germination index, mean germination time, and germination phenotypes. Combined omics analyses revealed that the common and unique pathways were induced by NaCl and BW. Starch and sucrose metabolism were the only commonly enriched pathways in which the genes were significantly changed. Additionally, amino sugar and nucleotide sugar metabolism, and ascorbate and aldarate metabolism were preferably enriched in the NaCl group. However, glutathione metabolism tended to enrich in the BW group where glutathione peroxidase, peroxiredoxin 6, and glutathione S-transferase were significantly regulated. These findings suggest that the candidates involved in carbohydrate metabolism and antioxidant defense can regulate the salt responses of seed initial imbibition, which provide valuable insights into the molecular mechanisms underlying the effect of artificial and natural salt stresses.
... Further, with the help of these structural genetic markers, it would be interesting to screen genetically high-AN-yielding lines quickly. Screening of high AN producing varieties using sequence characterized ampli ed region (SCAR) marker has also been reported in A. annua (Pandey and Pandey-Rai 2016). Developed SCAR markers could be considered as complementary tools in identi cation of high-AN producer lines, and their use could greatly reduce the time and labor needed for screening, land usage, and other costs associated with the breeding of A. annua for high AN. ...
Chapter
Chronic inflammation is involved in many pathological conditions such as rheumatoid arthritis, atherosclerosis, type 2 diabetes, asthma, obesity, inflammatory bowel diseases, neurodegenerative diseases, and cancer. Currently, there is a demand for effective anti-inflammatory drugs with low toxicity and reduced side effects. In this context, A. annua appears to be a promising source of anti-inflammatory compounds, as numerous studies on different models of inflammation certify. Further research is needed in order to understand the underlying mechanism of action for each compound. The anti-inflammatory activity of extracts is also promising, considering the fact they have been demonstrated to be more efficient than isolated compounds, although they raise the problem of reproducibility and standardization. In addition, no clinical trials were yet undertaken to assert the anti-inflammatory effect of A. annua in the human body.
Article
Full-text available
Medicinal plants remain a valuable source for natural drug bioprospecting owing to their multi-target spectrum. However, their use as raw materials for novel drug synthesis has been greatly limited by unsustainable harvesting leading to decimation of their wild populations coupled with inherent low concentrations of constituent secondary metabolites per unit mass. Thus, adding value to the medicinal plants research dynamics calls for adequate attention. In light of this, medicinal plants harbour endophytes which are believed to be contributing towards the host plant survival and bioactive metabolites through series of physiological interference. Stimulating secondary metabolite production in medicinal plants by using endophytes as plant growth regulators has been demonstrated to be one of the most effective methods for increasing metabolite syntheses. Use of endophytes as plant growth promotors could help to ensure continuous supply of medicinal plants, and mitigate issues with fear of extinction. Endophytes minimize heavy metal toxicity in medicinal plants. It has been hypothesized that when medicinal plants are exposed to harsh conditions, associated endophytes are the primary signalling channels that induce defensive reactions. Endophytes go through different biochemical processes which lead to activation of defence mechanisms in the host plants. Thus, through signal transduction pathways, endophytic microorganisms influence genes involved in the generation of secondary metabolites by plant cells. Additionally, elucidating the role of gene clusters in production of secondary metabolites could expose factors associated with low secondary metabolites by medicinal plants. Promising endophyte strains can be manipulated for enhanced production of metabolites, hence, better probability of novel bioactive metabolites through strain improvement, mutagenesis, co-cultivation, and media adjustment.
Chapter
Malaria diseases are caused by Plasmodium, either P. falciparum or non-P. falciparum requires antimalarial treatment guidelines. Artemisinin-based combination (ACT) therapies are recommended as first-line therapy for uncomplicated P. falciparum and second-line therapy for non-P. falciparum. There are five ACTs approved by WHO: artesunate-amodiaquine, artemether-lumefantrine (AL), artesunate-mefloquine, artesunate-sulfadoxine-pyrimethamine, and dihydroartemisinin-piperaquine (DHA + PPQ). Quinine derivates are considered the first-line choice for non-P. falciparum infection. Intravenous artesunate is the first-line therapy for severe P. falciparum. Amodiaquine + sulfadoxine-pyrimethamine (SP) monthly is recommended for all children aged <6 years in high malaria-endemic areas. Children with severe malaria should take injectable artesunate. The intermittent preventive treatment in pregnancy (IPTp) using SP combination is recommended for all pregnant women. Antimalarial agents have a different site of action regarding on malaria parasite life cycle. The most common antimalarial agents have been clustered into four groups, including artemisinin derivatives, quinoline derivatives, an antifolate, and antimicrobial. Artemisinin, and its derivates, including dihydroartemisinin (DHA), artesunate, and artemether, has a sesquiterpene lactone group. The structure of chloroquine, amodiaquine, and mefloquine consists of quinine ring. Both artemisinin and quinine derivates activities focused on blood schizontocidal and gametocidal. The SP, as the most common antifolate, could inhibit tetrahydrofolate production of Plasmodium. Cycline is an antibiotic that interferes with Plasmodium in the sporogony, hepatic, and erythrocytic stages. This chapter discusses further the current treatment guideline; drug of choices, including mechanism of action, pharmacokinetic profiles, efficacy report, and specific notes; and ACT regimen doses recommendation. We also provide information related to first-line antimalarial in several counties that have high malaria endemic.
Article
Full-text available
Chinese herbal medicines (CHMs), with a wide range of bioactive components, are considered to be an important source for new drug discovery. However, the process to isolate and obtain those bioactive components to develop new drugs always consumes a large amount of organic solvents with high toxicity and non-biodegradability. Natural deep eutectic solvents (NADES), a new type of green and designable solvents composed of primary plant-based metabolites, have been used as eco-friendly substitutes for traditional organic solvents in various fields. Due to the advantages of easy preparation, low production cost, low toxicity, and eco-friendliness, NADES have been also applied as extraction solvents, media, and drug delivery agents in CHMs in recent years. Besides, the special properties of NADES have been contributed to elucidating the traditional processing (also named Paozhi in Chinese) theory of CHMs, especially processing with honey. In this paper, the development process, preparation, classification, and applications for NADES in CHMs have been reviewed. Prospects in the future applications and challenges have been discussed to better understand the possibilities of the new solvents in the drug development and other uses of CHMs.
Article
Full-text available
Artemisinin, a sesquiterpene lactone endoperoxide isolated from the herb Artemisia annua L. (Asteraceae), is a highly potent antimalarial compound, which is efficient against multidrug-resistant strains of Plasmodium falciparum. The promotion of artemisinin-based combination therapies (ACTs) by the WHO during the past years lead to a strong pressure on the world market of artemisinin. The scarcity of artemisinin caused a price increase that strongly renewed the interest for Artemisia annua culture at a large scale. The use of varieties with high artemisinin content is a key factor for the development of such cultures. The new hybrids recently obtained by Mediplant, with artemisinin contents nearing 2%, are being presented.
Article
Full-text available
The biological actions of artemisinin (ART), an antimalarial drug derived from Artemisia annua, remain poorly understood and controversial. Besides potent antimalarial activity, some of artemisinin derivatives (together with artemisinin, hereafter referred to as ARTs), in particular dihydroartemisinin (DHA), are also associated with anticancer and other antiparasitic activities. In this study, we used baker’s yeast Saccharomyces cerevisiae as cellular and genetic model to investigate the molecular and cellular properties of ARTs. Two clearly separable pathways exist. While all ARTs exhibit potent anti-mitochondrial actions as shown before, DHA exerts an additional strong heme-dependent, likely mitochondria-independent inhibitory action. More importantly, heme antagonizes the mitochondria-dependent cellcidal action. Indeed, when heme synthesis was inhibited, the mitochondria-dependent cellcidal action of ARTs could be dramatically strengthened, and significant yeast growth inhibition at as low as 100 nM ART, an increase of about 25 folds in sensitivity, was observed. We conclude that ARTs are endowed with two major and distinct types of properties: a potent and specific mitochondria-dependent reaction and a more general and less specific heme-mediated reaction. The competitive nature of these two actions could be explained by their shared source of the consumable ARTs, so that inhibition of the heme-mediated degradation pathway would enable more ARTs to be available for the mitochondrial action. These properties of ARTs can be used to interpret the divergent antimalarial and anticancer actions of ARTs.
Article
Full-text available
Many drugs that target transforming growth factor-β (TGFβ) signalling have been developed, some of which have reached Phase III clinical trials for a number of disease applications. Preclinical and clinical studies indicate the utility of these agents in fibrosis and oncology, particularly in augmentation of existing cancer therapies, such as radiation and chemotherapy, as well as in tumour vaccines. There are also reports of specialized applications, such as the reduction of vascular symptoms of Marfan syndrome. Here, we consider why the TGFβ signalling pathway is a drug target, the potential clinical applications of TGFβ inhibition, the issues arising with anti-TGFβ therapy and how these might be tackled using personalized approaches to dosing, monitoring of biomarkers as well as brief and/or localized drug-dosing regimens.
Article
Full-text available
Significance Evolution of malaria parasite drug resistance has thwarted efforts to control this deadly disease. Use of drug combinations has been proposed to slow that evolution. Artemisinin is a favorite drug in the global war on malaria and is frequently used in combination therapies. Here we show that using the whole plant ( Artemisia annua ) from which artemisinin is derived can overcome parasite resistance and is actually more resilient to evolution of parasite resistance; i.e., parasites take longer to evolve resistance, thus increasing the effective life span of the therapy.
Article
The abiotic stress-induced expression pattern of three genes responsible for artemisinin accumulation was preliminarily elucidated in cultured Artemisia annua plants. The semi-quantitative RT - PCR determination showed that upon exposure to chilling, heat shock or UV light, the transcription levels of amorpha-4, 11-diene synthase gene (ADS) and cytochrome P450 monooxygenase gene (CYP71AV1) were up-regulated; In contrast, under the circumstance without stress treatments, the expression levels of ADS and CYP71AV1 genes were relatively low, while the mRNA generated by the transcription of cytochrome P450 reductase gene (CPR) remained stable under pre-and post-treatments. The induced outcome by chilling stress was also supported by the measurement data from real-time fluorescent quantitative RT - PCR. Consequently, for A. annua shoots acclaimed by low temperature, the transcripts of ADS and CYP71AV1 genes were 11 folds and 7 folds higher than those of the control, but the CPR mRNA copy was almost equivalent to the control. Furthermore, a transient pre-chilling treatment led to an elevated artemisinin content up to 7.5-8.8 mg/g dry weight (DW), i.e., increased by 66.7%-95.6% as compared with the control, thereby providing an open space for enhanced artemisinin production by the environmental stresses.
Article
Artemisinin is extracted from a traditional Chinese medicinal herb Artemisia annua L., which is regarded as the most efficient drug against malaria in the world. In recent years, attention has been paid to increase the artemisinin content through transgenic methods because of the low content of artemisinin in wild plants. In this article, three functional artemisinin-related genes namely dxr, cyp71av1 and cpr, were used to genetically modify the artemisinin biosynthesis pathway in A. annua. Four independent transgenic lines of A. annua plants with overexpression of dxr and five independent transgenic A. annua lines with overexpression of both cyp71av1 and cpr were obtained and confirmed through genomic PCR. All the transgenic A. annua plants with overexpression of the dxr gene showed higher levels of artemisinin than the wild type. The content of artemisinin in the Line D24 with overexpression of dxr (1.21±0.01 mg.g-1 DW) was more than two times compared with that in the wild type (0.52±0.01 mg.g-1 DW). All the five lines of co-overexpressing cyp71av1 and cpr had an increase of artemisinin production compared with the wild-type A. annua. Line No. 16 with overexpression of cyp71av1 and cpr had the highest content of artemisinin (2.44±0.13 mg.g-1 DW) at nearly three times of the wild-type A. annua (0.91±0.02 mg.g-1 DW). Thus, the present study demonstrated that genetic modification of the upstream 2-C-methyl-D-erythritol 4-phosphate pathway or metabolic engineering of the artemisinin-specific pathway could, respectively, enhance artemisinin biosynthesis. These strategies could be applied to develop transgenic A. annua with higher levels of artemisinin.
Article
Two field experiments were carried out in cool temperate maritime latitudes in NW Tasmania (41°S) to assess whether wilting and drying Artemisia annua plants in the field after harvest had any detrimental effects on artemisinin (the source of important antimalarial drugs) or its precursor artemisinic acid. A third field experiment studied the effect of steam distillation of A. annua for its essential oil, prior to oven drying, on artemisinin and artemisinic acid. In the first two experiments whole plants were cut off at the base and left in situ for 1, 3 and 7 days (Experiment 1) and for 7, 14 and 21 days (Experiment 2). Experiment 2 included two additional treatments: (i) shade drying whole plants under ambient conditions in the field for 21 days and (ii) drying leaves, detached at harvest, for 21 days under ambient conditions inside in the dark. The effects of all of these treatments were compared with oven drying (35°C) leaves which had been detached immediately after harvest. Field drying for 1, 3 or 7 days had no adverse effect on either artemisinin or artemisinic acid in Experiment 1 and all leaf concentrations were similar to oven drying. Field drying for 7 days in Experiment 2 also gave artemisinin and artemisinic acid levels similar to oven drying. However there was a trend for sun-, shade- and dark drying for 21 days to give higher artemisinin than oven drying although artemisinic acid was unaffected. Distillation of A. annua plants for oil extraction, prior to oven drying at 35°C, resulted in nil to negligible leaf concentration of artemisinin but artemisinic acid was unaffected. Field drying may be a way of reducing the cost of antimalarial drugs and the dual production of oil and artemisinic acid is a possibility.